234
Role of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling in coastal marine sediments Yvonne Antonia Lipsewers

Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Role of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling

in coastal marine sediments

Yvonne Antonia Lipsewers

Page 2: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

THIS RESEARCH WAS FINANCIALLY SUPPORTED BY

DARWIN CENTER FOR BIOGEOSCIENCES

NIOZ – ROYAL NETHERLANDS INSTITUTE FOR SEA RESEARCH

ISBN: 978-90-6266-492-4

Cover photos: S. Rampen, L. Villanueva, M. van der Meer

Printed by Ipskamp Printing, The Netherlands

Page 3: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Role of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling in coastal marine sediments

De rol van chemolithoautotrofe micro-organismen die deelnemen in de stikstof- en zwavelcyclus in mariene

kustsedimenten (met een samenvatting in het Nederlands)

Proefschrift

ter verkrijging van de graad van doctor aan de Universiteit Utrecht op gezag van de rector magnificus, prof.dr. G.J. van der Zwaan, ingevolge het besluit van het college

voor promoties in het openbaar te verdedigen op dinsdag 5 december 2017 des ochtends te 10.30 uur

door

Yvonne Antonia Lipsewers

geboren op 18 juli 1977 te Salzkotten, Duitsland

Page 4: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Promotor: Prof. dr. ir. J.S. Sinninghe Damsté

Copromotor: Dr. L. Villanueva

Page 5: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

„Hinterher ist man immer schlauer…“

Für meine Familie

Page 6: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 7: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Table of contents

Chapter 1 – Introduction .....................................................................................................1

Chapter 2 – Seasonality and depth distribution of the abundance and activity of am-monia oxidizing microorganisms in marine coastal sediments (North Sea) ..........23

Chapter 3 – Lack of 13C-label incorporation suggests low turnover rates of thaumar-chaeal intact polar tetraether lipids in sediments from the Iceland Shelf .............53

Chapter 4 – Abundance and diversity of denitrifying and anammox bacteria in season-ally hypoxic and sulfidic sediments of the saline Lake Grevelingen .......................79

Chapter 5 – Impact of seasonal hypoxia on activity and community structure of chem-olithoautotrophic bacteria in a coastal sediment .......................................................107

Chapter 6 – Potential recycling of thaumarchaeotal lipids by DPANN archaea in sea-sonally hypoxic surface marine sediments ...................................................................147

Chapter 7 – Synthesis ...........................................................................................................167

Summary ..................................................................................................................................187

Samenvatting ...........................................................................................................................191

References ...............................................................................................................................195

Acknowledgements ...............................................................................................................223

Page 8: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 9: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

1

1. Introduction

Marine sediments are dynamic habitats shaped by interactions of biotic and abiotic processes including the redox reactions performed by microorganisms to gain energy for growth (Schrenk et al., 2010). The biogeochemical cycling of carbon-, nitrogen- and sulfur compounds in marine sediments is tightly coupled in space and time. Due to electron donor and acceptor availability, an overlap of the activity of metabolically diverse microorganisms involved in carbon-, nitrogen- and sulfur cycling occurs.

Aerobic mineralization processes of organic matter reaching the sediment surface are mainly mediated by the metabolism of phototrophic and chemoorganoheterotrophic microorganisms. These processes result in the consumption of oxygen (O2) within the first few millimeters/cen-timeters of marine sediments. Underneath the O2 penetration depth, anaerobic mineralization processes occur due to the vertical distribution of electron acceptors in the sediment: O2 → ni-trate (NO3

−) → manganese (IV) oxide (MnO2) → ferric iron (Fe3+) → sulfate (SO42−) → carbon

dioxide/bicarbonate (CO2/HCO3−) reflecting the gradual decrease of the redox potential. Most

microorganisms involved in anaerobic mineralization processes grow chemoorganoheterotro-phically, such as most denitrifying bacteria, manganese- or iron reducing bacteria, most sulfate reducing bacteria, fermenting bacteria, methane oxidizing bacteria and archaea. As a result, re-duced inorganic compounds, such as ammonium (NH4

+), nitrite (NO2−), NO3

−, ferrous iron (Fe2+) and reduced sulfur species, such as sulfide (H2S), elemental sulfur (S0), iron sulfide (FeS), thiosulfate (S2O3

2−) and sulfite (SO32−), accumulate in the sediment (Jørgensen, 2006; Herbert,

1999), which can subsequently be reoxidized by chemolithoautotrophic microorganisms. Chemolithoautotrophic microorganisms (bacteria and archaea) gain energy for growth using

chemical compounds as energy source (chemo-) in contrast to phototrophic organisms using light. Chemolithoautotrophs are able to use reduced inorganic compounds (litho-) as electron donor and inorganic carbon, CO2 and/or HCO3

−, as their only carbon source (-autotroph). The electron donors used by chemolithoautotrophic microorganisms comprise reduced nitrogen and sulfur compounds, as well as Fe2+, H2 and carbon monoxide (CO). Other microorganisms are able to assimilate organic carbon, while using inorganic electron donors, which is termed chemolithoheterotrophy, whereas those able to use inorganic or organic compounds as carbon source while utilizing inorganic electron donors are called mixotrophs. Many microorganisms use organic electron donors and fix organic carbon in a process known as chemoorganoheter-otrophy. Most chemolithoautotrophs are facultative chemolithoautotrophs, i.e. are able to grow chemolithoautotrophically as well as using a different metabolism. As chemolithoautotrophs are not able to synthesize adenosine triphosphate (ATP) from nicotinamide adenine dinucleotide (NADH) oxidation, most obligate chemolitoautotrophs couple the oxidation of their electron donors to the reduction of quinone or cytochromes supplying reducing equivalents for carbon fixation and biosynthesis through reverse electron transport (Kim and Gadd, 2008). Many per-form aerobic respiration using O2, while anaerobic chemolithoautotrophs use NO3

−, SO42−, CO2

or metal oxides as electron acceptors (Jørgensen, 2006). Chemolithoautotrophic microorganisms are metabolically versatile and able to use a broad

variety of electron donors for inorganic oxidation in coastal marine sediments (Table 1). Aer-obic and anaerobic ammonia oxidizing bacteria, sulfide oxidizing and sulfate reducing bacteria are believed to be the main contributors to chemolithoautotrophic carbon fixation in coastal

Page 10: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

2

Tabl

e 1:

Ver

satil

ity o

f ch

emol

ithoa

utot

roph

ic re

oxid

atio

n re

actio

ns/p

roce

sses

per

form

ed b

y m

icro

orga

nism

s in

coas

tal m

arin

e se

dim

ents

(mod

ified

fr

om Jø

rgen

sen,

200

6).

Ele

ctro

n do

natin

g re

actio

nPr

oces

sC

hem

olith

oaut

otro

phic

mic

roor

gani

sms

(e.g

.)

NH

4+ +

3 /2 O

2 →

NO

2− +

2H

+ +

H2O

amm

onia

oxi

datio

nam

mon

ia o

xidi

zing

bac

teria

(AO

B) (N

itros

omon

as),

amm

onia

oxi

dizi

ng a

rcha

ea (A

OA

) (N

itros

opum

ilus)

NO

2− +

½ O

2 → N

O3−

nitri

te o

xida

tion

nitri

fyin

g ba

cter

ia (T

hiob

acill

us, N

itrob

acter

)

NH

4+ +

NO

2− →

N2 +

2H

2Oan

aero

bic

amm

onia

oxi

datio

nan

amm

ox b

acte

ria (S

calin

dua

spp.

)

H2S

+ 2

O2 →

SO

42− +

2H

+

(oth

er p

ossib

le d

onor

s: SO

32−, S

2O32−

, S0 )

sulfu

r oxi

datio

nco

lorle

ss su

lfur b

acte

ria (T

hiob

acill

us, B

eggia

toa)

H2S

+ N

O3−

+H

+ +

H2O

→ S

O42−

+ N

H4+

(oth

er p

ossib

le d

onor

s: S0 ,

S 2O32−

)

sulfi

de o

xida

tion

coup

led

to

nitra

te re

duct

ion

to a

mm

onia

(D

NRA

)

deni

trify

ing

sulfu

r bac

teria

(Thi

obac

illus

den

itrifi

cans

),

nitra

te re

duci

ng su

lfur b

acte

ria (T

hiom

arga

rita)

,

larg

e su

lfur o

xidi

zing

bac

teria

(Beg

giatoa

)

4Fe2+

+ O

2 → 4

Fe3+

+ 2

H2O

(or M

n2+)

iron

and

man

gane

se o

xida

tion

iron

bact

eria

(Ferr

obac

illus

, She

wane

lla)

2H2 +

O2 →

2H

2Ohy

drog

en o

xida

tion

hydr

ogen

oxi

dizi

ng b

acte

ria (H

ydro

genob

acter

th

ermop

hilu

s)

4H2 +

SO

42– +

2H

+ →

4H

2O +

H2S

hydr

ogen

oxi

datio

n co

uple

d to

su

lfate

redu

ctio

nso

me

sulfa

te re

duci

ng b

acte

ria (D

esulfo

vibrio

spp.)

H2 +

CO

2 → C

H4

met

hano

gene

sism

etha

noge

nic

arch

aea

4H2 +

2C

O2 →

CH

3CO

OH

+ 2

H2O

acet

ogen

esis

acet

ogen

ic b

acte

ria

Page 11: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

3

Introduction

marine sediments. In addition, many archaea are chemolithoautotrophs and representatives of the phylum Thaumarchaeota (formerly known as ‘marine group I Crenarchaeota’) have been determined to oxidize ammonia to gain energy for inorganic carbon fixation (Wuchter et al., 2003 and 2006; Könneke 2005, Brochier-Armanet et al., 2008) suggesting the contribution of archaea to the reoxidation of reduced inorganic compounds and inorganic carbon fixation in coastal marine sediments.

Chemolithoautotrophy in coastal sediments is generally considered to be insignificant in terms of biogeochemistry, due to the low growth yields of these microorganisms and the competition with photo- and heterotrophic processes, which are the main source of labile organic matter in coastal sediments (Jørgensen and Nelson, 2004). However, although chemolithoautotrophic microorganisms might play a minor role in terms of growth, they are able to reoxidize important amounts of reduced inorganic substrates (Cypionka, 2006). The dependency of chemolitho-autotrophic microorganisms on the availability of reduced inorganic compounds produced by chemoorganoheterotrophic microorganisms suggests a tight coupling between different micro-bial metabolic pathways in marine sediments. Unfortunately, direct measurements of chemo-lithoautotrophic activity are rare. Two rate measurements of dark CO2 fixation in shallow sub-tidal sediments of the Baltic Sea suggest that up to 20% of the inorganic carbon produced by mineralization appears to be reoxidized by chemoautotrophic microorganisms (Enoksson and Samuelsson, 1987; Thomsen and Kristensen, 1997). Only recently, chemolithoautotrophic sulfur oxidizing Gammaproteobacteria have been shown to contribute substantially to CO2 fixation by combined fluorescence in situ hybridization and microautoradiography (MAR FISH) analysis in the oxygen-, nitrate- and sulfide transition zone (0−3 cm) of intertidal sandy sediments (Lenk et al., 2011). Inorganic carbon fixation in coastal and ocean margin sediments, characterized by high organic carbon loadings, has been estimated to account for about 48% of the total oceanic chemoautotrophic carbon fixation rate (Middelburg, 2011). Considering the combination of high respiration rates and dominantly anoxic conditions, coastal and ocean margin sediments are presumably a hot spot for chemolithoautotrophy. The activity of chemolithoautotrophic microorganisms in coastal sediments has global implications for the cycling of carbon and the availability of nitrogen and sulfur compounds in marine environments. However, the diversity of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well as the temporal and spatial distribution of their abundance and ac-tivity, is still poorly understood, especially in subsurface sediment layers. In the upper marine sediments, dark CO2 fixation is believed to be predominantly driven by oxidation of reduced nitrogen and sulfur compounds.

1.1. Chemolithoautotrophic contribution to nitrogen cycling in marine coastal sedimentsInorganic nitrogen concentrations (NH4

+, NO3−) in marine sediments are determined by the

sedimentation and decomposition rate of organic matter, as well as by diffusive transport rates and bioturbation. The cycling of nitrogen in coastal marine sediments consists of a series re-dox reaction performed by phototrophic, heterotrophic and autotrophic microorganisms and underlies complex regulatory mechanisms, comprising physicochemical and biological factors (Herbert, 1999; Joye and Anderson, 2006). The complexity of the cycle is revealed in Fig. 1.

Nitrogen fixation, e.g. the conversion of molecular nitrogen (N2) to bioavailable organic ni-

Page 12: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

4

Figu

re 1

: Nitr

ogen

cyc

ling

in c

oast

al m

arin

e se

dim

ents

(sim

plifi

ed a

nd m

odifi

ed fr

om H

erbe

rt, 1

999;

Joye

and

And

erso

n, 2

006)

. Das

hed

lines

indi

cate

di

ffus

ive,

adve

ctiv

e, or

sur

face

exc

hang

e be

twee

n co

mpa

rtm

ents

; sol

id li

nes

repr

esen

t mic

robi

al m

edia

ted

proc

esse

s. Bo

ld p

roce

ss in

dica

tes

chem

olith

o-au

totro

phic

con

tribu

tion.

Page 13: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

5

Introduction

trogen compounds (Norg, e.g. proteins), is the initial step of the nitrogen cycling in marine sedi-ments. During the mineralization of organic nitrogenous matter ammonification, the release of NH4

+ occurs. The NH4+ released by ammonification is available for primary production and for

ammonia oxidation in the O2 transition zone of the sediment (Herbert, 1999). The fate of NH4+

regenerated in sediments may either be the adsorption to particles, the flux from the sediment to the overlying water, the consumption by biological processes within the sediment or at the sediment–water interface, the oxidation to NO2

− and then NO3− by nitrifying microorganisms,

or the transformation to molecular nitrogen (N2) via the denitrification and anammox process (Joye and Anderson, 2006).

1.1.1. Nitrification Nitrification is a two-step process, first ammonium is oxidized to NO2

− by chemolithoautotro-phic ammonium oxidizing bacteria (AOB) and archaea (AOA), which is then further oxidized to NO3

− by nitrite oxidizing bacteria (NOB) (Kim and Gadd, 2008). Aerobic ammonium ox-idation is a two-step process, in which ammonium is oxidized to hydroxylamine catalyzed by the enzyme ammonia monooxygenase (AMO). In bacteria, hydroxylamine is further oxidized to NO2

− by the enzyme hydroxylamine oxidoreductase (HAO), whereas in archaea, the oxida-tion of hydroxylamine to NO2

− may be catalyzed via nitroxyl oxidoreductase (NxOR) (Schlep-er and Nichol, 2010). In coastal marine sediments, AOB belong to the Gammaproteobacteria (Nitrosococcus oceanus) and to Betaproteobacteria (Nitrosomonas spp. and Nitrosospira spp.), based on the phylogenetic analysis of 16S rRNA genes (Herbert, 1999). Ammonia oxidizing archaea (AOA) belong to the distinct phylum of Thaumarchaeota (Brochier-Amanet et al., 2008). Most AOBs are obligate chemolithoautotrophs, whereas Thaumarchaeota can grow chemolithoau-totrophically (e.g. “Candidatus Nitrosopumilus” maritimus; Könneke et al., 2005) and possibly mixotrophically (e.g. “Nitrososphaera viennensis”; Tourna et al., 2011), besides heterotrophic growth cannot be ruled out (Pester, 2011). The communities of AOB and AOA seem to be favored by different physicochemical parameters, such as salinity, ammonia and O2 concentrations (Mosier and Francis, 2008; Santoro et al., 2008), with AOA preferring lower salinities and lower ammonia concentrations. The oxidation of NH4

+ plays an important role in generating nitrogen oxides (NO2

− and NO3−) for microorganisms utilizing these to gain energy for carbon fixation, such as

anammox and denitrifying bacteria.

1.1.2. Anammox Bacteria of the phylum Planctomycetes have been discovered to perform anaerobic ammonia oxidation (anammox) reaction, which consist on conversion of NH4

+ and NO2− into N2 while

growing chemolithoautotrophically (Strous et al., 1999, Kuypers et al., 2003). In addition to NO2

−, anammox bacteria are also able to use Fe3+, manganese oxides and NO3− as electron

acceptors in their metabolism (Strous et al., 2006). All five described anammox genera belong to the order Brocadiales and are related to the order Planctomycetales (Jetten et al., 2009), but at present only “Candidatus Scalindua” has been detected in marine sediments and O2 minimum zones (OMZs) (Daalsgard et al., 2005; Woebken et al., 2008; Lam et al., 2009). The contribution of the anammox process to the total nitrogen removal in marine sediments has been estimated to be between 0–80% (Dalsgaard, 2005).

Page 14: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

6

1.1.3. Denitrification After the consumption of dissolved O2 in the sediment pore water, organic matter decompo-sition continues via denitrification, anammox or dissimilatory nitrate reduction to ammonium (DNRA) (Zumft, 1997; Seitzinger, 2006). Whereas DNRA results in bioavailable ammonia, de-nitrification leads to the release of gaseous end products such as nitric oxide (NO), nitrous oxide (N2O), and dinitrogen gas (N2) to the atmosphere. Dissimilatory denitrification, the microbial oxidation of organic matter with NO3

− or NO2− as terminal electron acceptor, is considered to

be the sequence of nitrate respiration, nitrite respiration combined with NO reduction, and N2O respiration: NO3 → NO2 → NO → N2O → N2. The denitrification process is performed by facultative chemoorganoheterotrophic bacteria, some archaea and foraminifera (Zumft, 1997; Risgaard-Petersen et al., 2006), and occurs as well in chemolithoautotrophic organisms perform-ing the oxidation of an inorganic compound such as H2S coupled to nitrogen oxide reduction. Recently, sulfide-dependent denitrification, performed by autotrophic members of Alpha-, Beta- Gamma-, and Epsilonproteobacteria (especially Thiobacillus denitrificans and Sulfurimonas denitrifi-cans), has been suggested to play a major role in the denitrification process in the oxygen- and sulfide- transition zone of coastal marine sediments (Shao, 2010).

The processes of nitrification and denitrification in sediments are the major source of NO3−

for denitrification in some estuaries and shelf regions. In systems with low NO3− concentrations

(<10 μM) in the oxygenated bottom water, coupled nitrification/denitrification accounts for 90% of the NO3

− that is required for denitrification (Risgaard-Petersen, 2003). However, cou-pled nitrification/denitrification is a mechanism that can be responsible for substantial nitrogen removal from the sediment, which is in turn not available for primary producers. Nitrogen re-moval by denitrification, including anammox, might control the rate of eutrophication in coastal marine systems that receive large amounts of anthropogenic nitrogen (Herbert, 1999).

1.2. Chemolithoautotrophic sulfur cycling in coastal marine sediments Below the oxygen penetration depth (OPD) and after the depletion of electron acceptors, such as oxygen, NO3

−/NO2−, and metal oxides, sulfate reduction to H2S becomes the prevalent mi-

crobial redox process in marine sediments (Jørgensen, 2006). Sulfate reduction, the reduction of elemental sulfur (S0) and sulfur disproportionation can result in H2S production and upward diffusion. Most H2S is reoxidized by chemical and microbial mediated sulfur oxidation or buried as iron sulfide and pyrite (FeS, FeS2).

1.2.1. Sulfate reducing microorganisms Sulfate reducing microorganisms have a key role in the sulfur cycle, they use SO4

2– as a terminal electron acceptor in the degradation of organic matter, which results in the production of H2S (Muyzer and Stams, 2008), which subsequently can be oxidized aerobically and anaerobically by sulfur oxidizing bacteria (e.g. Thiobacillus or Beggiatoa spp.). During dissimilatory sulfate reduc-tion, SO4

2− is activated by the enzyme ATP sulfurylase, catalyzing the formation of adenosine phosphosulfate (APS) from SO4

2− and phosphate (ATP). The SO42− in APS is directly reduced

to SO32− by the cytoplasmic enzyme adenosine-5’ phosphosulfate (APS) reductase, which is then

further reduced to H2S by the enzyme sulfite reductase. Sulfate reducing microorganisms are metabolically and genetically divers, some grow chemoorganoheterotrophically by using acetate, propionate, butyrate or lactate as electron acceptor, whereas many chemolithoautotrophic repre-

Page 15: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

7

Introduction

sentatives use hydrogen as electron donor and SO42− as electron acceptor while utilizing CO2 as

carbon source (e.g. Desulfovibrio spp.). Sulfate reducing bacteria belong to the class of Deltapro-teobacteria, as well as to the phyla of the Firmicutes and Nitrospirae. Within the archaea, sulfate reducing representatives belong to the genus Archaeglobus within the phylum of Euryarchaetoa, and to the phylum of Crenarchaeota (Muyzer and Stams, 2008). The activity of sulfate reducing microorganisms is especially important in organic-rich sediments underlying highly productive waters of continental shelves and slopes (Jørgensen, 1982a, b; Jørgensen and Kasten, 2006). It has been estimated that sulfate reduction can account for more than 50% of the organic carbon mineralization in marine sediments (Jørgensen, 1982a, b).

1.2.2. Sulfur oxidizing microorganisms in marine sediments Sulfur oxidation, i.e. the oxidation of reduced sulfur compounds, such as H2S, SO3

2−, S2O32−,

S0, to SO42− can be performed by taxonomical divers chemolithoautotrophic bacteria and some

archaea, using mainly O2 and NO3− as electron acceptor, which ultimately links the carbon-,

nitrogen- and sulfur cycling in marine sediments. Sulfide oxidation starts with the conversion of H2S (sulfur, thiosulfate) to sulfite. Some sulfur oxidizing chemolithoautotrophs use the enzyme sulfite oxidase, which transfers electrons directly to cytochrome c generating ATP during elec-tron transport and proton motive force formation. Others use the enzyme adenosine-5’ phos-phosulfate (APS) reductase, which is an essential enzyme of sulfate reducing bacteria, which results in SO4

2− production in sulfur oxidizers (Meyer and Kuever, 2007b). Chemolithoautotrophic bacteria belonging to the class of Betaproteobacteria (Thiobacillus),

Gammaproteobacteria (Thiomicrosprira), and to the Epsilonproteobacteria (Acrobacter, Sulfuri-monas, Sulfurovum), have been shown to contribute to chemolithoautotrophic sulfide oxidation in coastal marine sediments (Jørgensen and Nelson, 2004; Campbell et al., 2006; Lenk et al., 2011). Chemolithoautotrophic sulfur oxidizing bacteria are taxonomically broad, while chemolithoau-totrophic sulfur oxidizing archaea seem to be restricted to Sulfolobales (Crenarchaeota) using metal oxides, Thermoplasmatales (Euryarchaeota) using elemental sulfur, and only two genera seem to be able to oxidize H2S, i.e. Acidianus sulfidivorans and Ferroglobus placidus, using NO3

− and Fe3+ as electron acceptors (Hafenbradl et al., 1996; Plumb et al., 2007).

Additionally, large sulfur bacteria, Gammaproteobacteria of the order Thiotrichales (Beggiatoa, Thioploca, Thiomargarita), contribute substantially to sulfide oxidation in coastal marine sediments. Under low O2 conditions, Beggiatoaceae are able to store NO3

− in an internal vacuole at the sediment surface, before gliding into anoxic and sulfidic layers, where they oxidize H2S to S0 with the stored NO3

−. Here, S0 is stored in globules and transported back to the surface to complete the oxidation to sulfate. Most members of the marine Beggiatoaceae family are chemolithoau-totrophs, some species are able to use an organic carbon source for mixotrophic or chemooar-ganoheterotrophic growth (Mußmann et al., 2003; Jørgensen and Nelson, 2004).

1.3. Autotrophic CO2 fixation pathways of microorganisms Up to date six different CO2 fixation pathways used by taxonomically diverse bacteria and ar-chaea have been discovered (Fig. 2). In the Calvin-Benson-Bassham (CBB) cycle, CO2 reacts with ribulose 1,5-bisphosphate to gain two 3-phosphoglycerate, from which the sugar is formed. Apart from eukaryotes such as plants and algae, the CBB cycle is utilized by cyanobacteria, many aerobic or facultative aerobic Proteobacteria belonging to the Alpha-, Beta-, and Gamma- classes

Page 16: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

8

(including AOB), Sulfobacillus spp., iron and sulfur oxidizing Firmicutes, some mycobacteria, and bacteria of the phylum Chloroflexi (Oscillochloris) (Berg et al., 2010). The CBB cycle prevails un-der oxic conditions, whereas the more energy efficient and less O2 sensitive reductive tricarbox-ylic acid (rTCA) cycle is restricted to microaerobic and anaerobic conditions (Berg et al., 2010). The rTCA cycle reverses the reactions of the oxidative citric acid (Krebs-) cycle and operates in sulfate reducing Delta- and sulfur oxidizing Epsilonproteobacteria, green sulfur bacteria and in microaerophilic bacteria of the phylum Aquificae (Hügler et al., 2005; Campbell et al., 2006).

The strictly anaerobic reductive acetyl-coenzyme A (acetyl-CoA) pathway is a noncyclic path-way, in which one CO2 molecule is reduced to CO and one to a methyl group, from which acetyl-CoA is synthesized subsequently. The acetyl-CoA pathway functions in acetogenic and methanogenic bacteria, as well as in anammox bacteria (Planctomycetes), sulfate reducing Del-taproteobacteria (Desulfobacterium sp.), and in Euryarchaeota (Archaeglobales). The 3-hydroxy-propionate bicycle is currently restricted to members of the Chloroflexacae family. The 3-hydroxy-propionate/4-hydroxybutyrate cycle occurs in aerobic Crenarchaeota (Sulfolobales) and marine Thaumarchaeota (AOA), whereas the dicarboxylate/4-hydroxybutyrate cycle functions in anaer-obic Crenarchaeota of the orders Thermoproteales and Desulfurococcales. The prevalence of CO2 fixation pathways utilized by chemolithoautotrophic microorganisms in seasonally hypoxic coastal marine sediments is poorly known at present (Berg et al., 2010, and 2011).

Thaumarchaeota

Figure 2: Phylogenetic classification of autotrophic bacteria and archaea; green: Calvin-Benson-Bassham cycle, red: reductive tricarboxylic acid cycle, blue: reductive acetyl-CoA cycle, orange: 3-hydroxypropionate bicycle, purple: 3-hydroxypropionate-/ 4-hydroxybutyrate cycle, brown: dicarboxylate/4-hydroxybutyrate cycle (Figure modified from Hügler and Sievert, 2011).

Page 17: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

9

Introduction

1.4. Physicochemical parameters affecting microbial chemolithoautotrophy in coastal marine sedimentsVarious abiotic and biotic factors control the abundance and activity of chemolithoauto trophic microorganisms involved in carbon-, nitrogen-, and sulfur cycling in marine sediments. Gener-ally, O2 is one of the most important ecological factors influencing chemolithoautotrophy, espe-cially in the oxygen transition zones of coastal marine sediments, where O2 is depleted within the first millimeters. The O2 concentration shapes the distribution of different autotrophic carbon fixation pathways, as well as the availability of suitable electron donors and acceptors, metals and coenzymes (Berg, 2011). Whereas nitrification is predominantly an obligate aerobic process, denitrification is a facultative anaerobic process, most denitrifiers prefer to respire O2 instead of NO3

− when O2 concentrations exceed 20 mM (Tiedje, 1982). Anammox bacteria are obligate anaerobes and despite their O2 tolerance, the activity seems to be limited to anoxic conditions (Dalsgaard et al., 2005; Risgaard-Petersen et al., 2004, and 2005).

Coastal regions, especially estuaries and coastal areas with limited water exchange, often ob-tain large amounts of anthropogenic nitrogen leading to eutrophication, (Howarth and Marino, 2006). Hypoxia is one of the common features of eutrophication and the phenomenon of seasonal hypoxia has expanded over the last decades in coastal regions. Seasonal hypoxia causes a decrease in the availability of electron acceptors, i.e. O2 and NO3

− in the bottom water (Rosen-berg and Diaz, 2008), suggesting consequences for chemolithoautotrophic microorganisms in-volved in carbon-, nitrogen-, and sulfur cycling in marine coastal sediments and thus for the biogeochemistry in marine sediments (Middleburg and Levin, 2009).

Extensive sulfate reduction leads to the accumulation of free H2S in anoxic sediments, which is diffusing upwards (Jørgensen and Nelson, 2004). Sulfide has been shown to inhibit ammonia oxidation (Joye and Hollibaugh, 1995), as well as enzymes catalyzing the last steps of denitrifica-tion (Sørensen, 1980; Porubsky et al., 2009). The effect of H2S on anammox is relatively unclear, but there is certain evidence, that the activity of anammox bacteria is inhibited in sulfidic waters of oxygen minimum zones (Daalsgaard 2003; Jensen et al., 2008), suggesting that the abundance and activity of anammox bacteria might as well be inhibited by rising H2S concentrations in coastal marine sediments.

However, it is well known that high concentrations of free H2S are toxic for benthic animals, such as bioturbating organisms (Vaquer-Sunyer and Duarte, 2008). The key role of bioturbation in nutrient cycling is well established (Meysman et al., 2006) and has been suggested to extend the area of O2 availability in deeper sediment layers, supporting aerobic microbial processes, such as nitrification, which in turn might fuel anammox and denitrification (Meyer et al., 2005). Recently, bioturbation has been shown to induce changes in the sedimentary microbial commu-nity composition, resulting in distinct communities within the burrow (Laverock et al., 2010).

1.5. Methods to study microbial chemolithoautotrophyThe majority of microorganisms that occur in natural environments cannot be cultured or iso-lated using traditional cultivation techniques, which limits our knowledge of environmental mi-crobial diversity (Rappé et al., 2003). In the last decades, a wide range of cultivation independent approaches have been developed to understand the abundance, diversity and metabolic activity of microbial communities in the environment using a wide range of organic and inorganic

Page 18: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

10

biomarker molecules, such as DNA/RNA and lipids, as well as the incorporation of labeled substrates into biomarker molecules.

1.5.1. Abundance and diversity of chemolithoautotrophic microorganisms using DNA-based methods The analyses of DNA sequences of microorganisms has expanded our knowledge of the tax-onomic and functional diversity of marine microbes to a large extent. Molecular methods re-vealed a diverse community of microorganisms with diverse metabolic capabilities, new species and novel pathways of organic matter cycling (Delong, 1992; Fuhrman et al., 1992). Advances in molecular microbiology allow for DNA sequence analysis and comparison to sequences that are already available in sequence databases. The development of quantitative PCR (qPCR) for the real time quantification of selected DNA sequences (genes) can be used to estimate the abundance of specific groups of microorganisms. In general, two kinds of genes are normally targeted, the 16S rRNA gene and functional genes.

The 16S ribosomal RNA is part of the 30S small subunit of a prokaryotic ribosome, and the 16S rRNA gene sequence has been used to reconstruct the microbial phylogeny (Woese and Fox, 1977). The 16S rRNA gene sequence is especially useful because as it is present in all Domains of life, i.e. Bacteria, Archaea and Eukarya. Variable regions of the 16S rRNA gene are diagnostic for specific groups of organisms, whereas highly conserved regions can be used to assess the diversity of the microbial community. Sequencing of the highly conserved and species specific 16S rRNA gene has been extensively used to characterize the microbial community in marine surface and subsurface sediments (Madsen, 2008). Nevertheless, the rapid development of new generation sequencing methods allows for high throughput sequencing, and 16S rRNA gene am-plicon sequencing has been shown to be a powerful tool to study the diversity of the microbial community as well as the relative abundance of specific groups in a high resolution in the marine environment (Gilbert and Dupont, 2011; Klindworth et al., 2013; Cruaud et al., 2014).

As the 16S rRNA gene gives only limited and indirect information about the metabolism, i.e. the contribution to CO2 fixation and reoxidation pathways, functional gene analysis, i.e. genes encoding for key-enzymes involved in different CO2 fixation and reoxidation pathways, is a powerful tool to study the diversity and abundance of chemolithoautotrophic microorganisms. Gene sequences, i.e. the cbbL gene coding for ribulose 1,5-bisphosphate carboxylase/oxygenase (RuBisCO) and the aclB gene coding for the ATP-citrate lyase, functioning in the CBB and rTCA cycle respectively, have been used to assess the diversity of chemolithoautotrophic microorgan-isms in intertidal mudflat sediments as well as in hydrothermal vent sediments and in sediments with an oxygen-sulfide chemocline (Nigro and King, 2007; Campbell et al., 2003, and 2006). However, the temporal and spatial distribution of the diversity and the abundance of chemolith-oautotrophic microorganisms using the CBB or rTCA cycle have rarely been studied in coastal marine sediments. Some important key-enzymes and their corresponding functional genes for inorganic carbon assimilation and reoxidation pathways are summarized in Table 2.

The key-enzyme of aerobic ammonia oxidation is a membrane-bound putative ammonia monooxygenase (AMO), catalyzing the oxidation of ammonium to hydroxylamine in Bacteria (Rotthauwe et al., 1997) and Archaea (Venter et al., 2004; Könneke et al., 2005; Yakimov et al., 2009). The metabolic gene amoA, encoding AMO, has been used to analyze the diversity and abundance of AOA and AOB in estuarine sediments (Beman and Francis, 2006; Mosier and

Page 19: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

11

IntroductionTa

ble

2: Im

port

ant f

unct

iona

l gen

es o

f C

O2 fi

xatio

n an

d re

oxid

atio

n pa

thw

ays.

Prim

er se

quen

ces a

re re

port

ed in

the

give

n re

fere

nces

.

Path

way

Key

enz

yme

Func

tiona

l ge

neR

efer

ence

s

CO2 fixation

Cal

vin-

Bens

on-B

assh

am (C

BB) c

ycle

/re

duct

ive

pent

ose

phos

phat

e cy

cle

ribul

ose1

,5-b

ispho

spha

te c

arbo

xyla

se/

oxyg

enas

e (R

uBisC

O)

cbbL

Nig

ro a

nd K

ing,

200

7

redu

ctiv

e tri

carb

oxyl

ic a

cid

(rTC

A) c

ycle

/ A

rnon

–Buc

hana

n cy

cle

ATP

-citr

ate

lyas

e (A

CL)

aclB

Cam

pbel

l et a

l., 2

003

reoxidation

Am

mon

ia o

xida

tion

arch

aeal

am

mon

ia m

onoo

xyge

nase

(A

MO

)am

oAYa

kim

ov e

t al.

(201

1)

β-pr

oteo

bact

eria

l am

mon

ia

mon

ooxy

gena

se (A

MO

)am

oARo

tthau

we

et a

l. (1

997)

; H

orne

k et

al.

(200

6)

Ana

erob

ic a

mm

onia

oxi

datio

n (a

nam

mox

)

hydr

azin

e sy

ntha

se (H

ZS)

hzsA

Har

hang

i et a

l., 2

012

cyto

chro

me

cd1-

cont

aini

ng n

itrite

re

duct

ase

(NIR

)ni

rSLa

m e

t al.,

200

9

Den

itrifi

catio

ncy

toch

rom

e cd

1-co

ntai

ning

nitr

ite

redu

ctas

e (N

IR)

nirS

Brac

ker e

t al.,

199

8

Den

itrifi

catio

n co

uple

d to

sulfu

r oxi

datio

nad

enos

ine-

5’ p

hosp

hosu

lfate

(APS

)re

duct

ase

aprA

Mey

er &

Kue

ver,

2007

c

Page 20: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

12

Francis, 2008; Abell et al., 2010) and in sandy coastal sediments (Santoro et al., 2008). Howev-er, the spatial and temporal distribution of the diversity and abundance of microbial ammonia oxidizers, as well as of anammox bacteria, denitrifiers, sulfate reducers and sulfur oxidizers, in seasonally hypoxic sediments, is still unknown. The hzsA and nirS gene of anammox bacteria, encoding for the hydrazine synthase (HZO) and cytochrome cd1-containing nitrite reductase (NIR), have been proposed as suitable biomarker to detect anammox bacteria diversity and abundance in the marine environment (Kartal et al., 2011; Harhanghi et al., 2012; Lam et al., 2009) as well as in marine sediments (Li et al., 2011). The cytochrome cd1-containing nitrite re-ductase, coded by the nirS gene, operates as well in the dissimilatory denitrification process, cata-lyzing the reduction of NO2

− to NO, the first gaseous product in the sequence of denitrification (Zumft, 1997; Braker, 1998). The nirS gene has been used to study the diversity and to estimate the abundance of denitrifying microorganisms in estuarine sediments (Smith at al., 2007; Mosi-er and Francis, 2010). The diversity of sulfate reducing and sulfur oxidizing microorganisms have been studied using the aprA gene encoding the disssimilatory adenosine-5’ phosphosulfate (APS) reductase, which is a key enzyme in sulfate reduction and sulfur oxidation processes. The APS reductase catalyzes the conversion of APS to AMP and SO3

2− in sulfate reducing micro-organisms, and operates in the reverse direction in sulfur oxidizing microorganisms. The highly conserved aprA gene has been shown to be a suitable biomarker to profile the community struc-ture of sulfate reducing and sulfur oxidizing microorganisms in marine and estuarine sediments (Meyer and Kuever, 2007c, Muyzer and Stams, 2008).

1.5.2. Archaeal and bacterial membrane lipids as taxonomic markers Membrane lipids are considered as taxonomic markers that can be used to estimate the abun-dance, distribution, and physiological status of certain microbial groups (White, 2007). In ad-dition, these lipids can be preserved in sediments upon cell depth and thus being used as bio-markers of the presence of certain organisms both in present and past ecosystems. On the contrary, other biomolecules like DNA have a more rapid turnover and they cannot be used for this purpose. In recent years, intact polar lipids (IPLs) are increasingly applied for tracing ‘living’ bacteria and archaea in the environment (Lipp et al., 2008; Rossel et al., 2008; Lipp and Hinrichs, 2009). IPLs with polar head groups are present in living cells but upon cell lysis the polar head groups are lost, releasing the core lipids (CLs) that may be preserved in the fossil record. Since IPLs degrade relatively quickly after cell death (Harvey et al., 1986), it is possible to associate the presence of IPLs in the environment with the occurrence of their living producers (Zink et al., 2003; Biddle et al., 2006; Huguet et al., 2010; Schubotz et al., 2009). Specifically, IPLs with phos-pho- head groups have been shown to degrade rapidly after cell death, while IPLs with glycosidic polar head groups might be preserved in the sediment record (White et al., 1979; Harvey et al., 1986; Bauersachs et al., 2010; Lengger et al., 2012).

Membrane lipids from bacteria and eukaryotes are composed of fatty acid chains that are linked to the glycerol moiety through ester bonds and organized in a bilayer structure. On the other hand, archaeal membrane lipids are characterized by ether bonds, isoprene-based alkyl moieties, and an opposite stereochemistry of the glycerol phosphate backbone, i.e. sn-glycer-ol-1-phosphate (G1P). Archaeal membrane lipids are typically a variation of two main struc-tures, sn-2,3-diphytanylglycerol diether (archaeol) with phytanyl (C20) chains in a bilayer structure, and sn-2,3-dibiphytanyl diglycerol tetraether (glycerol dibiphytanyl glycerol tetraether, GDGT), in which the two glycerol moieties are connected by two C40 isoprenoid chains, allowing the

Page 21: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

13

Introduction

formation of monolayer membranes (Koga and Morii, 2007). GDGTs have been found to be ubiquitous in the marine water column and sediments (Schouten et al., 2000). Isoprenoid GDG-containing 0–8 cyclopentane moieties (Fig. 3) are commonly found in different archaeal groups, such as Thaumarchaeota (Wuchter et al., 2006), extremophiles (Huber et al., 1998; Lattuati et al., 1998; Ellen et al., 2009), methanogens (Koga et al., 1998; Gattinger et al., 2002; Zhang et al., 2011), and anaerobic methane-oxidizing archaea, ANME (Pancost et al., 2001). However, the isoprenoid GDGT crenarchaeol (Sinninghe Damsté et al., 2002a), containing four cyclopen-tane moieties and a cyclohexane moiety, is considered to be characteristic of Thaumarchaeota (Schouten et al., 2000; Sinninghe Damsté et al., 2002a; Zhang et al., 2006; Pester et al., 2011; Pitcher et al., 2011a). The culture-based GDGT compositions of the main archaeal groups, which are potentially involved in chemolithoautotrophy and sulfur metabolism in marine sedi-ments, are summarized in Table 3.

The distribution of thaumarchaeotal GDGTs in the marine environment has been shown to be affected by temperature, i.e. with increasing temperature there is an increase in the relative abundance of cyclopentane-containing GDGTs (Schouten et al., 2002; Wuchter et al., 2004, and 2005). Based on this relationship, the TEX86 paleotemperature proxy was developed and calibrated using sea surface temperature and it has been widely applied for more than a decade

Table 3: Membrane lipid composition of the main groups of archaea involved in chemolithoautotro-phy and sulfur metabolism in marine sediments (modified from Villanueva et al., 2014). Species specific GDGT occurrence and distributions of pure cultures are described in detail in Schouten et al. (2013) and references therein.

Phylum Order Metabolism Membrane lipid

Thaumarchaeota

(AOA)

Cenarchaealesammonia oxidizer

autotroph

GDGT-0 to 4,

crenarchaeol

Nitrosopumilalesammonia oxidizer

autotroph/mixotroph

GDGT-0 to 4,

crenarchaeol

Nitrosospheralesammonia oxidizer

autotroph

GDGT-0 to 4,

crenarchaeol

Crenarchaeota

Desulfurococcales sulfur dependent GDGT-0 to 4,

Sulfolobales sulfur dependentGDGT-0 to 4,

GDGT-5 to 8

Thermoproteales sulfur dependentGDGT-0 to 4,

GDGT-5 to 8

Euryarchaeota

Thermoplasmatales sulfur dependentGDGT-0 to 4,

GDGT-5 to 8

Thermococcales sulfur dependentarchaeol,

GDGT-0Archaeglobales sulfur dependent GDGT-0

ANME-clustermethane oxidation

chemoorganoautotrophGDGT-0 to 4

Page 22: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

14

(Schouten et al., 2002). However, evidences increasingly demonstrate that temperature is not the only variable affecting the GDGT distribution (Schouten et al., 2013, Pearson et al., 2013). For example, TEX86 responds to temperature in a different way in enrichments (Schouten et al., 2007a) and cultures (Pitcher et al., 2011a) than in suspended particulate matter (SPM) or when fossilized in surface sediments (Kim et al., 2010). It is also known that TEX86 does not reflect surface temperatures but rather subsurface temperatures (Huguet et al., 2007; Lopes dos Santos et al., 2010; Kim et al., 2010), and that archaeal production in the deeper water column may also contribute to the signal of GDGTs in sediments (Pearson et al., 2001; Shah et al., 2008). In addi-tion, the TEX86 signal may also be affected by input from other archaea, which could potentially synthesize GDGTs used in TEX86. For example, it has been suggested that in situ produced lipids of benthic and methanogenic archaea might contribute to the GDGT pool in marine sediments, which might be relevant for the reliability of the TEX86 proxy (Weijers et al., 2006a; Blaga et al., 2009; Sinninghe Damsté et al., 2012a). In addition, some studies have suggested that benthic archaea might be able to recycle core lipid GDGTs in marine sediments (Takano et al., 2010; Liu et al., 2011).

Recent advances in high performance liquid chromatography (HPLC)- mass spectrometry (MS) methods have allowed the analysis of core GDGTs as well as IPL GDGTs. IPL GDGTs have been analyzed by separating them from the core lipids using silica gel chromatography and hexane/EtOAc mixtures as eluent (Oba et al., 2006; Pitcher et al., 2009) followed by acid hydrolysis and high pressure liquid chromatography coupled via an atmospheric pressure chem-ical ionization interface to a mass spectrometer (HPLC−APCI−MS) (Hopmans et al., 2000; Sturt et al., 2004; Schouten et al., 2007a). With this method, IPL derived core GDGTs can be quantified using internal standards, but it does not allow to specify the kind of head group that was attached to the GDGT. In contrast, the development of HPLC combined with electrospray chemical ionization mass spectrometry (HPLC−ESI−MS), allows to analyze GDGT IPLs, with

Figure 3: Structures of isoprenoid GDGTs (0–8) and crenarchaeol (mod. from Schouten et al., 2013).

Page 23: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

15

Introduction

the polar head group still attached to the core (Rütters et al., 2002; Zink et al., 2003; Sturt et al., 2004). Core and intact polar GDGTs have been analyzed by HPLC-ESI-MS2 in selected reac-tion monitoring (SRM) (Schouten et al., 2008a), in AOA sediment enrichment cultures, includ-ing “Candidatus Nitrosoarchaeum limnia” (Pitcher et al., 2011a), as well as in marine sediments (Lengger et al., 2012). Polar headgroups of creanarchaeol have been seen to be mainly com-prised by monohexose (MH), dihexose (DH), phosphohexose (PH) and hexose-phosphohexose (HPH) (Fig. 4) (Pitcher et al., 2011a). Results suggest that HPH-crenarchaeol is the most suitable lipid biomarker to trace living Thaumarchaeota of marine group I.1a due to the labile nature of the phosphate bond (Harvey et al., 1986; Schouten, et al., 2010) in comparison with glycosidic crenarchaeol-IPL-GDGTs. Additionally, it has been shown that the HPH-crenarchaeol con-centration corresponds well to DNA-based thaumarchaeal abundance (Pitcher, 2011a and b; Schouten et al., 2012a; Lengger et al., 2012).

Bacterial membrane lipids are composed by fatty acids linked by ester bonds to the glycer-ol-3-phosphate backbone. The phospholipid-linked fatty acids (PLFAs) have been extensively used to characterize bacteria and eukaryotic microorganisms in different environmental settings including estuarine sediments (Guckert et al., 1985). PLFAs have complex structures with dif-ferent carbon lengths, unsaturations, methylations and isomers (Vestal and White, 1989) and in some cases, PLFA signatures are specific enough to pinpoint shifts in the bacterial and eukary-otic community composition (Boschker and Middleburg, 2002) (Table 4), but it does not offer a high taxonomic resolution compared to DNA/RNA-based methods. PLFA concentrations can be analyzed by gas chromatography (GC)-flame ionization detection (FID) and MS according to a method previously described (Boschker et al., 1998).

Other membrane lipids, present in microbial cells can be used as biomarkers of the presence of their producer. For the case of chemolithoautotrophic microorganisms of importance in ma-rine sediment, the specific lipids contained in anammox bacteria are a clear example. Anammox bacteria contain a membrane bound compartment in the cytoplasm, the anammoxosome, in

Figure 4: Main head groups of crenarchaeol detected in cultures of “Ca. Nitrosopumilus maritimus”, “Ca. Nitrososphaera gargensis” and “Ca. Nitrosoarchaeum limnia”, as well as in marine sediments.

Page 24: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

16

which the anammox reaction takes place (van Niftrik et al., 2008). The anammoxosome consists of ladderane lipids with three or five linearly arranged cyclobutane rings (Sinninghe Damsté et al., 2002b). Ladderane lipids can be either ester- or ether bound to glycerol and major polar headgroups are phosphatidylcholine (PC), phosphoetanolamine (PE) and phosphoglycerol (PG) (Boumann et al., 2006; Rattray et al., 2008). Whereas ladderane core lipids have been shown to be suitable biomarkers for anammox bacteria abundance in the fossil sediment record, the C20-[3]-monoether ladderane lipid attached to a PC head group has been shown to be a suitable

Table 4: PLFA signatures of the main groups of chemoautotrophic bacteria in marine sediments. The rel-ative contributions of the main fatty acids to the total PLFA composition and complete PLFA signatures are described in the listed references (mod. from Vasquez-Cardenas, 2016). For details on the nomencla-ture of PLFAs see supplementary information.

Metabolism PLFA signature Reference

Ammonium oxidizers16:1ω7c

16:0Blumer et al., 1969

Nitrite oxidizers16:1ω7c

16:018:1ω7c

Lipski et al., 2001Blumer et al.,1969

Sulfate reducers(Deltaproteobacteria)

14:015:0i15:0

16:1ω7c16:0

i17:1ω7c17:1ω6c18:1ω7c

Taylor and Parks, 1983Edlund et al., 1985Suzuki et al., 2007Pagani et al., 2011

Mat forming sulfur oxidizers(Gammaproteobacteria)

14:016:1ω7c16:1ω7t

16:018:1ω7c

Zhang et al., 2005Li et al., 2007

Sulfur oxidizers(Epsilonproteobacteria)

14:016:1ω7c

16:018:1ω7c

Inagaki et al., 2003Takai et al., 2006

Donachie et al., 2005

Iron and Sulfur oxidizers

16:1ω7c16:0

cy17:0cy19:0

Knief et al., 2003

Page 25: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

17

Introduction

biomarker for living anammox bacteria in continental shelf sediments (Jaeschke et al., 2010). Ladderane core lipids can be analyzed as fatty acid methyl esters (FAMEs) by HPLC-APCI-MS2 in selective reaction monitoring (SRM) mode as previously described (Hopmans et al., 2006) and modified to include short chain fatty acids (Rush et al., 2012). The PC-monoether ladderane lipid can be analyzed by HPLC-MS2 (Jaeschke et al., 2009), and quantified using an external PC-mon-oether ladderane standard. Main structures of ladderane core lipids and the intact PC-mono-ether ladderane lipid are summarized in Fig. 5.

1.5.3. Determining the activity of chemolithoautotrophic microorganisms The determination of active (chemolithoautotrophic) microorganisms in marine sediments has been a challenge and has been addressed using different methods by targeting the metabolic process or the activity of specific groups or specific microorganisms.

In sediments, microsensor profiling, i.e. concentration measurement of respiratory gases O2 and H2S has been used to identify active microbial processes, such as photosynthesis, respira-tion and sulfur oxidation (Nielsen et al., 1990; Schramm et al., 1996). The activity of microbes participating in nitrogen cycling processes, such as nitrification, denitrification and anammox, have been estimated by isotope tracer experiments with 15N- and 14N-labeled NO3

−, NO2− and

NH4+ by the discrimination of isotope pairing in marine sediments (Nielsen, 1992; Rysgaard et

al., 1993). Microsensor profiling, as well as isotope tracer experiments, are adequate methods to measure overall process rates, but do not provide any taxonomic information about metaboli-cally active microorganisms.

In contrast, the expression of functional genes, i.e. the quantification of mRNA by reverse transcription (Holmes et al., 2004) and subsequent qPCR, has been proposed to be a good marker to determine the activity of specific microorganisms in the marine environment. The positive gene expression of functional genes may serve as evidence for active in situ processes and provide insight into the diversity of active microorganisms (Nogales et al., 2002; Wellington

Figure 5: Chemical structures of ladderane core lipids and the intact PC-monoether ladderane lipid.

Page 26: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

18

et al., 2003). For the case of chemolithoautotrophic microorganisms, the transcript abundance of the ar-

chaeal amoA gene has been quantified to estimate the activity of Thaumarchaeota in oceanic waters, showing that marine Thaumarchaeota are actively involved in ammonia oxidation in the water column of the Pacific Ocean, as well as in the Peruvian OMZ and in the Mediterranean Sea (Lam et al., 2009; Church et al., 2010; Yakimov et al., 2011). Additionally, gene expression of the bacterial and archaeal amoA gene has been analyzed in soil, as well as in estuarine sediments as a measure of the transcriptional activity, indicating the protein production potential (Nicol et al., 2008; Abell et al., 2011). In estuarine sediments, the expression of bacterial amoA genes was reduced at low dissolved O2 (DO) concentrations, whereas the archaeal amoA gene expression was not affected, suggesting that AOB and AOA show different responses to changes in sedi-ment DO, which may be a significant factor in niche partitioning of different ammonia oxidizer groups. The gene expression of other functional genes has also been determined as an indicator for the potential activity of specific groups. For example, the nirS transcript abundance of de-nitrifying bacteria has been quantified in estuarine sediments (Nogales et al., 2002; Smith et al., 2007; Bulow et al., 2008), showing the decrease of potentially active denitrifying bacteria from the head to the mouth of the estuary. The nirS transcript abundance of anammox bacteria has been used to estimate the potential activity of anammox bacteria, showing the contribution of anammox bacteria to nitrogen removal the Peruvian oxygen minimum zone (Lam et al., 2009).

Stable carbon isotopes have been also used to track the activity of microorganisms in marine sediments. In contrast to radioactive isotopes (14C), stable carbon isotopes (12C, 13C) do not decay in nature and are less harmful. In nature, stable carbon and nitrogen isotopes occur in different abundances. Whereas the isotope with the low mass (e.g. 12C) is highly abundant (98%), the isotope with the higher mass (e.g. 13C) represents less than 2% (Fry, 2006). The difference in the abundance gives the possibility to track the cycling of elements by the metabolic incorporation of 13C into specific molecules. The labeled substrate is added to an environmental sample (e.g. intact sediment core) and biomarkers are purified and analyzed following the consumption of the substrate. Stable isotope probing (SIP) has been applied to track the incorporation of 13C substrates into membrane lipids, such as archaeal GDGTs (Wuchter et al., 2003), IPL derived GDGTs (Pitcher et al., 2011b) and bacterial PLFAs (Boschker et al., 2002). The 13C incorpo-ration into the biphytanes of GDGTs and into PLFAs can be measured by chromatography combustion isotope ratio mass spectrometry (GC-c-IRMS) (Boschker et al., 1998; Wuchter et al., 2003; Pitcher et al., 2011b). GDGT SIP, i.e. the incorporation of 13C labeled bicarbonate into the biphytanyl chains of crenarchaeol revealed the capability of Thaumarchaeota to grow auto-trophically (Wuchter et al., 2003), which was later confirmed by the isolation and physiological characterization of Candidatus Nitrosopumilus maritimus (Könneke et al., 2005). SIP has been used to label membrane lipids, as well as to track the incorporation of 13C into nucleic acids, such as DNA (DNA-SIP) and RNA (RNA-SIP) (Radajewski et al., 2000, and 2003). PLFA-SIP is the most sensitive method, whereas the DNA-SIP is the less sensitive approach. In contrast to RNA and PLFA, the replication of DNA depends on cell division in presence of the labeled substrate, which makes it difficult to accomplish sufficient incorporation for separation of labeled DNA (Neufeld et al., 2006).

A method that does not imply extraction of labeled compounds, but rather visualizes active microbial cells that have been targeted with a sequence specific labeled probe under the micro-

Page 27: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

19

Introduction

scope, is fluorescence in situ hybridization (FISH). The combination of FISH with catalyzed reporter deposition (CARD-FISH) has been used to estimate the abundance of active microor-ganisms in the marine environment, targeting DNA and RNA molecules as well as rare mole-cules, such as mRNA and genes, with signal amplification using horseradish peroxidase (HRP) (Llobet-Brossa et al., 2002; Pernthaler et. al., 2002; Ishii et al., 2004). For example, the abun-dance of Crenarchaeota has been estimated by CARD-FISH in water samples of the Atlantic Ocean (Wuchter et al., 2006). CARD-FISH has been also combined with other approaches to determine metabolic activity in chemolithoautotrophic microorganisms. For example, catalyzed reporter deposition (CARD)-FISH combined with microautoradiography (MAR) demonstrated that novel groups of Gammaproteobacteria attached to sediment particles have been shown to incorporate 14C-labeled CO2 by MAR-FISH concluding that these groups performed a chemo-lithoautotrophic metabolism based on sulfur oxidation, and also suggesting that they play a more important role for H2S removal and primary production in marine sediments than previously thought (Lenk et al., 2011).

Another novel technique, providing information about elemental, isotopic and molecu-lar characteristics of single cells is nano-scale secondary ion mass spectrometry (NanoSIMS). NanoSIMS can be combined with SIP as well as with CARD-FISH methods. A combination of CARD-FISH and NanoSIMS has been used to track the fate of carbon and nitrogen in micro-bial aggregates (Behrens et al., 2008). However, the NanoSIMS method is cost-intense and time consuming regarding the sample preparation and fine tuning of the instrument.

1.6. Scope of this thesis The research described in this thesis is focused on the spatial and temporal distribution of chemolithoautotrophic microorganisms in coastal marine sediments. These microbes have not been studied extensively and there is a growing need to determine their role in the sedimentary carbon-, nitrogen-, and sulfur cycle, as well as to identify microbial key-players that are involved, especially in the oxygen transition zone of seasonally hypoxic sediments.

To this end, the diversity, abundance and activity of chemolithoautotrophic microorganisms, such as AOA, AOB, anammox and denitrifying bacteria, as well as sulfate reducing and sul-fur oxidizing microorganisms have been determined. Additionally, the diversity and activity of chemolithoautotrophic microorganisms involved in the CBB and rTCA cycle as well as the metabolism of Thaumarchaeota and the abundance of Archaea have been addressed. In the following studies, we focused on coastal marine sediments, exposed to summer hypoxia, i.e. (1) the seasonal hypoxic basin of the Oystergrounds in the central North Sea, (2) different sediment types (sand, clay, and mud) of the Iceland Shelf, and (3) seasonally hypoxic and sulfidic sedi-ments of saline Lake Grevelingen (The Netherlands). The work presented here is divided in the following chapters:

Chapter 2 shows the seasonal and depth distribution of the abundance and potential activity of archaeal and bacterial ammonia oxidizers (AOA and AOB) and anammox bacteria in coastal marine sediments of the southern North Sea. Specific intact polar lipids as well as the gene abundance and expression of the 16S rRNA gene, the ammonia monooxygenase subunit A (amoA) gene of AOA and AOB, and the hydrazine synthase (hzsA) of anammox bacteria were quantified in contrasting seasons (i.e. February, March and August). AOA, AOB and anammox bacteria were detected and transcriptionally active down to 12 cm sediment depth in all seasons.

Page 28: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 1

20

AOA seem to play a more dominant role in aerobic ammonia oxidation compared to AOB, as their abundance was higher in all seasons. Anammox bacteria were abundant and active even in bioturbated and oxygenated sediment layers, especially in August, suggesting that their growth and activity are favored by higher temperatures and lower O2 concentrations during summer stratification. The metabolic activity of the studied microbial groups is spatially coupled lead-ing to the assumption that the rapid consumption of O2 allows the coexistence of aerobic and aerobic ammonia oxidizers in the sediments of the Oyster Grounds in the southern North Sea.

In Chapter 3, we aimed to determine the metabolism of sedimentary Thaumarchaeota and bacteria by studying the incorporation of 13C-label from bicarbonate, pyruvate, glucose and ami-no acids into the thaumarchaeal IPL-GDGTs and bacterial/eukaryotal PLFAs during 4−6 days of incubation in intact marine sediment cores from three different sites on the Iceland shelf. The abundance of thaumarchaeal IPL-GDGTs, in particular HPH-crenarchaeol, were detected at all stations and the concentrations remained constant or decreased slightly during incubation, suggesting that living Thaumarchaeota were present in sediments at all stations. In organic-rich sediments a large uptake of 13C label from organic substrates in bacterial/eukaryotic IPL-derived fatty acids was found, implying that heterotrophic bacteria and potentially non-photosynthetic eukaryotes incorporated label into their biomass, whereas the incorporation of labeled bicar-bonate into the PLFAs of chemolithoautotrophic bacteria and archaea seem to be of minor rel-evance or under the detection limit of the method. In sandy sediments, some enrichment of 13C (1−4%) in thaumarchaeotal IPL-GDGTs was detected after the incubation with bicarbonate and pyruvate, but overall no substantial incorporation of 13C into the biphytanyl chains of HPH-cre-narchaeol has been detected. We conclude that the low incorporation rates detected suggest a slow growth of sedimentary Thaumarchaeota in contrast to other sedimentary organisms and/or a slow turn-over of thaumarchaeotal IPL-GDGTs in contrast to bacterial/eukaryotic PLFAs.

In Chapter 4 the abundance and diversity and potential activity of denitrifying and anammox bacteria, as well as of H2S dependent denitrifying bacteria in seasonally hypoxic and sulfidic sed-iments of saline Lake Grevelingen (The Netherlands) were studied. The depth distribution of 16S rRNA and nirS genes of denitrifying and anammox bacteria, as well as the aprA gene of sul-fur oxidizing and sulfate reducing bacteria, were estimated down to 5 cm sediment depth at three different locations before (March) and during (August) summer hypoxia. The abundance of denitrifying bacteria was relatively stable, whereas anammox bacteria were more abundant and active during summer hypoxia in sediments with lower H2S concentrations, suggesting a possible inhibition of anammox bacteria by sulfide. Heterotrophic and autotrophic denitrifiers as well as anammox bacteria and SRB/SOB spatially coexist, but nirS-type denitrifiers outnumbered anammox bacteria leading to the conclusion that anammox does not contribute significantly to nitrogen removal in Lake Grevelingen sediments. The transcriptional activity of denitrifying and anammox bacteria as well as of sulfur oxidizing, including sulfide-dependent denitrifiers, and sulfate reducing bacteria was potentially inhibited by the accumulation of H2S during sum-mer hypoxia. Both denitrifiers and anammox bacteria could be inhibited by the competition for NO3

− and NO2− with cable and Beggiatoa-like bacteria before summer hypoxia. Results suggest

that H2S detoxification mediated by sulfur oxidizers including sulfide-dependent denitrifiers is not sufficient to sustain the activity of denitrifying and anammox bacteria in Lake Grevelingen sediments.

In Chapter 5 the impact of seasonal hypoxia on the activity and the community structure of

Page 29: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

21

Introduction

chemolithoautotrophic bacteria was studied in seasonally hypoxic and sulfidic sediments of sa-line Lake Grevelingen (The Netherlands) before (March) and during summer hypoxia (August). Combined 16S rRNA gene amplicon sequencing and PLFA analysis revealed a major temporal shift in the community structure. Aerobic sulfur oxidizing Gamma- and Epsilonproteobacteria predominated the chemolithoautotrophic community in spring, whereas Deltaproteobacteria were prevalent during summer hypoxia. Chemolithoautotrophic rates, determined by PLFA-SIP, were three times higher in spring than during summer hypoxia, especially in surface sediments. The abundance of cbbL (RuBisCO) and aclB (ATP-citrate lyase) genes revealed a predominance of microorganisms using the CBB cycle compared to the rTCA cycle independent of the season and location. Gene abundances of both, the cbbL and aclB genes decreased significantly between the seasons, suggesting that the abundance of chemolithoautotrophic bacteria using the CBB or the rTCA cycle is inhibited by the limited availability of electron acceptors, such as O2 and NO3

–, regulated by summer hypoxia.

In Chapter 6 the archaeal abundance and community composition was studied in seasonally hypoxic and sulfidic surface sediments of Lake Grevelingen (The Netherlands). We combined archaeal 16S rRNA gene amplicon sequencing and the analysis of archaeal GDGT IPLs to determine the biological sources of these archaeal lipids under changing O2 and H2S condi-tions, before (March) and during summer hypoxia (August). In spring, the archaeal communi-ty was predominated by aerobic Thaumarchaeota, whereas during summer hypoxia, anaerobic members of the DPANN superphylum prevailed. The abundance of IPL-GDGTs decreased one order of magnitude between March and August, and polar head groups changed from phospho- to glycosidic head groups, suggesting that Thaumarchaeota might be in a stationary growth phase and/or a slow degradation of these GDGTs in the sediment. The switch from a Thaumarchaeota to a DPANN-dominated community during summer hypoxia together with a change in the lipid composition suggest that DPANN archaea may use fossil lipids found in the surface sediment to build up their membranes as they are unable to do so themselves, suggesting that DPANN archaea could be considered as important lipid recyclers in sediment systems and should be taken into account in organic geochemistry interpretations.

Page 30: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 31: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

23

2. Seasonality and depth distribution of the abundance and activity of ammonia oxidizing microorganisms in marine coastal sediments (North Sea)

Yvonne A. Lipsewers, Nicole J. Bale, Ellen C. Hopmans, Stefan Schouten, Jaap S. Sinninghe Damsté, and Laura Villanueva

Microbial processes such as nitrification and anaerobic ammonium oxidation (anammox) are important for nitrogen cycling in marine sediments. Seasonal variations of archaeal and bacte-rial ammonia oxidizers (AOA and AOB) and anammox bacteria, as well as the environmental factors affecting these groups, are not well studied. We have examined the seasonal and depth distribution of the abundance and potential activity of these microbial groups in coastal ma-rine sediments of the southern North Sea. This was achieved by quantifying specific intact polar lipids (IPLs) as well as the abundance and gene expression of their 16S rRNA gene, the ammonia monooxygenase subunit A (amoA) gene of AOA and AOB, and the hydrazine syn-thase (hzsA) gene of anammox bacteria. AOA, AOB and anammox bacteria were detected and transcriptionally active down to 12 cm sediment depth. In all seasons, the abundance of AOA was higher compared to the AOB abundance suggesting that AOA play a more dominant role in aerobic ammonia oxidation in these sediments. Anammox bacteria were abundant and active even in oxygenated and bioturbated parts of the sediment. The abundance of AOA and AOB was relatively stable with depth and over the seasonal cycle, while anammox bacteria abundance and transcriptional activity were highest in August. North Sea sediments thus seem to provide a common, stable, ecological niche for AOA, AOB and anammox bacteria.

Frontiers in Microbiology 5, 472, 2014

Page 32: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

24

2.1. IntroductionNitrogen is an essential component for living organisms and thus plays a critical role in con-trolling primary production (Gruber & Galloway 2008). Microbial aerobic and anaerobic nitro-gen cycling processes such as nitrification, denitrification and anaerobic ammonium oxidation are tightly coupled in marine sediments where 50% of marine nitrogen removal occurs (Zhang et al., 2013, and references therein). The discovery of ammonia-oxidizing archaea (AOA) (Kön-neke et al., 2005; Wuchter et al., 2006) belonging to the phylum Thaumarchaeota (Brochier-Ar-manet et al., 2008), and bacteria of the Planctomycetes phylum performing the anaerobic am-monia oxidation (anammox) reaction (Strous et al., 1999; Kuypers et al., 2003), has changed our perspective on the marine nitrogen cycle and the involved microbial key-players over the last decade. Ammonia oxidation is the first and rate-limiting step of nitrification in which ammonia is oxidized to nitrite by AOA and ammonia oxidizing bacteria (AOB) (see Prosser and Nicol, 2008 for a review). The anammox process involves the combination of ammonium with nitrite to form dinitrogen gas which is then released from the system (Kuenen et al., 2008).

Several studies focusing on ammonia oxidation in marine sediments have suggested that AOA have a more prominent role in marine nitrification than AOB based on the higher abundance of AOA amoA gene copies (Beman & Francis, 2006; Mosier & Francis, 2008; Santoro et al., 2008; Abell et al., 2010). Proposed environmental controls shaping the distribution and activity of AOA in marine and estuarine surface sediments include the presence of sulfide, salinity, dis-solved oxygen (DO) concentration, the availability of phosphorous, temperature and primary production (Bernhard et al., 2010; Dang et al., 2010; Erguder et al., 2009; Mosier & Francis, 2008; Sahan & Muyzer, 2008; Sakami et al., 2012; Bale et al., 2013). Beman et al. (2012) studied the seasonal distribution of AOA and AOB amoA gene abundance in Catalina Harbor down to 10 cm sediment depth and showed higher abundances of AOA and AOB amoA genes during summer compared to winter months as well as coinciding higher 15NH4

+ oxidation rates.For anammox bacteria, studies have shown that temperature, organic carbon content, nitrite

concentration, water depth, sediment characteristics and bioturbation play a role in their distri-bution, abundance and activity in marine and estuarine sediments (Dale et al., 2009; Dalsgaard & Thamdrup, 2002b; Engström et al., 2005; Meyer et al., 2005; Dang et al., 2010; Jaeschke et al., 2009; Li et al., 2010; Zhang et al., 2013; Laverock et al., 2013; Neubacher et al., 2011, 2013). However, most of the studies have focused on the spatial distribution of anammox bacteria abundance and activity only in surface sediments (Dalsgaard & Thamdrup, 2002a; Engström et al., 2005; Thamdrup & Dalsgaard, 2002; Nicholls & Trimmer, 2009; Bale et al., 2014). Limited studies have been focused on the abundance of anammox bacteria with sediment depth or the impact of seasonality on this group (Jaechke et al., 2009; Zhang et al., 2013). Anammox bac-teria abundance and activity has been previously detected using anammox specific ladderane biomarker lipids in combination with 15N-labeling experiments in continental shelf and slope sediments of the Irish and Celtic Sea down to eight centimeters depth with spatial differences (Jaeschke et al., 2009). Recently, anammox bacteria abundance was shown to be higher during summer compared to the winter months in surface sediments of three hyper-nutrified estuarine tidal flats of Laizhou Bay (Bohai Sea, China) through the quantification of anammox bacteria 16S rRNA gene (Zhang et al., 2013). A study by Laverock et al. (2013) indicated that the tem-poral variation in the abundance of N-cycling functional genes is directly influenced by biotur-bation activity varying between different bacterial and archaeal N-cycling genes. The relative

Page 33: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

25

Introduction/Material and Methods

abundance of AOB amoA genes is significantly affected by bioturbation whereas the relative abundance of AOA amoA seems to be controlled by other factors (Laverock et al., 2013). Inter-estingly, bioturbation and mixing can extend the area of nitrate reduction in sediments leading to the production of nitrite which would fuel the anammox process (Meyer et al., 2005).

The aim of our study was to determine differences in abundance and activity of ammonia oxidizers with sediment depth (down to 12 cm depth) in marine sediments from the southern North Sea, a continental shelf sea. To address this issue, we quantified the abundance of 16S rRNA gene of ammonia oxidizer groups (AOA, AOB, and anammox bacteria), metabolic genes (amoA of AOA and AOB, and hzsA gene of anammox bacteria), and the intact polar membrane lipids specific for anammox bacteria (i.e. ladderane lipids, Sinninghe Damsté et al., 2002b), and Thaumarchaeota (i.e. crenarchaeol; Sinninghe Damsté et al., 2002a). The potential transcrip-tional activity of the target genes was analyzed to elucidate the potential activity of the target organisms in the sediment. Our results suggest that both aerobic and anaerobic ammonia oxi-dizers coexist in the same niche and are actively involved in the nitrogen cycle in oxygenated and bioturbated sediments in the North Sea.

2.2. Material and Methods

2.2.1. Study site and sampling The Oyster Grounds, an almost circular depression with a maximum depth of 50 m located in the southern North Sea (Weston et al., 2008), is a temporary deposition center for sediment (Raaphorst et al., 1998), which plays an important role in the carbon and nitrogen cycle in the region (Weston et al., 2008). The sedimentation of organic matter in this region leads to organic rich and muddy sediment, which supports high levels of benthic fauna (Van Raaphorst, 1992; Duineveld, 1991).

Sediment cores were collected at a station (4˚33.01’ E, 54˚13.00’ N) in the Oyster Grounds during three cruises on board of the R/V Pelagia in February, May and August 2011. Sediment was collected with 10 cm diameter multicores. Bottom water of the overlaying water was col-lected using a syringe and filtered (through a 0.45 µm 25 mm Acrodisc HT Tuffryn Membrane syringe filter) for nutrient analysis. The cores were sliced into 1 cm slices using a hydraulic slicer or a manual slicer. Samples were collected for lipid and DNA/RNA analysis and kept at –80ºC (DNA/RNA) and –40ºC (lipids) until processing.

Bottom water temperature, water depth and salinity were measured using a SBE911+ conduc-tivity-temperature depth (CTD) system (Seabird Electronics, Inc., WA). The oxygen concentra-tion in the water column was measured using a dissolved oxygen (DO) sensor SBE 43 (Seabird Electronics, Inc., WA) integrated in the CTD system.

2.2.2. Physicochemical parameters Pore water was extracted from 2.5 mm sediment slices from 0 to 2 cm depth, and from 1 cm sediment slices from 2–12 cm depth by centrifugation (approx. 4000 g, 5 min, through a 0.45 µm 25 mm Acrodisc HT Tuffryn Membrane syringe filter), and the obtained concentrations of the first 2 cm were averaged for the first and second centimeter of the sediment to estimate pore water nutrient concentrations in a 1 cm resolution. Pore water samples were stored in pre-rinsed pony vials and NH4

+, NO3–, NO2

– and PO43– concentrations were analyzed as described by

Bale et al. (2013). The total organic carbon (TOC) and total organic nitrogen (TON) content

Page 34: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

26

per gram of freeze dried sediment were determined after acidification of freeze dried sediment (0.5−1 g) with 2N HCl. Residues were analyzed by using a Thermo Finnigan Delta plus iso-tope ratio monitoring mass spectrometer (irmMS) connected to a Flash 2000 elemental analyzer (Thermo Fisher Scientific, Milan).

2.2.3. Intact polar lipid extraction and analysis IPLs were extracted following Bale et al. (2013). Hexose-phosphohexose (HPH) crenarchaeol, a specific biomarker lipid for AOA (Schouten et al., 2008b), was analyzed using high performance liquid chromatography mass spectrometry (HPLC/MS) as described by Pitcher et al. (2011a). Due to the lack of a standard the results are reported here as response unit per gram of sedi-ment (r.u. g–1). The anammox bacteria specific intact ladderane phospholipid, C20-[3]-monoether ladderane, attached to a phosphatidylcholine (PC) headgroup (PC-monoether ladderane) was analyzed by HPLC/MS following Jaeschke et al. (2009) and quantified using an external standard consisting of isolated PC-monoether ladderane.

2.2.4. DNA/RNA extraction DNA and RNA from sediment cores were extracted by using the DNA and RNA PowerSoil® Total Isolation Kit, respectively (Mo Bio Laboratories, Inc., Carlsbad, CA). Nucleic acid con-centrations were quantified spectrophotometrically (Nanodrop, Thermo Scientific, Wilming-ton, DE) and checked by agarose gel electrophoresis for integrity. Extracts were kept frozen at –80°C. The RNA extracts were treated with RNase-free DNase (DNA-free™, Ambion Inc., Austin, TX). RNA quality and concentration were estimated by the Experion RNA StdSens Analysis Kit (Bio-Rad Laboratories, Hercules, CA). DNA contamination was checked by PCR using RNA as a template.

2.2.5. Reverse transcription (RT)-PCR Reverse transcription (RT) was performed with the Enhanced Avian First Strand synthesis kit (Sigma-Aldrich Co., St Louis, MO) using random nonamers as described previously (Holmes et al., 2004). Two negative controls lacking reverse transcriptase or RNA were included. PCR reac-tions were performed as described above to confirm the transcription to complementary DNA (cDNA) and the negative controls using the RT reaction as a template.

2.2.6. Quantitative PCR analysis Quanititative (qPCR) analyses were performed on a Biorad CFX96TM Real-Time System/C1000 Thermal cycler equipped with CFX ManagerTM Software. Gene copy numbers of the Thaumarchaeota group 1.1a were estimated by using the 16S rRNA gene primers MCGI-391F/MCGI-554R (Coolen et al., 2007). The AOA amoA gene was amplified using the primer com-bination AmoA-ModF/AmoA-ModR (Yakimov et al., 2011). Abundance of AOB 16S rRNA gene was estimated using primers CTO189F/CTO654R as described by Kowalchuk et al. (1997). Gene copy numbers of the AOB amoA gene were estimated using the primer set AOB-amoAF/AOB-amoAR new (Rotthauwe et al., 1997; Hornek et al., 2006). Abundance of the anammox bacteria 16S rRNA gene was estimated using primers Brod541F/Amx820R as described by Li et al. (2010). Additionally, the anammox bacteria hzsA gene was quantified using the primer combination hzsA_1597F/hzsA_1857R as described by Harhangi et al. (2012) (See Table S1 for details). All qPCR reactions were performed in triplicate with standard curves from 100 to 107 molecules per microliter. Standard curves were generated as described before (Pitcher et al.,

Page 35: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

27

Material and Methods

2011b). Gene copies were determined in triplicates on diluted DNA extracts (1:10) and on com-plementary DNA (cDNA) extracts. The reaction mixture (25 μL) contained 1 U of PicoMaxx high fidelity DNA polymerase (Stratagene, Agilent Technologies, Santa Clara, CA) 2.5 μL of 10x PicoMaxx PCR buffer, 2.5 μL 2.5 mM of each dNTP, 0.5 μL BSA (20 mg mL–1), 0.02 pmol μL–1 of primers, 10,000 times diluted SYBR Green® (Life Technologies, Carlsbad, CA) (optimized concentration), 0.5 μL 50 mM of MgCl2 and ultra-pure sterile water. All reactions were per-formed in iCycler iQTM 96-well plates with optical tape (Bio-Rad, Hercules, CA). Specificity of the reaction was tested with a gradient melting temperature assay. The cycling conditions for the qPCR reaction were the following: 95°C, 4 min; 40–45 × [95°C, 30 s; melting temperature (Tm), 40 s; 72°C, 30 s]; final extension 80°C, 25 s. Specificity for qPCR reaction was tested on agarose gel electrophoresis and with a melting curve analysis (50°C–95°C; with a read every 0.5°C held for 1 s between each read) in order to identify unspecific PCR products such as primer dimers or fragments with unexpected fragment lengths. Melting temperature, PCR efficiencies (E) and correlation coefficients for standard curves are listed in Table S2.

2.2.7. PCR amplification and cloning Amplifications of the anammox bacteria 16S rRNA gene, AOA and AOB amoA genes were performed with the primer pairs specified above (Table S1). PCR reaction mixture was the fol-lowing (final concentration): Q-solution (PCR additive, Qiagen, Valencia, CA) 1 ×; PCR buffer 1 ×; BSA (200 µg mL–1); dNTPs (20 µM); primers (0.2 pmol µL–1); MgCl2 (1.5 mM); 1.25 U Taq polymerase (Qiagen, Valencia, CA). PCR conditions for these amplifications were the fol-lowing: 95°C, 5 min; 35 × [95°C, 1 min; Tm, 1 min; 72°C, 1 min]; final extension 72°C, 5 min. PCR products were gel purified (QIAquick gel purification kit, Qiagen, Valencia, CA), cloned in the TOPO-TA cloning® kit (Life Technologies, Carlsbad, CA), and transformed in E. coli TOP10 cells following the manufacturer’s recommendations. Recombinant plasmidic DNA was sequenced using M13R (5’-CAG GAA ACA GCT ATG AC-3’) primer by Macrogen Inc. (Am-sterdam, The Netherlands).

2.2.8. Phylogenetic analysis Sequences were analyzed for the presence of chimeras using the Bellerophon tool at the Green-Genes website (http://greengenes.lbl.gov/). Sequences were aligned with Mega5 software (Ta-mura et al., 2011) by using the alignment method ClustalW. Partial sequences of the archaeal amoA gene generated in this study were added to the amoA gene reference tree provided in Pester et al. (2012) using the ARB Parsimony tool (Ludwig et al., 2004). Phylogenetic trees of AOB amoA and the anammox bacterial 16S rRNA genes were computed with the Neigh-bour-Joining method (Saitou and Nei, 1987) in the Mega5 software. The evolutionary distances were estimated using the Jukes-Cantor method (Jukes and Cantor, 1969) with a bootstrap test of 1,000 replicates. Sequences were deposited in NCBI with the following accession numbers: KJ807530–KJ807556 for partial sequences of the archaeal amoA gene, KJ807557–KJ807597 for partial bacterial amoA gene sequences and KJ807598–KJ807609 for partial anammox bacterial 16S rRNA gene sequences.

2.2.9. Statistical analysis Spearman’s rank order correlation coefficient (rs) analysis was performed using the SigmaPlotTM (12.0) Exact Graphs and Data Analysis (Systat Inc., San Jose, CA). Additional multivariate anal-

Page 36: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

28

ysis did not show additional correlations of environmental data with gene and expression data therefore results were not included in this study.

2.3. Results

2.3.1. Physicochemical conditionsOxygen concentrations of the water column corresponding to the sampling time and location of sediment core sampling are shown in Fig. S1. In February, oxygen concentrations decreased throughout the water column from 301.3 to 299 µM (300 µM on average). In May, the oxygen concentration was slightly lower compared to February (282 µM on average) and decreased be-tween 17.8–22.8 m water depth from about 287 to 278 µM. Lowest oxygen concentrations were detected in August compared to February and May (230 µM on average) and decreased between 17.8–24.8 m water depth from 242 to 217.25 µM.

Physicochemical parameters of the bottom water during the cruises of February (winter), May (spring) and August (summer) are shown in Table S3. Bottom water temperatures ranged from 5 to 15.4˚C. Salinity values were stable throughout the year ranging from 34.3–34.7 practical sa-linity units (psu). Ammonia (NH4

+) concentrations in the bottom water ranged between 1.6 and 3 µM with a maximum concentration in February. Bottom water nitrite (NO2

–) concentrations varied from 0.6 µM in February to 0.1 µM in May and August. Highest nitrate (NO3

–) concentra-tions were detected in February (1 µM) and lowest in August (0.5 µM).

Pore water concentrations of NH4+ ranged between 3–53 µM in all three seasons (Fig. 1). The

depth trend of ammonia concentrations in the pore water was similar in all seasons with lower concentrations in the upper 4 cm of the sediment (3–34 µM) compared to increasing concen-trations of the underlying layers (20–53 µM) down to 12 cm sediment depth. Highest NH4

+ concentrations were detected in August (40–53 µM) compared to February and May. Nitrite (NO2

–) concentrations ranged between 0.1–1.2 µM in all seasons (Fig. 1). Depth profiles of pore water nitrite concentrations showed different seasonal trends. In February, nitrite concentrations were relatively low (0.3 µM) in the first 3 cm of the sediment and increased with depth (0.6 µM) whereas in May and August, nitrite concentrations were higher between 1–2 cm sediment depth (0.8 µM) compared to the underlying sediment layers (0.4 and 0.2 µM, respectively). Pore water nitrate (NO3

–) concentrations strongly varied with sediment depth and season (Fig. 1). In Feb-ruary and May, nitrate concentrations were highly variable with depth and reached maximum values between 3−4 cm (38 µM) and 4−5 cm of the sediment (24 µM), while the concentration strongly decreased in the underlying layers. In August, lower nitrate concentrations were detect-ed compared to the other seasons with slightly elevated values in the first 2 cm of the sediment (7.3 µM) and lower concentrations in the underlying layers (1.3 µM). Phosphate concentrations ranged between 1.4–5.2 µM in all seasons (Fig. 1). Pore water phosphate concentrations were stable in the first 4 cm of the sediment (1.7 µM, averaged) in all three seasons. In February and May, phosphate concentrations increased slightly (3.4 µM, averaged) in the underlying sediment layers whereas the depth profile remained stable in August. TOC and TON percentages of the sediment were relatively stable in all seasons (between 0.15–0.38% and 0.02–0.08% respectively), with highest values in May between 1–2 cm and 4–5 cm sediment depth (Table S4).

2.3.2. Abundance, distribution, activity and diversity of AOA and AOBThe abundance of AOA in the sediments was determined by quantification of the 16S rRNA

Page 37: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

29

Material and Methods/Results

and amoA gene copies, as well as by quantification of HPH-crenarchaeol (Figs. 2A–C Table S5). The seasonal variations in depth profiles of AOA 16S rRNA and amoA gene abundance were generally similar (Figs. 2A, B). Both AOA gene abundances were higher between 0−5 cm (6 × 106 and 1.4 × 107 gene copies g–1, respectively) compared to 5−12 cm of the sediment (5 × 106 and 8.2 × 106 gene copies g–1, respectively). Gene copy numbers of the AOA amoA gene were higher in August in comparison to the values observed in February and May (1.4-fold and 2.7-fold higher, respectively) for the 0–5 cm interval. HPH-crenarchaeol concentration was variable over the first 5 cm of the sediment (between 3 × 106–17 × 106 r.u. g–1), but stable values (around 5.9 × 106 r.u. g–1) were detected in the layers underneath. No clear seasonal differences were detected (Fig. 2C).

The abundance of AOB was determined by quantifying the AOB 16S rRNA and amoA gene (Figs. 2D, E). Values were relatively stable with depth with seasonal variations between 0–4 cm of the sediment. Abundance of the AOB 16S rRNA gene showed values similar to the AOA 16S rRNA gene abundance in February and August (on average 5.1 × 106 and 5.8 × 106 gene copies g–1, respectively), while in May AOB 16S rRNA gene copy numbers were lower on average com-pared to the AOA 16S rRNA gene copy number (4.0 × 106 gene copies g–1, 1.4-fold lower). The AOB amoA gene abundance followed the same seasonal trends as that of the AOB 16S rRNA gene, AOA 16S rRNA and amoA gene abundances, with higher values in August, but absolute values were one order of magnitude lower than AOA amoA gene copy numbers (2.3 × 105–5.2 × 106 gene copies g–1). AOB amoA gene abundance was highest in the 0−5 cm depth interval except in May when values between 2−5 cm and in the 5−12 cm depth interval were lower than in the surface sediment.

Figure 1: Sediment pore water nutrients: A) NH4+; B) NO2

–; C) NO3– and D) PO4

3+ [µM] with sediment depth [cm] below sea floor [bsf].

Page 38: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

30

The RNA:DNA ratios of the AOA and AOB 16S rRNA and amoA genes were calculated as an indicator of the potential transcriptional activity of the targeted microbial group (Figs. 3A–F).In all three seasons, AOA 16S rRNA gene RNA:DNA ratio showed relatively stable values down core ranging between 0.9–5.0 and 1.2–3.4 in May and August but reached slightly higher values (2.2–8.2) in February (Fig. 3A). AOB 16S rRNA gene RNA:DNA ratio was between 0.1–2.1 with a minimum of 0.1 between 4−5 cm of the sediment in February and 0.1–3.7 with a minimum between 2−3 cm sediment depth in May. The AOB 16S rRNA gene RNA:DNA ratio was slightly higher (1.2–4.1) in August (Fig. 3C). Values were relatively stable with depth in August whereas in February and May values were more variable. Compared to the RNA:DNA ratio of the AOA 16S rRNA gene, the AOB 16S rRNA gene RNA:DNA ratio was lower (5-fold in February, 1.5-fold in May and 1.3-fold lower in August) in all seasons (Figs. 3A, C). AOA amoA gene transcripts (Fig. 3B) were only detected in February in the first 3 cm and in August throughout the core, and were under the detection limit in May. AOB amoA gene RNA:DNA ratio (Fig. 3D) varied between 4 × 10–5–2 × 10–2 in all seasons with a slight decrease with depth and highest values in August.

The diversity of AOA and AOB was evaluated in two selected depth intervals (0–1 and 9–10 cm sediment depth) of the sediment core recovered in August by targeting the AOA amoA gene and the AOB amoA gene, respectively (Figs. S2, S3). Amplified AOA amoA gene sequences were closely related to AOA amoA gene sequences previously isolated from marine and estuarine environments and did not cluster according to depth (Figs. S2A, B). Sequences were mainly affil-iated to the Nitrosopumilus subclusters 12 and 16, also known as the stable marine cluster, or to

Figure 2: Abundance of A) AOA 16S rRNA gene [copy number g–1]; B) AOA amoA gene [copy number g–1]; C) Hexose phosphohexose (HPH)-crenarchaeol [response units g–1], and D) AOB 16S rRNA gene [copy number g–1]; E) AOB amoA gene [copy number g–1] with sediment depth [cm] below sea floor [bsf].

Page 39: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

31

Results

the Nitrosopumilus subcluster 4.1, known as the estuarine cluster, according to the phylogenetic classification proposed by Pester et al. (2012). The AOB amoA gene sequences were affiliated to AOB amoA gene sequences previously isolated from estuarine sediments. Sequences were close-ly related to the AOB amoA gene sequences of Nitrospira sp. (accession number X90821) and Nitrosolobus multiformis (accession number AF042171). No clustering of AOB amoA sequences according to depth was observed (Fig. S3).

2.3.3. Abundance, distribution, activity and diversity of anammox bacteriaThe abundance and distribution of anammox bacteria were estimated by quantification of ana-mmox bacteria 16S rRNA and hzsA gene copy number, as well as by quantification of the PC-monoether ladderane (Figs. 4A–C). Depth profiles of anammox bacteria 16S rRNA and hzsA gene abundances followed similar trends. Gene copy numbers of the hzsA gene were one order of magnitude lower (between 3.5 × 105 and 2 × 106 copies g–1) than anammox bacteria 16S rRNA gene copy numbers in all three seasons (Figs. 4 A, B). Values were higher in August compared to those in February and May, especially for the anammox bacteria 16S rRNA gene (3.7-fold higher).

The RNA:DNA ratio for the anammox bacteria 16S rRNA gene varied between 1.6–34.6 (Fig. 3E) with higher values in August compared to February and May (on average 4-fold higher). The hzsA gene RNA:DNA ratio (Fig. 3F) depth profiles showed relatively stable values (between 0.008–0.125) throughout the core without significant seasonal differences. The PC-monoether ladderane concentration (Fig. 4C) was variable over the first 5 cm of the sediment (between 0.25–1.25 ng g–1), while in the underlying sediment layers values decreased slightly to more stable

Figure 3: Transcriptional activity given as RNA:DNA ratio of A) AOA 16S rRNA gene; B) AOA amoA gene; C) AOB 16S rRNA gene, D) AOB amoA; E) Anammox 16S rRNA gene, and F) Anammox bacteria hzsA gene with sediment depth [cm] below sea floor [bsf].

Page 40: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

32

values (approximately 0.36 ng g–1) down to 12 cm sediment depth. There was no clear seasonal difference observed.

The diversity of anammox bacteria was evaluated in two selected depths (0–1 and 9–10 cm) of the sediment core in August by targeting the 16S rRNA gene of anammox bacteria. The sequences were closely related to “Candidatus Scalindua marina” (accession number EF602038) and “Candidatus Scalindua brodae” (accession number AY254883) and to sequences previously detected in waters of marine oxygen minimum zones (Woebken et al., 2008). Anammox bacterial 16S rRNA gene sequences from different depths were closely related to each other (Fig. S4).

2.4. Discussion

2.4.1. Aerobic ammonia oxidizersIn this study, we defined two niches in the Oyster Grounds sediment based on the differences in physicochemical parameters and in the abundance of the target organisms, i.e. the upper 4 cm of the sediment, and the underlying sediment. Ammonia concentration in the pore water of the upper part of the sediment (0–4 cm) was consistently lower than in deeper sediment layers (4–12 cm) (Fig. 1), suggesting ammonia diffusion from below formed by mineralization of or-ganic matter by ammonification or dissimilatory nitrate reduction to ammonium (DNRA). In addition, the abundance of AOA and AOB (as well as anammox bacteria, see below) was higher and more affected by seasonality in the upper part than deeper in the sediment, which apparently is a more stable compartment. The higher variability observed in the upper part of the sediment is in agreement with the seasonal mixing of the bottom water layers and sediment deposition

Figure 4: Abundance of A) Anammox bacteria 16S rRNA gene [copy number g–1]; B) Anammox bacteria hzsA gene [copy number g–1], and C) PC-monoether ladderane [ng g–1] with sediment depth [cm] below sea floor [bsf].

Page 41: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

33

Results/Discussion

dynamics in this area (Greenwood et al., 2010). Oxygen availability is expected to play an important role in the dynamics of ammonia oxidiz-

ing microorganisms in marine sediments. In the summer months, water stratification leads to a decrease in the dissolved oxygen concentration in the bottom water in the Oyster Grounds area (Greenwood et al., 2010) (see Fig. S1). Indeed, bottom water hypoxia has been shown to cause a decrease in oxygen penetration depth and oxygen consumption in Oyster Grounds sediments by induced short-term hypoxia in intact sediment cores (Neubacher et al., 2011). Lohse et al. (1996) reported an oxygen penetration depth (OPD) of 1.7 cm in February and 0.5 cm in August in sediment at a station close to the Oyster Grounds using oxygen microelectrodes, suggesting per-manently anoxic conditions from a depth of approximately >2 cm throughout the year. How-ever, the OPD into the sediment must be interpreted with caution because the Oyster Grounds sediment is known to be an area of intense bioturbation by macro- and meiobenthos even in periods of low oxygen concentration in the bottom water (De Wilde et al., 1984; Duineveld et al., 1991), thus the import of oxygen by macrofaunal burrows into the deeper sediment could be relevant (Kristensen, 2000; Meysman et al., 2006). A study by Laverock et al. (2010) has also shown that bioturbation can induce changes in the microbial community composition, resulting in distinct communities within the burrow. However, in our study we did not observe changes in the diversity of amoA gene sequences of AOA or AOB, which might indicate that the changing physicochemical conditions in this system do not select for specific ammonia oxidizing micro-organisms. The depth and seasonal variability for AOB was more pronounced than for AOA in the first 4 cm of the sediment, which is consistent with the fact that the relative abundance of AOB is more significantly affected by bioturbation than AOA according to previous studies (Laverock et al., 2013).

AOA and AOB 16S rRNA and amoA genes and AOA-IPLs and were detected up to 12 cm depth in Oyster Grounds sediments. The detection of aerobic ammonia oxidizing microorgan-isms in marine sediment at depths where the oxygen concentration is expected to be undetect-able seems counterintuitive with their aerobic metabolism. However, the intense bioturbation reported in the Oyster Grounds sediments, as mentioned above, might provide enough oxygen to support the activity of these microbial groups. In fact, Beman et al. (2012) detected AOB and AOA amoA genes up to 10 cm sediment depth as well as readily detectable 15NH4

+ oxidation rates up to 9 cm sediment depth with maxima overlapping with a peak in pore water NO3

– + NO2

– concentrations between 3–7 cm sediment depth in bioturbated sediments of Catalina Island (CA, USA) . Thus, our results, and those of Beman et al. (2012), suggest that aerobic am-monia oxidizers are indeed living and actively involved in the marine sedimentary nitrogen cycle deeper (i.e. up to at least ca. 10 cm) in coastal marine sediments.

The detection of a potential transcriptional activity of the AOA 16S rRNA gene up to 12 cm depth supports that AOA are indeed active deeper in the marine sediment. The potential AOA 16S rRNA gene transcriptional activity was higher during winter, whereas the abundance was slightly higher during summer months. Previous studies showed higher AOA abundance during winter months in the in the North Sea water column (Wuchter et al., 2006; Herfort et al., 2007; Pitcher et al., 2011c) attributed to the higher availability of ammonia and the lack of competition for ammonia with phytoplankton. However, none of these studies targeted the transcriptional activity of AOA. Recently, Bale et al. (2013) observed that the abundance of the benthic Thaumarchaeota (given by HPH crenarchaeol) in the surface sediment (0–1 cm) at the

Page 42: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

34

Oyster Grounds was on average highest in the summer (August) and lowest in the winter (No-vember), which they attributed to the deposition of algal-bloom-derived organic matter taking place in late spring and summer resulting in the formation of ammonia. However, it should also be noted that Bale et al. (2013) did find that transcriptional activity of the AOA 16S rRNA gene was higher during winter in the surface sediment. The disparity in seasonality between the results of these two studies highlights the difference between the seasonal ammonia dynamics at the surface sediment/water interface (Bale et al., 2013) and with depth in the sediment (this study).

To complement the determination of potential transcriptional activity of AOA based on 16S rRNA gene, we also determined the AOA amoA gene RNA:DNA ratio, which was possible for samples from August throughout the core but only in the upper 3 cm depth in February because the quantity of transcripts in the other seasons and depths was under the detection limit of the qPCR assay.

The relationship of amoA gene expression and in situ ammonia oxidation is still unclear (Lam et al., 2007; Nicol et al., 2008; Abell et al., 2011; Vissers et al., 2013). However, recent studies have observed a good correlation between nitrification rates and abundance of AOA amoA gene and transcripts in coastal waters (Urakawa et al., 2014; Smith et al., 2014). Therefore, the detection of AOA amoA transcripts in our setting suggests that AOA were active in the summer throughout the sediment in comparison with the winter in which AOA would be only active in the upper part of the sediment (upper 2 cm).

AOB 16S rRNA gene abundance was in the same order of magnitude than AOA 16S rRNA gene abundance, however, AOB amoA gene abundance was one order of magnitude lower than AOA amoA gene abundance, which can be explained by biases introduced by the primers or by differences in copy number of 16S rRNA and amoA genes per cell. In fact, it is assumed that 3.3 ± 1.6 copies of 16S rRNA can exist per cell in AOB (Větrovský et al., 2013), whereas marine Crenarchaeota seem to contain only one copy per genome (Klappenbach et al., 2001; Laverock et al., 2013). For the amoA gene it is assumed that AOB genomes contain 2 to 3 copies (Arp et al., 2007) while AOA genomes contain only one copy. In addition, the transcriptional activity of the AOB 16S rRNA gene was lower compared to the one of AOA 16S rRNA gene in all seasons.

On the other hand, the abundance of AOB amoA gene transcripts was higher than for AOA and was detected up to 12 cm depth in all seasons which would suggest a higher AOB activity vs AOA. However, a recent study by Urakawa et al. (2014) in water of the Puget Sound Estuary did not observe a correlation between nitrification rates and AOB amoA transcripts. In addition, previous studies have detected the presence of AOB amoA transcripts in non-active starved cells (Bollmann et al., 2005; Geets et al., 2006; Ward et al., 2011) indicating that the AOB amoA tran-scriptional activity is not a good indicator of AOB nitrification activity due to a longer half-life of AOB amoA transcripts than that expected for AOA amoA transcripts.

Taking together the AOA and AOB amoA gene quantification results, AOA have a more im-portant role in the nitrification process in Oyster Grounds sediments than AOB. In addition, all available evidence points to the presence of some oxygen deeper in the sediments of the Oyster Grounds in all seasons probably due to the intense bioturbation in the area, which favors the activity of both AOB and AOA. In the case of AOA, and due to their higher affinity for oxygen (Martens-Habbena et al., 2009), the niche occupancy of AOA could be potentially larger than for AOB.

Page 43: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

35

Discussion

Previous studies observed a positive correlation between AOA abundance and low phosphate concentrations and suggested that phosphate plays a role in determining the niches of Thau-marchaeota (Dang et al., 2013; Erguder et al., 2009). In our study, AOA amoA gene transcript copy number and RNA:DNA ratio was negatively correlated with phosphate (rs= –0.745 and –0.739 respectively; both P ≤ 0.005) suggesting a lower expression of the AOA amoA gene with an increase in pore water phosphate concentrations. These results support the assumption that some ecotypes of AOA dominate over AOB in environments with low phosphate availability. Previously, homologues of the pst gene, encoding for the Pst phosphate ABC transporter, which is usually activated in phosphate deficient environments, was found in marine metagenomes of thaumarchaeotal origin (Rusch et al., 2007; Dang et al., 2013) emphasizing the importance of phosphate availability in marine environments. No correlations were found between AOB abundance or transcriptional activity with pore water nitrogen species, but a negative correlation with the sediment depth (rs= –0.646; P ≤ 0.005) was observed indicating that sediment depth is a factor determining the AOB abundance and activity in Oyster Grounds sediments which might be explained by the decreasing oxygen availability with depth.

2.4.2. Anaerobic ammonia oxidizersAnammox bacteria 16S rRNA and hzsA gene abundance was detected throughout the analyzed sediment with more pronounced depth and seasonal differences than those observed for AOA and AOB. This was not reflected in the abundance of PC-monoether ladderane lipid.

It is possible that the PC-monoether ladderane is partly derived from fossil material and thus reflects a long-term presence of anammox bacteria averaged throughout the entire year rath-er than an actively living anammox population. This possibility has been already suggested by Bransdsma et al. (2011) and Bale et al. (2014) who found dissimilarities between the PC-mon-oether ladderane concentration and gene copy numbers in surface sediments of the Gullmar Fjord and the Southern North Sea, respectively.

We also observed an up to 10-fold difference in the quantification of anammox bacteria 16S rRNA and hzsA gene abundance (Figs. 4A, B). This discordance between the anammox bacteria gene markers has been previously observed in environmental samples (Harhangi et al., 2012), and it could be due to primer biases as previously suggested (Bale et al., 2014). Both the anammox 16S rRNA and the hzsA gene revealed a higher abundance in August compared to February and May. This may indicate that anammox bacteria are more active and abundant at higher temperatures (15˚C), which is the anammox bacteria temperature optimum in temperate shelf sediments (Dalsgaard and Thamdrup, 2005). Additionally, the oxygen penetration depth is expected to decrease during water stratification between May and October, which would also favor anammox bacterial growth and activity (Dalsgaard and Thamdrup, 2002). Previous studies by Neubacher et al. (2013) observed an increase in anammox activity up to 82% compared to control treatments in mesocosms under sustained hypoxic conditions, which corresponds well with our findings of high anammox activity during summer stratification associated with lower oxygen concentrations in the bottom water (Fig. S1).

The presence of anammox markers in the upper layers of the sediment (0–4 cm) is remark-able as the metabolism of anammox bacteria is expected to be incompatible with the presence of oxygen. However, some studies have reported that anammox bacteria can cope with low oxygen concentrations (Kalvelage et al., 2011; Yan et al., 2012), and that anammox bacteria can

Page 44: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

36

be present and active at high oxygen levels by being dormant or using anaerobic niches (Kuypers et al., 2005; Woebken et al., 2007). It is also possible that a high activity of AOA and AOB and heterotrophic bacteria inhabiting these marine sediments would rapidly consume the oxygen available and allow the activity of anammox bacteria even in presence of oxygen. Some studies have also concluded that the potential for anammox bacteria activity is indeed not affected by bioturbation or physical mixing (Rooks et al., 2012). On the contrary, bioturbation and mixing can extend the area of nitrate reduction and thus the availability of nitrite, which fuels the ana-mmox process (Meyer et al., 2005). Our results thus suggest that anammox bacteria can tolerate the presence of oxygen and co-exist with aerobic ammonia oxidizers and potentially have an important role in the nitrogen cycle in marine sediments. In fact, the potential transcriptional activity of anammox bacteria 16S rRNA and hzsA gene was observed throughout the sediment in all seasons. The transcriptional activity of anammox bacteria 16S rRNA gene was higher in the summer as observed for the anammox bacteria abundance. A recent study by Bale et al. (2014) observed a good correlation between the rate of anammox and anammox bacteria 16S rRNA gene transcript abundance in North Sea sediments, which suggests that the transcription-al activity of anammox bacteria 16S rRNA gene is a good proxy of anammox bacteria activity, and confirm a higher anammox activity in the summer for the sediments analyzed in our study. Nevertheless, quantification results of anammox bacteria 16S rRNA gene transcripts have to be interpret with caution because there is evidence that the ribosome content does not decrease during the period of starvation (Schmid et al., 2005; Li et al., 2011). However, the increase of anammox bacteria 16S rRNA gene transcriptional activity during summer was not observed for the hzsA gene, which remained stable across sediment depth and also for the different seasons, and gene abundances were generally an order of magnitude lower. The low and constitutive transcriptional activity of the hzsA gene indicates that the hzsA gene transcript abundance does not reflect variations of anammox bacteria activity in environmental sediment samples, thus the hzsA gene transcript abundance does not seem to be an adequate biomarker for the estimation of the activity of anammox bacteria. Further experiments, especially with pure cultures and under controlled physiological conditions, should be performed to clarify the regulation of the expression of the hzsA gene.

No evidence of a niche separation of aerobic and anaerobic ammonia oxidizers was observed in the oxygen transition zone of the marine sediments. Surprisingly, aerobic and anaerobic am-monia oxidizers are present and active throughout the year in oxygenated sediment layers as well as in anoxic layers. These results seem to be counterintuitive with regards to the individual oxygen requirements of the targeted microbial groups. We hypothesize that the presence and activity of aerobic ammonia oxidizers (AOA, AOB) is supported by the oxygen supply in deep-er anoxic layers due to bioturbation. Likewise, the presence and activity of anammox bacteria in these sediments is then possible by the rapid consumption of oxygen by aerobic ammonia oxidizers and/or other groups, as well as the availability of nitrite provided either by ammonia oxidizers or by an active nitrate reduction present in bioturbated sediments (Meyer et al., 2005).

2.5. ConclusionOur study has unraveled the coexistence and metabolic activity of AOA, AOB, and anammox bacteria in bioturbated marine sediments of the North Sea, leading to the conclusion that the metabolism of these microbial groups is spatially coupled based on the rapid consumption of

Page 45: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

37

Discussion/Cocnlusion

oxygen that allows the coexistence of aerobic and anaerobic ammonia oxidizers. AOA outnum-bered AOB throughout the year which may be caused by the higher oxygen affinity of AOA compared to AOB. Anammox bacterial abundance and activity were higher during summer, indicating that their growth and activity are favored by higher temperatures and lower oxygen available in the sediments due to summer stratifying conditions in the water column. During the summer, anammox bacteria are probably not in competition with AOA and AOB for ammonia as the concentrations were relatively high in the sediment pore water most likely as a result of ammonification processes and the activity of denitrifiers and DNRA. Further studies are re-quired to get a complete picture of the nature of these interactions in the oxygen transition zone of coastal marine sediments.

Acknowledgements

We thank the captain and crew of the R/V Pelagia for logistic support and E. Panoto, J. van Bleijswijk, H. Witte, and A. Mets for technical support. We thank J. van Oijen and S. Ossebaar for nutrient sample collection and analysis. We thank T. Richter for providing and discussing the CTD data. This work was supported by a grant to J. S. Sinninghe Damsté from the Dutch Orga-nization for Scientific Research (NWO) (NICYCLE; 839.08.331) and a grant to Y.A. Lipsewers (grant number 3062) of the Darwin Center for Biogeosciences.

Page 46: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

38

Figure S1: Oxygen profiles [µM] of the water column in February, May and August.

Page 47: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

39

Supplementary Material

Nitrosopumilus subcluster 9 84

uncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sedimentuncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

Nitrosopumilus subcluster 1 16

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, estuary

uncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, estuary

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured archaeon, Nitrosopumilus cluster, coralNitrosopumilus maritimus SCM1, Nitrosopumilus cluster, marine

uncultured archaeon, Nitrosopumilus cluster, marineuncultured ammonia−oxidizing, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, estuaryMarine metagenome ctg_1101668569825, Nitrosopumilus clusteruncultured crenarchaeote, Nitrosopumilus cluster, freshwater

uncultured crenarchaeote, Nitrosopumilus clusteruncultured archaeon, Nitrosopumilus cluster, freshwater

uncultured crenarchaeote, Nitrosopumilus cluster, freshwateruncultured archaeon, Nitrosopumilus cluster, estuary

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, spongeuncultured crenarchaeote, Nitrosopumilus cluster, estuary

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sedimentuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sediment

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

1cm 1uncultured crenarchaeote, Nitrosopumilus cluster, estuary

Nitrosopumilus subcluster 10 5

uncultured archaeon, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sedimentuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sedimentuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sediment

uncultured archaeon, Nitrosopumilus cluster, coraluncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured crenarchaeote, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus cluster, hydrothermal sediment

uncultured archaeon, Nitrosopumilus cluster, marineuncultured crenarchaeote, Nitrosopumilus clusteruncultured crenarchaeote, Nitrosopumilus cluster

Nitrosopumilus subcluster 11 4

uncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, estuaryuncultured crenarchaeote, Nitrosopumilus cluster, wastewater

uncultured crenarchaeote, Nitrosopumilus clusteruncultured crenarchaeote, Nitrosopumilus cluster, marine

Nitrosopumilus subcluster 12, Oyster Grounds sediment, 8 seq 1cm, 12 seq 10 cm 25

1cm 31cm 4

uncultured archaeon, Nitrosopumilus cluster, marineuncultured archaeon, Nitrosopumilus cluster, coraluncultured ammonia−oxidizing, Nitrosopumilus cluster, aquarium biofilter

uncultured crenarchaeote, Nitrosopumilus cluster, estuaryNitrosopumilus subcluster 13 4uncultured crenarchaeote, Nitrosopumilus cluster, estuary

Nitrosopumilus subcluster 14 7

Nitrosopumilus subcluster 2 26

Nitrosopumilus subcluster 15 8

Nitrosopumilus subcluster 3 13

uncultured crenarchaeote, Nitrosopumilus cluster, spongeuncultured crenarchaeote, Nitrosopumilus cluster, spongeuncultured crenarchaeote, Nitrosopumilus cluster, sponge

Nitrosopumilus subcluster 16, Oyster Grounds sediment, 1 seq 1 cm 7

Nitrosopumilus subcluster 4, Oyster Grounds sediment 3 seq 1 cm 23

Nitrosopumilus subcluster 5 21

uncultured crenarchaeote, Nitrosopumilus cluster, spongeNitrosopumilus subcluster 6 14

Nitrosopumilus subcluster 7 36

uncultured crenarchaeote, Nitrosopumilus cluster, spongeuncultured crenarchaeote, Nitrosopumilus cluster, marine

uncultured ammonia−oxidizing, Nitrosopumilus cluster, aquarium biofilteruncultured crenarchaeote, Nitrosopumilus cluster, wastewateruncultured crenarchaeote, Nitrosopumilus cluster, wastewater

Nitrosopumilus subcluster 8 28

Nitrosotalea cluster 39

Nitrosocaldus cluster 2

315

Nitrososphaera sister cluster 14

bacterial amoA 8

0.10

Nitrososphaera cluster

Figure S2A: Phylogenetic tree of AOA amoA gene sequences retrieved in this study and closest relatives (Bold: sequences retrieved at 1cm sediment depth and sequences retrieved at 10 cm sediment depth; Au-gust 2011 and clusters containing sequences retrieved in this study, find details in Fig. S2B); the scale bar indicates 10% sequence divergence.

Page 48: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

40

Nitrosopumilus subcluster 4

0.10

Nitrosopumilus subcluster 12

10cm 101cm 61cm 1210cm 31cm 710cm 61cm 210cm 13uncultured crenarchaeote, Nitrosopumilus subcluster 12, marine10cm 21cm 910cm 1410cm 121cm 15

10cm 151cm 1410cm 7

10cm 910cm 1

10cm 111cm 8

uncultured crenarchaeote, Nitrosopumilus subcluster 12, marineuncultured archaeon, Nitrosopumilus subcluster 12, marineuncultured archaeon, Nitrosopumilus subcluster 12, marine

Nitrosopumilus subcluster 16uncultured crenarchaeote, Nitrosopumilus subcluster 16, marineuncultured crenarchaeote, Nitrosopumilus subcluster 16, marineuncultured ammonia−oxidizing, Nitrosopumilus subcluster 16, aquarium biofilter

uncultured archaeon, Nitrosopumilus subcluster 16, marineuncultured crenarchaeote, Nitrosopumilus subcluster 16, marine

10cm 5

1cm 131cm 51cm 1010cm 8uncultured crenarchaeote, Nitrosopumilus subcluster 4.1, marine

uncultured ammonia−oxidizing, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, marineuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, marineuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, marineuncultured ammonia−oxidizing, Nitrosopumilus subcluster 4.1, estuaryuncultured ammonia−oxidizing, Nitrosopumilus subcluster 4.1, estuary

uncultured crenarchaeote, Nitrosopumilus subcluster 4.1, marineuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuary

uncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuary

uncultured crenarchaeote, Nitrosopumilus subcluster 4.1uncultured crenarchaeote, Nitrosopumilus subcluster 4.1, wastewater

uncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuaryuncultured crenarchaeote, Nitrosopumilus subcluster 4.1, estuary

uncultured crenarchaeote, Nitrosopumilus subcluster 4, estuary

Figure S2B: Phylogenetic tree of AOA amoA gene sequences retrieved in this study and closest relatives (Bold: sequences retrieved at 1cm sediment depth and sequences retrieved at 10 cm sediment depth; Au-gust 2011) the scale bar indicates 10% sequence divergence.

Page 49: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

41

Supplementary Material

10cm 10cm 10cm 10cm

10cm 10cm 10cm

10cm 1cm 10cm 10cm 10cm 1cm KC736030.1 Uncultured Nitrosomonadaceae bacterium clone (Yangtze estuary sediments)

HM364248.1 Uncultured bacterium clone (estuary sediments Elkhorn Slough CA)

HM364256.1 Uncultured bacterium clone (estuary sediments Elkhorn Slough CA)

AY736864.1 Uncultured ammonia-oxidizing bacterium clone (Monterrey Bay CA)

10cm HQ247909.1 Uncultured ammonia oxidising bacterium clone (estuarine sediment)

1cm 10cm

1cm 10cm

10cm JQ010925.1 Uncultured beta proteobacterium clone (sediment China East Sea)

EU244548.1 Uncultured bacterium clone (sediments Jiaozhou Bay China)

AB373459.1 Uncultured ammonia-oxidizing bacterium clone (aquarium biofilter)

AB373510.1 Uncultured ammonia-oxidizing bacterium clone (aquarium biofilter)

KC735937.1 Uncultured Nitrosomonadaceae bacterium clone (Yangtze estuary sediments)

1cm JQ638783.1 Uncultured beta proteobacterium clone (estuarine sediment)

1cm 1cm 10cm

10cm AY785983.1 Uncultured beta proteobacterium clone (Guaymas Basin hydrothermal plumes particle attached)

1cm 1cm 1cm

1cm 1cm 10cm 1cm 10cm 1cm

1cm 10cm JQ345981.1 Uncultured bacterium clone (sediment of Chongming eastern tidal flat)

JX465211.1 Uncultured ammonia-oxidizing bacterium clone (Laizhou Bay sediment)

1cm 1cm 10cm

10cm 10cm 10cm

X90821.1 Nitrosospira sp.

AF042171.1 Nitrosolobus multiformis

AJ388578.1 uncultured proteobacterium clone (water column freshwater lake Medemblik)

AF272406.1 Nitrosomonas oligotropha

AF272398.1 Nitrosomonas halophila

U51630.1 Nitrosomonas eutropha

L08050.1 Nitrosomonas europaea

100

99100

97

95

93

86

7884

68

66

65

62

6095

89

59

56

5150

50

98

65

100

57

100

55

77

57

100

94

55

96

91

79

79

76

0.05

Figure S3: Phylogenetic tree of AOB amoA gene sequences retrieved in this study; the scale bar indicates 5% sequence divergence.

Page 50: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

42

EU

03

98

43

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d a

na

ero

bic

am

mo

niu

m-o

xid

izin

g b

ac

teri

um

)

EU

47

86

59

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d P

lan

cto

my

ce

tale

s b

ac

teri

um

)

EU

47

86

26

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d P

lan

cto

my

ce

tale

s b

ac

teri

um

)

GQ

35

07

11

.1 Candidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d p

lan

cto

my

ce

te)

EU

49

18

42

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d b

ac

teri

um

)

EU

47

86

40

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d P

lan

cto

my

ce

tale

s b

ac

teri

um

)

AY

25

48

83

.1 C

andidatus

Sc

ali

nd

ua

(Candidatus

Sc

ali

nd

ua

bro

da

e)

EF

60

20

38

.1 C

andidatus

Sc

ali

nd

ua

(Candidatus

Sc

ali

nd

ua

ma

rin

a)

1cm

1

cm

1

0cm

1

0cm

1

0cm

1

cm

1cm

10cm

10cm

1cm

HQ

16

37

32

.1 C

andidatus

Sc

ali

nd

ua

(u

nc

ult

ure

d p

lan

cto

my

ce

te)

1cm

1cm

80

74

50

50

68

0.0

02

Figu

re S

4: P

hylo

gene

tic tr

ee o

f an

amm

ox b

acte

ria 1

6S rR

NA

gen

e se

quen

ces r

etrie

ved

in th

is st

udy;

the

scal

e ba

r ind

icat

es 0

.2%

sequ

ence

div

erge

nce.

Page 51: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

43

Supplementary MaterialTa

ble

S1: P

rimer

pai

rs d

escr

ibed

in th

e te

xt, p

olym

eras

e ch

ain

reac

tion

(PC

R) c

ondi

tions

and

am

plic

on si

ze u

sed

in th

is st

udy.

Ass

ayTa

rget

Prim

er p

air

Tm

Am

plic

on

size

[bp*

]R

efer

ence

qPC

RA

OA

16S

rRN

A ge

neM

CG

I-39

1F (5

’-AA

GG

TTA

RTC

CG

AG

TGRT

TTC

-3’)

qPC

R 61

˚C16

3C

oole

n et

al.

(200

7)M

CG

I-55

4R (5

’-TG

AC

CA

CTT

GA

GG

TGC

TG-3

’)

qPC

R +

PC

R/cl

onin

gA

OA

amo

A g

ene

Am

oA-M

odF

(5’-G

GTC

TGG

YTW

AG

AC

GA

TG-3

’)qP

CR

59.5

˚C62

6Ya

kim

ov e

t al

. (20

11)

Am

oA-M

odR

(5’-G

CC

ATC

CA

TCTG

TAW

GTC

-3’)

PCR

55˚C

qPC

R +

PC

R/cl

onin

gA

nam

mox

bac

teria

16

S rR

NA

gene

Br

od54

1F (5

’-GA

GC

AC

GTA

GG

TGG

GTT

TGT-

3’)

qPC

R 59

˚C27

9Li

et a

l. (2

010)

Am

x820

R (5

’-AA

AA

CC

CC

TCTA

CTT

AG

TGC

CC

-3’)

PCR

58˚C

qPC

RA

nam

mox

bac

teria

hz

sA g

ene

hzsA

_159

7F (5

’-WTY

GG

KTA

TCA

RTA

TGTA

-3’)

qPC

R 51

˚C26

0H

arha

ngi e

t al

. (20

12)

hzsA

_185

7R (5

’-AA

ABG

GYG

AA

TCA

TART

GG

C-3

’)

qPC

RA

OB

16S

rRN

A g

ene

CTO

189F

_A (5

’-GG

AG

RAA

AG

CA

GG

GG

ATC

G-3

’)†

qPC

R 60

˚C46

5K

owal

chuk

et

al.

(199

7)C

TO18

9F_C

(5’-G

GA

GG

AA

AG

TAG

GG

GA

TCG

-3’)†

CTO

654R

(5’-C

TAG

CY

TTG

TAG

TTTC

AA

AC

GC

-3’)

qPC

R +

PC

R/cl

onin

gA

OB

amoA

gen

eA

OB-

amoA

F (5

’-GG

GG

TTTC

TAC

TGG

TGG

T-3’

)qP

CR

61.6

˚C49

1Ro

tthau

we

et a

l. (1

997)

, H

orne

k et

al.

(200

6)A

OB-

amoA

R ne

w (5

’-CC

CC

TCBG

SAA

AVC

CTT

CTT

C-3

’)PC

R 58

˚CPC

R co

nditi

ons:

95˚C

5 m

in; 4

0 ×

[95˚

C 1

min

, mel

ting

tem

pera

ture

(Tm) 4

0 s,

72˚C

1 m

in];

72˚C

5 m

in; q

PCR

cond

ition

s: 95

˚C 4

min

; 40

× [9

5˚C

30

s, T m

40

s, 72

˚C 3

0 s]

; 80˚

C 2

5 s;

* bas

e pa

irs (b

p); † T

he fo

rwar

d pr

imer

con

tain

s an

equi

mol

ar m

ixtu

re o

f th

e re

port

ed fo

rwar

d pr

imer

s.

Page 52: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

44

Tabl

e S2

: Qua

ntita

tive

poly

mer

ase

chai

n re

actio

n (q

PCR)

effi

cien

cies

(E) a

nd c

orre

latio

n co

effic

ient

s (R2 ).

Febr

uary

May

Aug

ust

qPC

R a

ssay

targ

et

mol

ecul

eT

m [˚

C]

E [%

]R

2T

m [˚

C]

E [%

]R

2T

m[˚

C]

E [%

]R

2

AO

A 1

6S rR

NA

gen

eD

NA

61.0

100.

00.

998

61.0

93.5

0.98

861

.094

.30.

999

(MC

GI-

391F

/MC

GI-

554R

)cD

NA

61.0

96.1

1.00

061

.098

.80.

998

61.0

97.2

0.99

9

AO

A a

moA

gen

eD

NA

59.5

100.

70.

998

59.5

99.7

0.98

859

.597

.80.

994

(Am

oA_m

odF/

Am

oA_m

odR)

cDN

A59

.599

.90.

983

59.5

107.

60.

984

59.5

96.4

0.99

5

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e D

NA

59.0

99.0

0.99

859

.096

.40.

995

59.0

98.9

0.99

3

(Bro

d541

F/A

mx8

20R)

cDN

A59

.098

.40.

999

59.0

93.7

0.99

659

.010

6.8

0.98

9

Ana

mm

ox b

acte

ria h

zsA

gen

eD

NA

51.0

82.0

0.99

751

.081

.10.

997

51.0

80.1

0.99

7

(hzs

A_1

597F

/hzs

A_1

857R

)cD

NA

51.0

80.0

0.99

851

.080

.20.

997

51.0

80.4

0.99

6

AO

B 16

S rR

NA

gen

eD

NA

60.0

86.6

0.99

860

.086

.60.

998

60.0

86.6

0.99

8

(CTO

189F

/CTO

654R

)cD

NA

60.0

84.0

0.99

860

.084

.00.

998

60.0

84.0

0.99

8

AO

B am

oA g

ene

DN

A61

.685

.40.

994

61.6

85.4

0.99

461

.685

.40.

994

(AO

B-am

oAF/

AO

B-am

oAR

new

)cD

NA

61.6

91.0

0.99

361

.691

.00.

993

61.6

91.0

0.99

3

qPC

R co

nditi

ons:

95˚C

4 m

in; 4

0 ×

[95˚

C 3

0 s,

mel

ting

tem

pera

ture

(Tm) 4

0 s,

72˚C

30

s]; 8

0˚C

25

s

Page 53: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

45

Supplementary MaterialTa

ble

S3: P

hysic

al (C

TD) a

nd c

hem

ical

(nut

rient

) pro

pert

ies o

f th

e bo

ttom

wat

er (B

W).

Para

met

er/

Am

mon

ium

N

itrite

Nitr

ate

Phos

phat

eTe

mpe

ratu

reSa

linity

wat

er d

epth

Sedi

men

t dbs

f [c

m]

[µM

][µ

M]

[µM

][µ

M]

[˚C

][p

su]

[m]

Feb-

11

BW

3.0

0.6

6.8

1.0

5.0

34.4

43.0

May

-11

BW

1.9

0.1

2.0

0.7

8.6

34.3

44.0

Aug

-11

BW

1.6

0.1

0.5

0.5

15.4

34.7

45.0

Page 54: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

46

Table S4: Total organic carbon (TOC) and total organic nitrogen (TON).

Parameter/ TOC TONSediment dbsf [cm] [%] [%]

Feb-110.5 0.22 0.021.5 0.22 bdl*2.5 0.21 0.043.5 0.22 0.044.5 0.26 0.036.5 0.20 0.028.5 0.28 0.0311.5 0.37 0.08

May-11 0.5 0.20 0.021.5 0.30 0.052.5 0.25 0.043.5 0.27 0.034.5 0.38 0.086.5 0.25 0.058.5 0.15 0.0311.5 0.21 0.02

Aug-110.5 0.26 0.031.5 0.27 0.052.5 0.23 0.023.5 0.20 0.024.5 0.31 0.066.5 0.25 0.038.5 0.28 0.0311.5 0.22 0.04

*bdl: below detection limit

Page 55: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

47

Supplementary MaterialTa

ble

S5a:

Spe

arm

an ra

nk o

rder

cor

rela

tion

of a

ll pa

ram

eter

s obs

erve

d in

this

stud

y. Pa

ram

eter

NH

4+N

O3-

PO4+

TO

NA

OA

16S

rRN

A g

ene

AO

A a

moA

gen

e

M]

[µM

][µ

M]

[%]

(RN

A:D

NA

ratio

)(D

NA

) [co

pies

g-1]

Sedi

men

t dep

th [c

m]

0.82

2***

0.80

1***

NH

4+ [µ

M]

0.73

4***

NO

3- [µM

]

PO4+

[µM

]

AO

A 16

S rR

NA

gen

e (D

NA

) [co

pies

g-1]

0.69

4***

AO

A 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

0.84

6***

AO

A a

moA

gen

e (D

NA

) [co

pies

g-1]

AO

A a

moA

gen

e (c

DN

A) [

copi

es g

-1]

AO

A a

moA

gen

e (R

NA

:DN

A ra

tio)

Cre

narc

haeo

l IPL

s tot

al [a

rea

g-1]

HPH

-cre

narc

haeo

l [ar

ea g

-1]

TOC

[%]

0.65

7***

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (R

NA

:DN

A ra

tio)

Ana

mm

ox b

acte

ria h

zsA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria h

zsA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox h

zsA

gen

e (R

NA

:DN

A ra

tio)

PC M

onoe

ther

ladd

eran

e [n

g g-1

]

AO

B 16

S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

B am

oA (D

NA

) gen

e [c

opie

s g-1]

AO

B am

oA (c

DN

A) g

ene

[cop

ies g

-1]

AO

B am

oA g

ene

(RN

A:D

NA

ratio

)

Cor

rela

tion

coef

ficie

nts

r s ≥ 0

.6 a

nd r s ≤

-0.6

with

P-v

alue

s ≤

0.0

5 w

ere

repo

rted

. The

deg

ree

of s

igni

fican

ce w

as in

dica

ted

by u

sing

a sin

gle

aste

risk

for

P-va

lues

≤ 0

.05

and

≥ 0

.01,

two

aste

risks

for P

-val

ues ≤

0.0

1 an

d th

ree

aste

risks

for P

-val

ues ≤

0.0

05.

Page 56: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

48

Tabl

e S5

b: S

pear

man

rank

ord

er c

orre

latio

n of

all

para

met

ers o

bser

ved

in th

is st

udy.

Para

met

erA

OA

am

oA g

ene

AO

A a

moA

gen

eA

nam

mox

16S

rRN

A g

ene

Ana

mm

ox 1

6S rR

NA

gen

e

(c

DN

A) [

copi

es g

-1]

(RN

A:D

NA

ratio

)(D

NA

) [co

pies

g-1]

(cD

NA

) [co

pies

g-1]

Sedi

men

t dep

th [c

m]

NH

4+ [µ

M]

NO

3- [µM

]-0

.647

***

-0.7

43**

*

PO4+

[µM

]-0

.745

***

-0.7

39**

*

AO

A 16

S rR

NA

gen

e (D

NA

) [co

pies

g-1]

AO

A 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

A a

moA

gen

e (D

NA

) [co

pies

g-1]

0.72

5***

AO

A a

moA

gen

e (c

DN

A) [

copi

es g

-1]

0.99

4***

0.62

9***

AO

A a

moA

gen

e (R

NA

:DN

A ra

tio)

Cre

narc

haeo

l IPL

s tot

al [a

rea

g-1]

HPH

-cre

narc

haeo

l [ar

ea g

-1]

TOC

[%]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (D

NA

) [co

pies

g-1]

0.79

7***

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (R

NA

:DN

A ra

tio)

0.71

7***

Ana

mm

ox b

acte

ria h

zsA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria h

zsA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox h

zsA

gen

e (R

NA

:DN

A ra

tio)

PC M

onoe

ther

ladd

eran

e [n

g g-1

]

AO

B 16

S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

B am

oA (D

NA

) gen

e [c

opie

s g-1]

AO

B am

oA (c

DN

A) g

ene

[cop

ies g

-1]

AO

B am

oA g

ene

(RN

A:D

NA

ratio

)

Page 57: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

49

Supplementary MaterialTa

ble

S5c:

Spe

arm

an ra

nk o

rder

cor

rela

tion

of a

ll pa

ram

eter

s obs

erve

d in

this

stud

y.

Para

met

erA

nam

mox

16S

rRN

AA

nam

mox

hzs

AA

nam

mox

hzs

AA

nam

mox

hzs

APC

-ME

ladd

eran

e

(R

NA

:DN

A ra

tio)

(DN

A) [

copi

es g

-1]

(cD

NA

) [co

pies

g-1]

(RN

A:D

NA

ratio

)[n

g g-1

]

Sedi

men

t dep

th [c

m]

-0.6

62**

*

NH

4+ [µ

M]

NO

3- [µM

]

PO4+

[µM

]

AO

A 16

S rR

NA

gen

e (D

NA

) [co

pies

g-1]

AO

A 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

A a

moA

gen

e (D

NA

) [co

pies

g-1]

0.66

2***

AO

A a

moA

gen

e (c

DN

A) [

copi

es g

-1]

0.70

8***

AO

A a

moA

gen

e (R

NA

:DN

A ra

tio)

Cre

narc

haeo

l IPL

s tot

al [a

rea

g-1]

0.79

7***

HPH

-cre

narc

haeo

l [ar

ea g

-1]

0.76

0***

TOC

[%]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (D

NA

) [co

pies

g-1]

0.84

8***

-0.7

51**

*

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

0.74

8***

0.72

9***

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (R

NA

:DN

A ra

tio)

Ana

mm

ox b

acte

ria h

zsA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria h

zsA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox h

zsA

gen

e (R

NA

:DN

A ra

tio)

PC M

onoe

ther

ladd

eran

e [n

g g-1

]

AO

B 16

S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

B am

oA (D

NA

) gen

e [c

opie

s g-1]

AO

B am

oA (c

DN

A) g

ene

[cop

ies g

-1]

0.71

3***

AO

B am

oA g

ene

(RN

A:D

NA

ratio

)

Page 58: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 2

50

Chapter 2 Supplementary MaterialTa

ble

S5d:

Spe

arm

an ra

nk o

rder

cor

rela

tion

of a

ll pa

ram

eter

s obs

erve

d in

this

stud

y.

Para

met

erA

OB

16S

rRN

A

AO

B 1

6S rR

NA

A

OB

16S

rRN

AA

OB

am

oA

AO

B a

moA

(D

NA

) [co

pies

g-1]

(cD

NA

) [co

pies

g-1]

(RN

A:D

NA

ratio

)(c

DN

A) [

copi

es g

-1]

(RN

A:D

NA

ratio

)

Sedi

men

t dep

th [c

m]

-0.6

46**

*

NH

4+ [µ

M]

NO

3- [µM

]

PO4+

[µM

]

AO

A 16

S rR

NA

gen

e (D

NA

) [co

pies

g-1]

AO

A 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

AO

A a

moA

gen

e (D

NA

) [co

pies

g-1]

AO

A a

moA

gen

e (c

DN

A) [

copi

es g

-1]

AO

A a

moA

gen

e (R

NA

:DN

A ra

tio)

Cre

narc

haeo

l IPL

s tot

al [a

rea

g-1]

HPH

-cre

narc

haeo

l [ar

ea g

-1]

TOC

[%]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

Ana

mm

ox b

acte

ria 1

6S rR

NA

gen

e (R

NA

:DN

A ra

tio)

Ana

mm

ox b

acte

ria h

zsA

gen

e (D

NA

) [co

pies

g-1]

Ana

mm

ox b

acte

ria h

zsA

gen

e (c

DN

A) [

copi

es g

-1]

0.65

6***

Ana

mm

ox h

zsA

gen

e (R

NA

:DN

A ra

tio)

PC M

onoe

ther

ladd

eran

e [n

g g-1

]

AO

B 16

S rR

NA

gen

e (c

DN

A) [

copi

es g

-1]

0.62

0***

AO

B am

oA (D

NA

) gen

e [c

opie

s g-1]

0.61

7***

AO

B am

oA (c

DN

A) g

ene

[cop

ies g

-1]

0.70

2***

AO

B am

oA g

ene

(RN

A:D

NA

ratio

)0.

887*

**

Page 59: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 60: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 61: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

53

3. Lack of 13C-label incorporation suggests low turnover rates of thaumarchaeal intact polar tetraether lipids in sediments from the Iceland Shelf

Sabine K. Lengger, Yvonne A. Lipsewers, Henk de Haas, Jaap S. Sinninghe Damsté and Stefan Schouten

Thaumarchaeota are amongst the most abundant microorganisms in aquatic environments, however, their metabolism in marine sediments is still debated. Labeling studies in marine sed-iments have previously been undertaken, but focused on complex organic carbon substrates which Thaumarchaeota have not yet been shown to take up. In this study, we investigated the activity of Thaumarchaeota in sediments by supplying different 13C-labeled substrates which have previously been shown to be incorporated into archaeal cells in water incubations and/or enrichment cultures. We determined the incorporation of 13C-label from bicarbonate, pyruvate, glucose and amino acids into thaumarchaeal intact polar lipid-glycerol dibiphytanyl glycerol tet-raethers (IPL-GDGTs) during 4–6 day incubations of marine sediment cores from three sites on the Iceland Shelf. Thaumarchaeal intact polar lipids, in particular crenarchaeol, were detected at all stations and concentrations remained constant or decreased slightly upon incubation. No 13C incorporation in any IPL-GDGT was observed at stations 2 (clay-rich sediment) and 3 (or-ganic-rich sediment). In bacterial/eukaryotic IPL-derived fatty acids at station 3, contrastingly, a large uptake of 13C label (up to + 80‰) was found. 13C was also respired during the experiment as shown by a substantial increase in the 13C content of the dissolved inorganic carbon. In IPL-GDGTs recovered from the sandy sediments at station 1, however, some enrichment in δ13C (1–4‰) was detected after incubation with bicarbonate and pyruvate. The low incorporation rates suggest a low activity of Thaumarchaeota in marine sediments and/or a low turnover rate of thaumarchaeal IPL-GDGTs due to their low degradation rates. Cell numbers and activity of sedimentary Thaumarchaeota based on IPL-GDGT measurements may thus have previously been overestimated.

Biogeosciences, 11, 201–216, 2014

Page 62: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

54

3.1. IntroductionThaumarchaeota are ubiquitous microorganisms (Hatzenpichler, 2012 and references cited therein) which have recently been discovered to form a separate phylum, the Thaumarchaeota (Brochier-Armanet et al., 2008; Spang et al., 2010). They were initially discovered as the Marine Group I Crenarchaeota in the coastal and open ocean (DeLong, 1992; Fuhrman et al. 1992; Karner et al., 2001) where they are abundantly present in the epipelagic zone (e.g. Francis et al., 2005; Beman et al., 2008) but also in deep water (Fuhrman and Davis, 1997; Herndl et al., 2005; Agogué et al., 2008) where they can be as abundant as the total bacterial population. Thaumar-chaeota have also been detected in marine sediments (Hershberger et al., 1996; Francis et al., 2005).

Before enrichment cultures became available, the metabolism of Thaumarchaeota was unclear. Hoefs et al. (1997) suggested, based on the 13C-enrichment of thaumarchaeal lipids in compari-son to algal biomarkers, the possibility of autotrophic assimilation of dissolved inorganic carbon (DIC) using a different pathway than via Rubisco, or, alternatively, the heterotrophic uptake of small organic molecules. Indeed, 13C-labeling studies with bicarbonate showed the incorporation of 13C-labelled DIC into thaumarchaeal membrane lipids and thus their autotrophic metabolism (Wuchter et al., 2003). Unambiguous proof for the growth of Thaumarchaeota on inorganic carbon came with the first enrichment culture of Nitrosopumilus maritimus, which was growing chemolithoautotrophically on bicarbonate, and gaining energy by the oxidation of ammonia to nitrite (Könneke et al., 2005). It was suggested that 3-hydroxypropionate/4-hydroxybutyrate pathway was used by these microorganisms (Berg et al., 2007), which was supported by genetic analyses of C. symbiosum (Hallam et al., 2006) and N. maritimus (Walker et al., 2010). However, in these two species, also genes involved in heterotrophic metabolisms were detected, which implies their potential for mixotrophy. Furthermore, the uptake of pyruvate into cell material in cultures of a Thaumarchaeote enriched from soil, Nitrososphaera viennensis, was demonstrated (Tourna et al., 2011).

Indeed, there is some environmental evidence for a mixotrophic metabolism of Thaumar-chaeota. Ouverney and Fuhrman (2000) as well as Herndl et al. (2005) showed, via MICRO-CARD-FISH using 3H-labeled amino acids, that thaumarchaeal cells present in sea water take up amino acids. There is also circumstantial evidence for mixotrophy of Thaumarchaeota, such as the radiocarbon values of archaeal lipids in the deep ocean (Ingalls et al., 2006), thaumarchaeal activity (as quantified by uptake of substrate measured by MAR-FISH) being correlated to the presence of urea (Alonso-Sáez et al., 2012), and the possession of genes for urea transporters (Baker et al., 2012). Mußmann et al. (2011) found thaumarchaeal cell numbers in waste reactors that were too high to be supported by the rates of ammonia oxidation, and a lack of incorpo-ration of 13C-labelled DIC into thaumarchaeal lipids during growth, also suggesting heterotro-phy of Thaumarchaeota. Circumstantial evidence, based on the correlation of organic carbon concentration with archaeal biomarker lipids, suggests that Archaea in deep subsurface marine sediments (<1 m) use organic carbon (Biddle et al., 2006; Lipp et al., 2008; Lipp and Hinrichs, 2009), however, this empirical correlation could also be due to preservation factors affecting biomarkers and total organic carbon in a similar way (Hedges et al., 1999; Schouten et al., 2010). These studies thus illustrate the potential for diversity of thaumarchaeal metabolisms.

Cell membranes of Thaumarchaeota consist of glycerol dibiphytanyl glycerol tetraether lip-ids (GDGT) in the form of intact polar lipids (IPL; Schouten et al., 2008; Pitcher et al., 2011a;

Page 63: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

55

Introduction/Material and Methods

Sinninghe Damsté et al., 2012; Fig. 1a). They possess GDGTs which are also present in other Archaea (Koga and Nakano, 2008), such as GDGT-0, -1, -2 and -3 (Fig. 1b), but also a specific lipid named crenarchaeol, which up to now has only been found in (enrichment) cultures of Thaumarchaeota (Sinninghe Damsté et al., 2002; de la Torre et al., 2008; Schouten et al., 2008; Pitcher et al., 2011b; Sinninghe Damsté et al., 2012). Stable isotope probing (SIP) experiments targeting the biphytanyl chains contained in crenarchaeol (Fig. 1c), have been undertaken in order to determine the uptake of different substrates. This revealed the incorporation of 13C from bicarbonate into crenarchaeol (Wuchter et al., 2003). Pitcher et al. (2011c) showed that this uptake was dependent on ammonia oxidation in North Sea water, as addition of inhibitors for ammonia oxidation, nitrapyrine (N-serve) or chlorate, resulted in a decreased incorporation of 13C-labelled DIC. Other SIP experiments carried out with organic substrates were less success-ful: Lin et al. (2012) achieved only a 2‰ enrichment in one of the biphytanyl chains of crenar-chaeol, but 2‰ depletion in the other biphytanyl chain when adding 13C -labeled phytodetrital organic carbon to sediment slurries. Takano et al. (2010) and Nomaki et al. (2011) also carried out benthic in situ-labeling experiments and did not find incorporation of 13C in the biphytanyl chains in an incubation experiment with 13C-labeled glucose in sediment from Sagami Bay, Ja-pan, after 405 days, although they noted an initial apparent increase in δ13C of one biphytane by ~10‰. This was in spite of an increase in molecular thaumarchaeal biomarkers within 9 days. Remarkably, however, they did report labeling of the glycerol moieties of the GDGT molecules after 405 days. Thus, not much evidence for substantial incorporation of carbon from organic substrates into thaumarchaeal lipids is observed in SIP experiments performed with sediments. It is possible that the types of substrate used in these studies were not suitable, i.e. not readily taken up by the Thaumarchaeota, and that other organic compounds would result in higher in-corporation. Interestingly, none of the previously conducted labeling studies in sediments used 13C-labelled DIC. Bicarbonate had previously unambiguously been shown to be incorporated by pelagic Thaumarchaeota (Wuchter et al., 2003; Pitcher et al. 2011c) and by Thaumarchaeota enriched from sediments (Park et al., 2010) and soils (Jung et al., 2011; Kim et al., 2012).

To shed further light on the metabolism of sedimentary Thaumarchaeota, we performed labeling experiments using 13C-labeled substrates, i.e. bicarbonate, pyruvate, amino acids and glu-cose in sediment cores from the Iceland shelf and determined the degree of 13C incorporation into the most abundant GDGT-lipids, i.e. crenarchaeol which is specific for Thaumarchaeota and GDGT-0, which can also be produced by other Archaea. All these compounds have pre-viously been shown to be taken up by Thaumarchaeota either in the environment (Ouverney and Fuhrman, 2000; Wuchter et al. 2003; Herndl et al., 2005; Takano et al., 2010; Pitcher et al. 2011c) or enrichment cultures (Park et al., 2010; Jung et al., 2011; Tourna et al., 2011; Kim et al., 2012) and were thus deemed the most suitable substrates and most likely to result in uptake by sedimentary Thaumarchaeota. The results are discussed with respect to the uptake of substrates as well as the turnover rates of thaumarchaeal GDGTs.

3.2. Material and Methods

3.2.1. Sampling stationsSediment cores were sampled at three stations, located on the Iceland Shelf, during the ‘Long Chain Diols’ cruise on the R/V Pelagia in July 2011 (Fig. 2). Station 1 (63°21’N 16°38’W), south of Iceland, was at 240 m water depth and consisted of dark, sandy sediment.

Page 64: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

56

Station 2 (66°18’N 13°58’W) is located north-east of Iceland, at 261 m water depth, and the sediment was sandy-clayish. Sediment at Station 3 (67°13’N 19°07’W), north of Iceland, at 503 m water depth, and further offshore than the other two stations, consisted of dark, soft, black mud. From each station, multicores were retrieved in polycarbonate core tubes. Ten cores from each station were retrieved in tubes of 10 cm diameter with drilled holes downwards, spaced 1 cm apart, and used for incubation. Cores sampled with tubes of 5 cm diameter were used for oxygen microprofiling, pore-water extraction, and four cores from each station were sliced and stored for density and total organic carbon (TOC) analysis.

3.2.2. Core characterizationFor each station, control cores were sliced into 1 cm thick layers from 0–10 cm core depth. The water content of each sediment was determined by weighing prior to and after freeze-drying. To-tal organic carbon was determined from freeze-dried sediments following overnight acidification with 2N HCl, subsequent washes with bidistilled H2O and water removal by freeze-drying. The decalcified sediments were measured on a Flash EA 1112 Series (Thermo Scientific, Waltham, MA) analyzer coupled via a Conflo II interface to a Finnigan Deltaplus mass spectrometer. Stan-dard deviations from three measurements ranged from 0.0 to 0.6% TOC.

Two cores (5 cm diameter) from each cast were used for oxygen microprofiling with an OX-100 Unisense oxygen microelectrode (Unisense, Aarhus, DK; Revsbech, 1989). A two-point

b

a

Crenarchaeol O

O

O

O

HO

OH

c

Monohexose-crenarchaeol

Dihexose-crenarchaeol

Hexose, phosphohexose-crenarchaeol

OO

O

O

O

O

OHO

HO

OHO OH

HOHO OH

O

OO

O

O

O

O

O PO

OO

HOHO

HO OH

OHOH

OH

HO

OO

O

O

O

O

HO

HOHO OH

O

O

O

O

HO

OH

OH

OH

acyclic biphytane

bicyclic biphytane

tricyclic biphytane

HI, ∆

4

GDGT-0

OO

OO

HO

OH

OO

OO

HO

OH

LiAlH , ∆

Figure 1: Structures of crenarchaeol and GDGT-0 (a), IPL-crenarchaeol species (b) and biphytanyl chains released from crenarchaeol and GDGT-0 by chemical degradation (c).

Page 65: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

57

Material and Methods

calibration was carried out with bottom water from the same station bubbled with air at 4°C for at least 5 min (100% saturation) and a solution of 0.1 M sodium ascorbate in 0.1 N NaOH in Milli-Q Water (0% saturation) and the electric current measured in pA was converted to mg L–1 O2. The sensor was moved with a micromanipulator one mm at a time.

Pore water was extracted on board from slices of several cores by centrifugation. Five mL from each depth was preserved air-free with added HgCl2 for analysis of the δ13C of the dis-solved inorganic carbon (DIC) at 4°C, and 3 mL were preserved at –20°C for analysis of phos-phate, ammonia, nitrite and nitrate.

Dissolved inorganic phosphate, ammonia, nitrite and nitrate were measured on a Traacs 800 Autoanalyzer (Seal Analytical, Fareham, UK). Ortho-phosphate was measured using potassium antimonyltartrate as a catalyst and ascorbic acid as a reducing reagent which resulted in for-mation of a blue (880 nm) molybdenum phosphate-complex at pH 0.9–1.1 (Murphy and Riley, 1962). Ammonia was measured via formation of the indo-phenol blue-complex (630 nm) with phenol and sodium hypochlorite at pH 10.5 in citrate (Helder and de Vries, 1979). Dissolved inorganic nitrite concentrations were measured after diazotation of nitrite with sulphanilamide and N-(1-naphtyl)-ethylene diammonium dichloride to form a reddish-purple dye measured at 540 nm. Nitrate was first reduced in a copperized Cd-coil using imidazole as a buffer, measured as nitrite and concentrations were calculated by subtraction of nitrite (Grasshoff et al., 1983).

δ13C of DIC was measured in pore water from all stations, and bottom water samples from the experiments, on a Thermo GasBench II coupled to a Thermo Delta Plus and is reported in

Iceland100

500Den

mark S

trait

1000

1000

20 ° 10 °

200060 °

30 °

Iceland Basin

20003000

Faroes

Jan Mayen

Greenland

Iceland2000

100

3000

Station 3

Station 2

Station 1

Figure 2: Map of the sampling stations (1, 2 and 3) on the Iceland shelf.

Page 66: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

58

Lower part of splitable core(upper part removed)

90°

Air in (pipette)

Drilled holes 2.5 mmfilled with silicon

Stopper

Stopper

Yellow tape

Distance 1 cm

Airpump

Filter 0.2 mm

Water bath for optimum temperature (4°C)

Figure 3: Scheme of the incubation set-up of the cores.

the delta notation against the Vienna Pee-Dee Belemnite (V-PDB) standard. To this end, 0.5 mL of each sample was injected into gastight exetainers, containing 0.1 mL of 80% H3PO4, that had been flushed for 15 min each with He. Performance was monitored by analysis of an in-house laboratory standard of Na2CO3 with a δ13C of –0.57‰ after every 12 measurements and an in-house laboratory standard of CaCO3, with a δ13C value of –24.2‰ at the start and the end of every series. Measured values of these standards were usually not deviating by more than 0.5‰ from the calibrated (on NBS-19) values.

3.2.3. Incubation conditionsTen cores from each station were used for incubations (Table 2). For this, splitable polycarbonate core tubes with 10 cm diameter were prepared by drilling two rows of 1 mm holes spaced at distances of 1 cm. The two rows were 90° contorted; holes were filled with silicon and covered with plastic non-permeable tape (see setup in Fig. 3). The cores were, after recovery from the multicorer, transferred as quickly as possible to a temperature-controlled room kept at 4°C, and the bottom water was sampled with a sterile pipette for determination of pH, δ13C of DIC and nutrients as described above. Four whole cores at each station were incubated with 13C-la-belled substrates, i.e. bicarbonate, pyruvate, amino acids and glucose, and one whole core (‘back-ground’) was injected with MilliQ water only. Four more cores were incubated with solutions containing one of the substrates and a nitrification inhibitor, and one (‘background + inhibitor’) with MilliQ water and the nitrification inhibitor. The incubation solutions were injected into

Page 67: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

59

Material and Methods

the seven injection ports from the surface downwards, i.e. the first 8 cm were supplied with the labeled substrate. Injection solutions contained 99% 13C-labelled substrates as supplied by Cam-bridge Isotope Laboratories (Andover, MA, USA). The substrates were dissolved in MilliQ-H2O in concentrations of 3.7 mg mL–1 for NaHCO3

– (Bic), 1.01 mg mL–1 for Glucose (Glu), 0.89 mg mL–1 for a mix of 16 algal amino acids (AA; 97–99% labeled; CLM 1548) and 1.24 mg mL–1 of sodium pyruvate (Pyr). 200 μL were delivered to each slice by injection through the two 90° contorted injection ports, which corresponded to a total addition of 1.6 g 13C m–3 sediment for the bicarbonate, and 1.2 g 13C m–3 for the organic compounds. The nitrification inhibitor was Nitrapyrine (N-Serve, Pestanal) at 11.5 μg mL–1, which is known to inhibit archaeal ammonia oxidation (Park et al., 2010; Pitcher et al., 2011c). As Nitrapyrine cannot be dissolved in pure water, 5% of EtOH was added and the solutions were kept at 30°C to prevent precipitation of the Nitrapyrine. Nitrapyrine-free incubation solutions were kept at the same conditions.

After injection, the cores were stored in the dark at 4°C in a water bath - boxes with core holders filled with water for 62 h (38 h for Station 3), after which the bottom water was sampled for 13C-DIC and N- and P-nutrients as described above and re-injected with a fresh solution of the substrates in water, with or without nitrapyrine added. During the incubation time, the cores were covered lightly with a rubber stopper, so that the escape of air was possible, and aerated through a sterile plastic pipette, with air pumped by an aquarium pump and filtered through a 0.2 mm filter for sterilization purposes. After 144 h (96 h for Station 3), bottom water samples for 13C-DIC (dissolved inorganic carbon) and N- and P-nutrients were taken again, and oxygen microprofiles were measured as described above on all cores (except for Station 2, where cores were too short and thus surfaces too low in the core tubes, to allow measurements in more than three cores). Subsequently, the upper 10 cm of the sediment were sliced in 1 cm resolution and each slice mixed thoroughly, split in half and both parts were frozen at –80°C immediately. The incubations at station 3 were shorter due to time constraints.

3.2.4. Lipid extraction, separation and measurementsThe freeze-dried sediments of 0–1 cm, 1–2 cm and 7–8 cm depth intervals of the incubation and background cores were ground and extracted by a modified Bligh-Dyer extraction meth-od (Pitcher et al., 2009). Briefly, they were extracted ultrasonically three times in a mixture of methanol/dichloromethane (DCM)/phosphate buffer (2:1:0.8, v:v:v), centrifuged and the sol-vent phases were combined. The solvent ratio was then adjusted to 1:1:0.9, v:v:v, which caused the DCM to separate. Liquid extraction was repeated two more times, the DCM fractions were combined, the solvent was evaporated and bigger particles were removed over cotton wool. An aliquot of the extracts was stored for high performance liquid chromatography /electron spray ionization tandem mass spectrometry (HPLC/ESI-MS2) analysis, and another aliquot was sepa-rated into core lipid (CL)- and intact polar lipid (IPL)-GDGTs by silica column separation with hexane/ethyl acetate (1:2, v:v) for the CL-fraction and MeOH to elute the IPL-fraction. 0.1 mg C46 internal standard (Huguet et al., 2006) were added to the CL-fractions. The IPL-fraction was split in three aliquots: 5% were used to determine the carry-over of CL-GDGTs into the IPL-fraction and measured directly via HPLC/APCI-MS, 60% were kept for ether cleavage in order to determine the 13C incorporation into the biphytanes (see below), and 35% were subject-ed to acid hydrolysis under reflux for 2 h with 5% HCl in MeOH, followed by addition of water and extraction (3 ×) of the aqueous phase at an adjusted pH of 4–5 with DCM. This yielded the IPL-derived GDGTs which were measured via HPLC/APCI-MS. The C46 internal standard was

Page 68: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

60

used to quantify the IPL-derived GDGTs and the carry over, but was added after the aliquot for 13C analysis was taken in order to prevent any possible interference of alkyl chains released from the internal standard molecule.

3.2.5. δ13C analysis of biphytanes Aliquots (60%) of the IPL-fractions were subjected to chemical treatment in order to cleave GDGTs, releasing the biphytanyl chains (Fig. 1c) according to Schouten et al. (1998). For this, the IPL fraction was refluxed in 57% HI for 1 h to break the ether bonds and produce alkylio-dides and subsequently extracted three times with hexane. The hexane phase was washed with 5% NaS2O7 and twice with water. The alkyliodides were purified over Al2O3 with hexane/DCM (9:1, v:v), hydrogenated with H2/PtO2 in hexane for 1 h (Kaneko et al., 2011) and eluted over Na2O3 with DCM. As the H2/PtO2 hydrogenation-procedure has never been tested for stable carbon isotope analysis, we used four CL-fractions of extracts from Station 1 and Station 2, split them in half and hydrogenated one half with the conventional LiAlH4 reduction method de-scribed by Schouten et al. (1998), the other half with the method described here and by Kaneko et al. (2011). The conventional method consisted of LiAlH4 in 1,4-dioxane for 1 h under reflux, then the remaining LiAlH4 was reacted with ethyl acetate, bidistilled H2O was added and the bi-phytanes were extracted with DCM from the dioxane/H2O mixture. Additional purification was achieved by elution over an Al2O3 column using hexane. The results showed that the δ13C values of the biphytanes were not significantly different when H2/PtO2 treatment and LiAlH4 hydro-genation were compared (Table 1). However, when yields were compared, it was obvious that hydrogenation with H2/PtO2 gave better results. This is probably due to the fact that the H2/PtO2 treatment requires fewer workup procedures after hydrogenation (only one drying step) than the LiAlH4 treatment, which requires two column separations in addition to a drying step.

3.2.6. δ13C of phospholipid-derived fatty acidsIn order to demonstrate that the incubation experiments did result in label uptake by the sedi-mentary microbial community, we determined the uptake of 13C-label by bacteria and eukaryotes by isotopic analysis of polar lipid-derived fatty acids (PLFA). For this, one third of the extract of the 0–1 cm sediment slice from Station 3, from the cores incubated with bicarbonate, glucose, amino acids and the background core (all without nitrapyrine), were used and analyzed for PLFA according to Guckert et al. (1985). Briefly, the aliquots were separated over a silica column with DCM, acetone and methanol, respectively, with the methanol-fraction containing the PLFA. PLFA fractions were saponified in methanolic KOH (2N) and, after adjusting to pH 5, the re-sulting fatty acids were extracted with DCM and methylated with BF3-MeOH (with a δ13C of –25.4 ± 0.2 ‰V-PDB) before measurements by gas chromatography – isotope ratio mass spec-trometry (GC-irMS). The δ13C values for the fatty acids were corrected for the added carbon.

3.2.7. HPLC-APCI-MS and HPLC-ESI-MS2The CL-fractions, hydrolyzed IPL- fractions and the IPL-fractions themselves (to determine the carry-over of CL into the IPL fraction) were analyzed by HPLC-APCI-MS as described previously (Schouten et al., 2007). HPLC-ESI-MS2 in selected reaction monitoring (SRM) mode, as described by Pitcher et al. (2011b), was used to analyze selected IPL-GDGTs. IPL-GDGTs monitored were monohexose (MH)-crenarchaeol, dihexose (DH)-crenarchaeol and hexose, phosphohexose (HPH)-crenarchaeol. Performance was monitored using a lipid extract of sed-

Page 69: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

61

Material and Methods

iment known to contain the monitored GDGTs that was injected every 8 runs. Response areas per mL sediment (mL sed) are reported, as absolute quantification was not possible due to a lack of standards. The values reported are the averages of duplicate injections and their standard deviations.

3.2.8. GC-MS and GC-irMSGC-MS was used to identify the biphytanes and phospholipid-derived fatty acid methyl esters (FAME) using a TRACE GC with a DSQ-MS. The gas chromatograph was equipped with a fused silica capillary column (25 m, 0.32 mm internal diameter) coated with CP Sil-5 (film thick-ness 0.12 mm). The carrier gas was helium. The compound specific isotope composition of the biphytanes and fatty acids was measured with an Agilent 6800 GC, using the same GC column conditions, coupled to a ThermoFisher Delta V isotope ratio monitoring mass spectrometer. The isotopic values were calculated by integrating the mass 44, 45 and 46 ion currents of the peaks and that of CO2-spikes produced by admitting CO2 with a known 13C-content into the mass spectrometer at regular intervals. The performance of the instrument was monitored by

Table 1: δ13C values of the acyclic, bicyclic, and tricyclic biphytanes obtained via H2/PtO2 and LiAlH4 reduction of alkyliodides formed via HI ether cleavage.

δ13C (‰)acyclic bicyclic tricyclic

Sample H2/PtO2 LiAlH4 H2/PtO2 LiAlH4 H2/PtO2 LiAlH4

Station 1 0–1 cm –20.3 ± 0.4 –20.2 ± 0.0 –19.7 ± 0.7 –20.0 ± 0.6 –19.5 ± 1.1 –21.2 ± 0.8

Station 1 1-–2 cm –19.9 ± 0.6 –22.0 ± 0.2 –18.4 ± 0.2 –20.2 ± 1.6 –19.0 ± 0.2 –19.8 ± 0.6

Station 2 0–1 cm –22.7 ± 0.7 –22.0 ± 0.4 –24.2 ± 1.0 –22.4 ± 0.3 –22.6 ± 1.6 –22.0 ± 0.4

Station 2 1–2 cm –24.0 ± 1.0 –25.6 ± 0.2 –22.3 ± 1.4 –25.6 ± 0.8 –22.6 ± 0.2 –25.0 ± 0.6

Total amount / μgacyclic bicyclic tricyclic

Sample H2/PtO2 LiAlH4 H2/PtO2 LiAlH4 H2/PtO2 LiAlH4

Station 1 0–1 cm 0.79 0.28 0.27 0.16 0.31 0.18

Station 1 1–2 cm 0.42 0.16 0.14 0.07 0.15 0.08

Station 2 0–1 cm 0.45 0.34 0.17 0.16 0.23 0.23

Station 2 1–2 cm 0.13 0.05 0.04 0.0 0.06 0.0

Page 70: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

62

Table 2: The bottom water NO3– concentrations and δ 13DIC changes over the incubation time (Δt–t0),

as well as oxygen penetration depths measured after the incubations. In brackets, the average values of ox-ygen penetration depth for the cores before incubation. BG: background, Bic: bicarbonate, Pyr: pyruvate, +Inh: nitrification inhibitor nitrapyrine added.

Station IncubationNO3

– Δδ13DICt–t0 O2 Pen depth

[mM] [‰] [mm]t=0h t=62 h t=144h t=62 h t=144h t=144h

(before: 17)

1

BG 17.9 19.1 4.1 0.6 2.8 11

Bic 15.9 16.2 8.2 66.5 156.2 16

Pyr 18.1 19.2 5.9 109.1 457.0 16

BG+Inh 15.5 16.3 13.9 –0.5 2.6 20

Bic+Inh 11.5 12.0 3.8 135.3 275.5 20

Pyr+Inh 25.4 25.6 14.3 72.7 671.1 16

(before: 13)

2

BG 12.14 11.37 0.09 –0.5 2.9 9

Bic 5.55 4.05 0.07 336.5 606.0 6

Pyr 13.77 12.48 0.12 366.6 1570.7 6

BG+Inh 10.72 10.45 0.19 0.1 3.5 5

Bic+Inh 7.22 6.79 0.09 255.4 570.1 6

Pyr+Inh 5.01 3.17 0.06 511.3 1136.4 16

(before: 16)

3

BG 13.84 0.00 0.13 1.5 –2.3 n.d.

Bic 13.83 12.60 0.08 558.2 814.3 n.d.

Pyr 6.45 4.23 0.12 1108.9 1496.3 n.d.

Glu 14.14 12.03 0.11 90.4 194.2 n.d.

AA 15.34 13.55 0.15 82.7 201.9 n.d.

BG+Inh 16.06 13.39 0.11 0.3 4.9 8

Bic+Inh 15.23 14.05 0.18 244.8 520.3 n.d.

Pyr+Inh 8.92 6.68 0.10 682.7 1644.5 8

Glu+Inh 5.86 4.77 0.03 110.8 301.2 10

AA+Inh 14.36 13.41 0.09 62.2 368.2 n.d.

n.d. - could not be determined.

Page 71: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

63

Material and Methods

cδ C - DIC [‰]13PO [µmol . L ]

NH [µmol . L ] NO [µmol . L ]-1

2

3

NO [µmol . L ]4-1

-1

-13-

4

-

-+O [mg/L]2

0 2 4 6 8 10 12

Sed

imen

t dep

th [c

m]

0 2 4 6 8 10-2

0

2

4

6

8

10

0 10 20 30 40 50

0.0 0.2 0.4 0.6

0 2 4 6 8 10121416

-4 -3 -2 -1 0 1 2

b δ C - DIC [‰]13PO [µmol . L ]

NH [µmol . L ] NO [µmol . L ]-1

2

3

NO [µmol . L ]4-1

-1

-13-

4

-

-+O [mg/L]0 2 4 6 8 10 12

-2

0

2

4

6

8

10

Sed

imen

t dep

th [c

m]

2

0 2 4 6 8 10 12

0 20 40 60 80

0.0 0.2 0.4 0.6

0 4 8 12 16

-4 -3 -2 -1 0 1 2

O [mg/L]2

0 2 4 6 8 10 12

-2

0

2

4

6

8

10

Sed

imen

t dep

th [c

m]

PO [µmol . L ]0 4 8 12 16

NH [µmol . L ] 0 20 40 60 80

NO [µmol . L ]-1

2

3

δ C - DIC [‰]NO [µmol . L ]0.0 0.2 0.4 0.6

0 10 20 30 40

4-1

-1

-13-

4

-

-+

-4 -3 -2 -1 0 1 2

13a

Figure 4: Oxygen microprofiles and pore water results vs. depth at station 1 (a), 2 (b) and 3 (c).

Page 72: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

64

daily injections of a standard mixture of a C20 and a C24 perdeuterated n-alkane. Two replicate analyses (one in the case of the fatty acids) were carried out and the results were averaged in order to obtain the average and standard deviations. The stable carbon isotope compositions are reported in the delta notation against the V-PDB standard. Some values of the biphytanes could not be determined due to low concentrations, mainly in the background cores at station 2. For calculations of incorporation, the background values from station 1 were used.

3.3. Results

3.3.1. Sediment characteristicsThe sediments at stations 1–3 consisted of sand, sandy clay, and dark and soft mud, respectively. Oxygen penetration depth was on average 17 mm at station 1, while oxygen penetrated slightly less deep at station 2 (13 mm on average) and station 3 (15.5 mm on average; Fig. 4). At station 1, organic carbon contents were low with 0.6 % at 0–1 cm depth and 0.3% at 1–2 cm and 6–7 cm depth. Organic carbon contents were higher at station 2 and amounted to 1.4–1.7% and were highest at station 3 (2.0–2.5%). The pH of the bottom water hardly differed between stations and was typically marine, i.e. 8.2 at station 1, 8.0 at station 2, and 7.9 at station 3.

Pore water concentrations of nutrients, as shown in Fig. 4, showed similar trends with depth at the three stations. Inorganic phosphate concentrations were 1 μM in the bottom water and gen-erally increased with depth up to 7–14 μM. Ammonia concentrations increased as well with sed-iment depth, from 0.19 to 1.3 μM in bottom water to 44–70 μM at depth. Nitrite concentrations were an order of magnitude higher in the pore water (0.32–0.62 μM) than in the bottom water (0.055 μM), and increased with sediment depth. Nitrate concentrations were high in the bottom water, 14 μM, and generally decreased quickly with depth in the sediment to values between 0.6 and 3 μM. The δ13C of the DIC showed also similar profiles at the three stations. It decreased from 1‰ in the bottom water to –3‰ in the pore water. At station 2, δ 13C of the DIC could only be measured in the bottom water and the four uppermost pore water samples, but, in these samples, showed the same decrease with depth as observed at Station 1, from –1.3 ‰ to –2.9 ‰.

3.3.2. GDGT-abundanceGDGT-concentrations averaged over all cores from the incubations (10 per station) are shown with standard deviation in order to give an impression of the range of concentrations found (Fig. 5). GDGT-0 and crenarchaeol were present as a core lipid in concentrations of 0.4 to 0.8 μg mL sed–1, and 0.08 to 0.12 μg mL sed–1 for the minor isoprenoid GDGTs (i-GDGTs; GDGT-1, -2 and -3). Concentrations were similar at all stations and did not show substantial changes with sediment depth. IPL-derived GDGTs were present in lower concentrations and amounted to 0.09 to 0.3 (GDGT-0), 0.07 to 0.18 (crenarchaeol) μg mL sed–1. Concentrations were generally lower at station 1 compared to stations 2 and 3. Direct analysis of individual IPL-crenarchaeol showed that the head groups mainly consisted of monohexose (MH)- and hexose, phospho-hexose (HPH)- headgroups, with only a small contribution of dihexose (DH)- head groups (Fig. 5c). The proportion of HPH- to the total peak area of IPL-crenarchaeol was ca. 80–90% in the surface sediments (0–1 and 1–2 cm) of all stations. At station 1 and 2, this decreased to 10–25% in the deepest sediment slice analyzed, 6–7 cm. At station 3, HPH-crenarchaeol dominated at all sediment depths (Fig. 4).

Page 73: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

65

Material and Methods/Results

Tabl

e 3:

Bac

kgro

und

(BG

) val

ues o

fδ 13

C o

f th

e bip

hyta

nes (

BP) m

easu

red

in th

e sed

imen

t (i.e

. δ13

C o

f th

e bi

phyt

anes

that

wer

e inc

ubat

ed ju

st w

ith w

ater

, bu

t und

er th

e sa

me

cond

ition

s). T

hese

val

ues

wer

e us

ed fo

r de

term

inat

ion

of th

e Δ13

C: B

G v

alue

s w

ere

subt

ract

ed fr

om th

e 13

C o

f th

e co

rres

pond

ing

valu

e m

easu

red

from

the

core

s inc

ubat

ed w

ith la

bel (

Fig.

6). N

umbe

rs in

dica

te th

e de

pth

inte

rval

the

sam

ple

was

com

ing

from

, + :

nitri

ficat

ion

inhi

bito

r ni

trapy

rine

adde

d fo

r inc

ubat

ion,

- : n

o in

hibi

tor a

dded

. Val

ues a

t sta

tion

2 co

uld

not b

e de

term

ined

due

to lo

w c

once

ntra

tions

.

Stat

ion

1St

atio

n 2

Stat

ion

3

δ13C

[‰]

δ13C

[‰]

δ13C

[‰]

Sedi

men

t(c

m)

acyc

lic

BP

bicy

clic

B

Ptr

icyc

lic

BP

acyc

lic

BP

bicy

clic

B

Ptr

icyc

lic

BP

acyc

lic

BP

bicy

clic

B

Ptr

icyc

lic

BP

0–1

–22.

0 ±

0.1

–20.

5 ±

1.1

–19.

9 ±

0.3

n.d.

*n.

d. *

n.d.

*–2

0.4

± 0

.2–2

0.2

± 1

.2–1

9.8

± 0

.8

1–2

–21.

4 ±

0.5

–20.

7 ±

0.1

–20.

2 ±

0.0

n.d.

*n.

d. *

n.d.

*–1

9.6

± 1

.2–1

8.3

± 1

.4–1

7.7

± 1

.9

6–7

–21.

4 ±

0.2

–22.

4 ±

1.9

–21.

5 ±

2.3

n.d.

*n.

d. *

n.d.

*–2

0.6

± 0

.4–2

0.1

± 0

.4–1

9.2

± 0

.6

+

0–1

n.d.

*n.

d.*

n.d.

*n.

d. *

n.d.

*n.

d. *

–19.

3 ±

0.6

–19.

7 ±

0.6

–19.

4 ±

0.3

1–2

–22.

0 ±

0–2

1.2

± 0

.4–2

1.3

± 0

–22.

9 ±

0.9

–21.

4 ±

1.4

–20.

3 ±

0.9

–19.

4 ±

0.1

–19.

9 ±

0.4

–19.

2 ±

0.2

6–7

–22.

4 ±

0

–20.

0 ±

0.6

–22.

2 ±

0–2

2.0

± 0

.0–2

1.0

± 0

.4–2

0.0

± 0

.4–2

2.4

± 0

.0–2

1.4

± 0

.4–2

0.7

± 0

.1

* n.d

. – c

ould

not

be

dete

rmin

ed.

Page 74: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

66

0-1 cm 1-2 cm 6-7 cm0-1 cm 1-2 cm 6-7 cm0-1 cm 1-2 cm 6-7 cm

0-1 cm 1-2 cm 6-7 cm

0-1 cm 1-2 cm 6-7 cm

0-1 cm 1-2 cm 6-7 cm0-1 cm 1-2 cm 6-7 cm

0-1 cm 1-2 cm 6-7 cm0-1 cm 1-2 cm 6-7 cm

IPL-

cren

arch

aeol

[%]

IPL-

GD

GT

conc

entra

tion

[µg

(mL

sed)

]

-1C

L-G

DG

T co

ncen

tratio

n[µ

g (m

L se

d)

]

Depth interval

-1

0.0

0.1

0.2

0.3

0.4

HPH-crenarchaeol

MH-crenarchaeolDH-crenarchaeol

GDGT-0 Crenarchaeol

Station 3Station 2Station 1

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0

20

40

60

80

100c

b

a

Figure 5: GDGT-0 and crenarchaeol concentrations (averages), for CL-GDGTs (a) and IPL-derived GDGTs (b) at all three stations and core depths. Proportions of head groups of IPL-crenarchaeol at all three stations (c).

Page 75: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

67

Results

3.3.3. Incubation experimentsIncubations experiments were started directly after core recovery and initial sampling by injec-tion of solutions, which contained 13C-bicarbonate, pyruvate, glucose , amino acids or just water (incubated for comparison and further called the “background core”) over the first 7 cm of the cores. The cores were incubated at in situ temperatures (4°C) for 6 days (4 days for station 3) and subsequently sliced and stored for analysis of 13C-enrichment in lipids. Oxygen penetration depths were measured and bottom water was sampled at the start of the experiment, after 48 hours (36 for Station 3) and at the end of the incubations. At all stations, nutrient concentrations of the bottom water did not change over the 6 days (5 days for station 3) of incubation, except for nitrate, which generally showed a decrease (Table 2). The bottom water pH remained mostly constant, except at station 1, where it changed during the incubation from 8.2 to 8.0 in all incu-bations, similar to the other sediment cores and perhaps indicating that the initial value of 8.2 might have been due to an error in the measurement. The δ13C of the DIC in the bottom water substantially increased in the cores where 13C-labeled compounds and bicarbonate was added, but not in the background cores (Table 2). In cores from station 2, oxygen penetration depths decreased to 6–9 mm in all incubation experiments except for the core incubated with pyruvate and inhibitor, as well as in cores from station 3, where a decrease in oxygen penetration depth to 8–10 mm was observed (Table 2). In all cores from the three stations, no major differences in IPL-crenarchaeol concentrations upon incubation could be detected (Figs. S1–S3).

The δ13C-values of the biphytanes released from the IPL-GDGTs of the background cores were ranging from –17 to –24‰ at all stations (Table 3), which is within the natural variation observed for biphytanes sourced by Thaumarchaeota in marine sediments (Hoefs et al., 1997; Könneke et al., 2012; Schouten et al., 2013). No substantial 13C enrichment of biphytanes was observed in nearly all of the incubation experiments (Fig. 6). The only exception was the tricy-clic biphytane in some of the incubations at station 1 (with 13C-labelled substrates bicarbonate, bicarbonate and inhibitor, and pyruvate and inhibitor at nearly all depths) and one incubation (with pyruvate and inhibitor at 6–7 cm depth) at station 2 with enrichments ranging from 2 to 4‰. However, due to the large standard deviations in the δ13C values of the background cores, only the enrichment of 2–5 ‰ in the 1–2 cm slices of station 1 is statistically significant at a 95% confidence level (t(2)= 6.6–9.0, P=0.05). To determine whether the added 13C-labelled substrate was incorporated into sedimentary microbial biomass at all, phospholipid-derived fatty

Table 4: δ13C values of fatty acids measured in the 0-1 cm slice in cores from station 3, incubated with-out inhibitor, for 5 days.

δ13C [‰]Fatty acid BG Bic Glu AA

C16:0 –25.5 ± 0.2 –25.3 ± 0.6 20.6 ± 0.5 30.0 ± 3.9

C18:0 –24.6* –24.5 ± 0.2 66.2 ± 0.4 63.1 ± 2.1

* Single measurement

Page 76: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

68

0-1* 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7

∆ δ

C

[‰

]

-5

0

5

Acyclic biphytane

Bicyclic biphytaneTricyclic biphytane

13in

c-B

G

Depth [cm]

Experiment

0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7* 0-1 1-2* 6-7

0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7

0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7 0-1 1-2 6-7

∆ δ

C

[‰

]13

inc-

BG

Pyruvate + InhPyruvateBicarbonate + InhBicarbonate

Pyruvate + InhPyruvateBicarbonate + InhBicarbonate

Pyruvate + InhPyruvateBicarbonate + InhBicarbonate

Depth [cm]

Depth [cm]-5

0

5

AA + InhAAGlucose + InhGlucose

∆ δ

C

[‰

]13

inc-

BG

-5

0

5

∆ δ

C

[‰

]13

inc-

BG

Depth [cm]

-5

0

5(d) Station 3

(c) Station 3

(b) Station 2

(a) Station 1

Figure 6: Difference in δ13C-values (Δδ13Cinc-bg) of the acyclic, bicyclic and tricyclic biphytanes in ‰ in the cores incubated with labeled substrate versus the 13C-values in the background cores (incubated with and without inhibitor) at station 1 (a), station 2 (b) and station 3 (c) from the cores incubated with bicarbonate and pyruvate, with and without inhibitor, and the cores incubated with glucose and amino acids with and without inhibitor, at station 3 (d).

Page 77: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

69

Results/Discussion

acids, stemming from bacteria or eukaryotes, were analyzed from the surface sediments of the bicarbonate, amino acids and glucose incubation experiments. δ13C values of the most common fatty acids, C16:0 and C18:0 were enriched in 13C by up to 80‰ in the amino acid and glucose incu-bation experiments compared to the background (Table 4). The cores incubated with 13C-labeled bicarbonate that were analyzed showed no 13C-enriched fatty acids.

3.4. Discussion

3.4.1. Biomarkers indicative for live sedimentary ThaumarchaeotaThe only biomarker lipids up to now which are probably indicative for live Thaumarchaeota are crenarchaeol IPLs (Pitcher et al., 2011b). Of these, the ones with a glycosidic head group likely contain a substantial fossil component, while the hexose, phosphohexose (HPH)-crenarchaeol is more labile and its concentration corresponds well to DNA-based thaumarchaeal abundance (Pitcher 2011a and 2011b; Schouten et al., 2012; Lengger et al., 2012a). The presence and dom-inance of HPH-crenarchaeol in the sediments strongly suggests that live Thaumarchaeota were present in sediments at all the stations (Fig. 5). This was expected for the surface sediment, as oxygen and ammonium were present (Martens-Habbena et al., 2009; Erguder et al., 2009; Hat-zenpichler, 2012). The ammonium concentration profiles, with concentrations decreasing up-wards, agree with aerobic ammonia oxidation proceeding in the oxygenated parts, using the am-monium diffusing from below. However, HPH-crenarchaeol was present down to 7 cm depth, below the oxygen penetration depth, which could indicate that either live Thaumarchaeota were present at depth, using a metabolic process different from aerobic ammonia oxidation, or that this HPH-crenarchaeol is not stemming from live Thaumarchaeota at depth and was partly preserved. In any case, it is not yet known what exactly the preferred environment for Thau-marchaeota is, as they have been found to be active in low and high ammonia and low and high oxygen environments (Martens-Habbena et al., 2009; Erguder et al., 2009; Hatzenpichler, 2012).

3.4.2. Activity indicators and label uptake during incubationSeveral parameters can be monitored during the incubation experiments to cast a light on sed-imentary microbial activities stimulated by the addition of substrates. During incubation, a de-crease of oxygen penetration depths would indicate an increased aerobic metabolism at the surface, utilizing oxygen and preventing it from penetrating deeper, which was expected in the slices where organic carbon was added. However, at station 1, this was not observed and oxygen penetration depths only decreased in the background core (Table 2). This could be due to the aeration carried out by bubbling air into the bottom water of the core, and oxygen penetrating deeper into the sandy sediment. However, the cores from station 2 showed a decrease of oxygen penetration depth after incubation, except for the core incubated with pyruvate and inhibitor. Also the cores from station 3, i.e. those where oxygen penetration depth could be measured, showed a decrease, indicating that the oxygen was consumed upon incubation, independent of the type of substrate added to the cores. Ammonia and nitrite concentrations in the bottom water showed no changes during the incubations. No significant differences can be seen if the cores incubated with and without nitrification inhibitor are compared, indicating that either the communities are not affected by it, or potentially that the inhibitor was not as effective in a sedimentary environment as in water incubations (e.g. due to adherence to particles). Nitrate concentrations were, at station 2 and station 3, below detection limit after 6 and 5 days, respec-

Page 78: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

70

tively, suggesting consumption of nitrate (Table 2). This may indicate a strong contribution of denitrification to organic carbon processing in these sediments in the sections below oxygen penetration depth. The increase in δ13C of the DIC in the bottom water of all cores compared to the background cores (Table 2) clearly showed that the organic substrates added were respired. Previous labeling experiments with similar substrates have been successful in detecting uptake of 13C-label into bacterial PLFA (e.g. Guilini et al., 2010), resulting in a highly labeled C16:0, among other PLFAs. Indeed, the strong labeling of the bacterial PLFA C16:0 at station 3 indicated that the organic substrates were readily taken up by bacteria, and potentially non-photosynthetic eukaryotes (PLFA C16:0, C18:0), and incorporated into their biomass. Intriguingly, no incorporation was observed upon labeling with bicarbonate, which is in contrast to other studies with oligotro-phic sediment cores who used similar concentrations (e.g. Guilini et al., 2010). The amount of PLFA generated by chemoautotrophs present might have been too low to allow unambiguous detection of the incorporation in this case.

The most abundant archaeal lipids were crenarchaeol, a lipid specific for Thaumarchaeota (Sinninghe Damsté et al., 2002; de la Torre et al., 2008; Schouten et al., 2008; Pitcher et al., 2011b; Sinninghe Damsté et al., 2012), and GDGT-0, a lipid produced by Thaumarchaeota as well as some other Archaea (cf. Koga and Nakano, 2008). Their concentrations did not substantially change upon incubation, suggesting that Thaumarchaeota abundances did not change (Fig. 5 showing an average of all incubated cores, with standard deviations). However, this could also indicate that the majority of IPLs in these sediments were fossil (cf. Schouten et al., 2010; Logemann et al., 2011; Lengger et al., 2012a) and not part of living Thaumarchaeota, and thus not affected at all by the incubation. This could include HPH-GDGTs as well, though it is also possible that the bias during analysis, i.e. loss during silica column chromatography (cf. Lengger et al., 2012b), resulted in an underrepresentation of HPH- derived GDGTs.

The biphytanes analyzed included a-, bi- and tricyclic biphytane, which are mainly derived from the two most abundant GDGTs: GDGT-0 (acyclic; general archaea), and crenarchaeol (bi- and tricyclic; Thaumarchaeota). Importantly, in nearly all of the incubation experiments, the δ13C values of all measured biphytanes, including those derived from crenarchaeol, changed by less than 2‰ and sometimes even decreased in 13C-content compared to the background val-ues (Fig. 6). In some of the experiments, mainly at station 1, a slight enrichment in the tricyclic biphytane, derived from crenarchaeol, of up to 4 ± 2‰ was observed though this difference is not statistically significant for individual experiments. Furthermore, the bicyclic biphytane, also predominantly derived from crenarchaeol, shows, in general, lower 13C-enrichments than the tricyclic biphytane and also no statistically significant enrichments for individual experiments. Thus, there is no evidence for substantial uptake of 13C-label in archaeal GDGTs, contrary to what is observed for the PLFA.

This apparent lack of uptake could be due to three reasons. Firstly, it is possible that the Thau-marchaeota were active, but the substrates were not taken up. As the metabolism of the different Thaumarchaeota could be diverse, this is possible, but, in case of the bicarbonate incubations, unlikely, as in sea water incubations as well as enrichment cultures – including those of sedimen-tary Thaumarchaeota – 13C from bicarbonate has been shown to be readily incorporated into biphytanes (Wuchter et al., 2003; Park et al., 2010; Pitcher et al., 2011c). Secondly, it is possible that most of the IPLs measured (i.e. MH- and DH-GDGTs, assuming the chromatographic discrimination against the HPH-GDGTs) are fossil, and thus only a small amount of these IPLs

Page 79: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

71

Discussion

are part of live and active sedimentary Thaumarchaeota. Finally, it is also possible that a higher proportion of the IPLs are part of sedimentary Thaumarchaeota, but that their metabolism and growth rates are so slow (as opposed to Thaumarchaeota in sea water and enrichment cultures), that no incorporation could be detected over the time scales investigated. Thus, our incubation times of 96–144 h may have been too short. However, longer incubation times inevitably result in spreading of the label over the whole sedimentary food web as e.g. discussed by Radajewski et al. (2003) for stable isotope probing of nucleic acids. This is why this study, and nearly all other incubation studies (e.g. Boschker et al. 1998, Guilini et al. 2010, cf. Middelburg 2013 and references cited therein), including those for sediments (e.g. Middelburg et al. 2000, Moodley et al. 2002), use only a short time interval (mostly just a few days) to avoid this label scrambling. Indeed, our results show that the uptake in bacteria is occurring within a few days suggesting that, if active, microbes rapidly take up label.

3.4.3. Estimating growth rates of Thaumarchaeota and turnover rates of IPL-GDGTsTo investigate whether growth rates as well as degradation rates of IPL-GDGTs were respon-sible for the apparent lack of substantial 13C-labeling, we used data from our experiments to estimate minimal growth rates and the turnover time of IPL-GDGTs. The greatest label in-corporation was found at station 1 in the sandy sediment, indicating that Thaumarchaeota may have been active there. If all IPL-GDGT biphytanes are stemming from living cells, and an enrichment of 2‰ of biphytanes is assumed, this can be converted to growth rates by using a formula used for growth of phytoplankton based on the isotope dilution theory of Laws (1984; Eq. (1)), with µ representing the growth rate, tinc the incubation time, P* the atom%-excess of the product, in our case the biphytane, and A* the atom%-excess of the substrate added (cf. alkenone lipids for eustigmatophyte algae; Popp et al., 2006):

(1)

Unfortunately, as natural concentrations of the added substrates were not determined, A* could not be calculated for the incubation experiments with pyruvate, glucose and amino acids. In the bicarbonate experiments, we calculated A* based on the amount of added 99%-labeled DIC and natural δ13C and concentration of DIC (Table 2). For P* we used the highest enrich-ment observed at 6–7 depth with no inhibitor, i.e. 4‰. This gives an estimate of the maximum µ as 6.3 × 10–4 d–1, corresponding to a generation time of 3 yr. Clearly, for the other bicarbon-ate experiments which resulted in even less label incorporation, generation times will be much higher (e.g. for a 2 ‰ enrichment, 9 yr). Interestingly, growth rates as low as estimated here have been predicted by Valentine (2007), who stated that the archaeal domain, other than bacteria and eukaryotes, has adapted to low energy conditions and is thus characterized by the exceptional ability to live in low energy environments. They are thus not fast in adapting to settings with higher concentrations of available carbon and redox substrates. However, this has not been observed for pelagic Thaumarchaeota as SIP experiments showed a substantial enrichment in 13C in IPL-GDGTs (Wuchter et al., 2003; Pitcher et al., 2011c) in incubation experiments of 168 and 24 h duration, respectively. Indeed, the generation times calculated from enriched North Sea water labeling experiments conducted by Wuchter et al. (2003; A*=5%; P*=0.46% – 400 ‰ enrichment) and Pitcher et al. (2011c; A*=9%; P*=0.07% – 44‰ enrichment) were at least an order of magnitude lower than observed here – 50 and 88 d, respectively. Also sedimentary

Page 80: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

72

Thaumarchaeota have been shown to incorporate 13C from bicarbonate in enrichment cultures (Park et al., 2010). The growth rates reported here might be an underestimation if a large pro-portion of archaeal cells is fossil; and the use of lipid turnover times might be more suitable, as discussed below. Slow growth could potentially be due to the experiments not being carried out in situ on the sea floor. However, incubations were carried out in whole, intact sediment cores, in the dark, immediately after recovery, and at in situ temperatures in order to minimize the devia-tions from in situ conditions, although, admittedly, the pressure was at atmospheric conditions. Therefore, this is unlikely to explain the low growth rates.

The calculations above assume exponential growth of sedimentary Thaumarchaeota and an immediate translation of this signal into the lipid pool. If, on the other hand, steady state con-ditions are assumed, i.e. cell numbers and lipid concentrations are not increasing but remain constant, the turnover time for the sedimentary lipid pool tto can be calculated according to Lin et al. (2012) :

*

*

PtAt inc

to⋅

= (2)

Using Eq. (2), the turnover time of IPL-crenarchaeol in sediments would thus be at least ~4 yrs based on the 4‰ enrichment of the tricyclic biphytane in the bicarbonate incubation experi-ment. Since labeling was not detected in the other sediments, this turnover time estimate is likely to be a minimal estimate. This large turnover time is not an unusual postulate for thaumarchaeal lipids in sediments, as also Lin et al. (2012) found very low 13C incorporation rates of labeled glucose into sugar moieties of glycosidic GDGTs over >400 d in slurry incubations of subsur-face sediments and used these to estimate turnover times of diglycosyl-lipids of 1,700–20,500 years. Similarly, recently published degradation experiments by Xie et al. (2013) using 14C-labeled IPL-diether lipids found very high turnover times and estimated half-lives of up to 310 kyr un-der energy-deprived subsurface conditions. This is in strong contrast to GDGTs in the water column: Using the SIP data of Wuchter et al. (2003) and Pitcher et al. (2011c) from incubations of North Sea water, we obtained tto of 76 and 128 d, respectively, for the thaumarchaeal GDGTs.

Our results are similar to those obtained by in situ-incubations carried out by Takano et al. (2010) and Nomaki et al. (2011), i.e. they did not observe significant incorporation after 405 days into the biphytanyl chains either (even though one single data point after 9 days does show a 10‰ enrichment), though they did observe incorporation into the glycerol moiety. A possibility, as suggested earlier, is that biphytanyl chains are not or hardly produced by sedimentary Thau-marchaeota, but recycled, and the sugar headgroups and/or the glycerol are only newly gener-ated and 13C-labeled. In this case, isotope ratios of whole ether lipids instead of the biphytanes would enable detection of the label, but this involves laborious clean-up of the isolated ether lipids (Takano et al., 2010). However, as also the incorporation into the sugar headgroups of GDGTs, as reported by Lin et al. (2012), is similarly low, it suggests that there is indeed a lack of, or only a very slow 13C-uptake into IPL-GDGTs.

It seems that stable carbon isotope probing using biphytanes is most likely not the method of choice for thaumarchaeal activity measurements in sediments, where conditions are favorable for preservation of fossil GDGTs, thus creating a large background signal. In the water column, however, where growth of Thaumarchaeota is faster and lipids are likely degraded on shorter time scales, due to the abundant presence of oxygen and lack of mineral matrix protection (Keil

Page 81: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

73

Discussion/Conclusion

et al., 1994; Hedges and Keil, 1995), stable isotope incubations are useful for determining ac-tivities of Thaumarchaeota. In sediments that contain a high proportion of fossil IPL-GDGTs which are not part of live Archaea, the incorporation of 13C into the lipids of the few active Thaumarchaeota is probably not enough to change the δ13C of the IPL-GDGTs to a significant degree over the commonly used time scales of SIP experiments.

3.5. ConclusionIncubations of sediment cores from the Iceland Shelf were used in order to determine the me-tabolism of sedimentary Thaumarchaeota. Thaumarchaeal IPLs were present in sediment cores recovered from the Iceland Shelf and the presence of labile IPL-biomarkers for Thaumarchae-ota, such as HPH-crenarchaeol, suggested the presence of living Thaumarchaeota. However, incubations with 13C-labeled substrates bicarbonate, pyruvate, glucose and amino acids did not result in any substantial incorporation of the 13C into the biphytanyl chains of these lipids. Turn-over times and generation times of thaumarchaeal lipids were estimated to be at least several years, suggesting slow growth of sedimentary Thaumarchaeota in contrast to other sedimentary organisms, and/or a slow degradation of IPL-GDGTs, in contrast to bacterial or eukaryotic PLFAs.

Acknowledgements

We thank Dr. Y. Takano and one anonymous reviewer for their useful comments which greatly improved the manuscript. The authors would like to thank the Master and crew of the R/V Pelagia and the participants of the Long Chain Diols Cruise 64PE341, in particular M. Baas, S. Rampen, M. Besseling, L. Handley and W. Lenthe, as well as the NIOZ technical department (Y. Witte, J. Schilling). We would also like to thank L. Moodley, H.T.S. Boschker, L. Pozzato, J. Mid-delburg, G. Duineveld and M. Lavaleye for helpful advice in planning the experiments and K. Koho / Utrecht University and L. Stal and N. Bale for providing equipment. We are grateful for analytical assistance to J. van Ooijen, S. Crayford, A. Mets, M. Verweij and J. Ossebaar. S.K.L. was funded by a studentship from the Darwin Center for Biogeosciences to S. S. This is a publication nr. DW-2013-1007 of the Darwin Center for Biogeosciences.

Page 82: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

74

Figure S1: IPL-crenarchaeol profiles of the incubated cores (station 1) - background (BG), bicarbonate (Bic) and pyruvate (Pyr) without (left plot) and with nitrification inhibitor (right plot). Top graphs shows the values in SRM area . mL–1 from 0–1 cm depth, graphs in the middle the profiles from 1–2 cm depth and at the bottom the profiles from 6–7 cm depth

7

6

5

4

10

10

10

10

7

6

5

4

10

10

10

10

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

IPL

abun

danc

e[S

RM

are

a . m

L ]-1

IPL

abun

danc

e[S

RM

are

a . m

L ]-1

IPL

abun

danc

e[S

RM

are

a . m

L ]

7

6

5

4

10

10

10

10

-1

Pyr + InhPyr Bic + InhBic BG + InhBG

6-7 cm

1-2 cm

0-1 cm

Hexose, phosphohexose-crenarchaeolDihexose-crenarchaeolMonohexose-crenarchaeol

Page 83: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

75

Supplementary Material

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

MH DH HPHMH DH HPH MH DH HPHMH DH HPH MH DH HPHMH DH HPH

IPL

abun

danc

e[S

RM

are

a . m

L ]-1

IPL

abun

danc

e[S

RM

are

a . m

L ]-1

IPL

abun

danc

e[S

RM

are

a . m

L ]

5

4

3

10

10

10

5

4

3

10

10

10

5

4

3

10

10

10

-1

Pyr + InhPyr Bic + InhBic BG + InhBG

6-7 cm

1-2 cm

0-1 cm

Hexose, phosphohexose-crenarchaeolDihexose-crenarchaeolMonohexose-crenarchaeol

Figure S2: IPL-crenarchaeol profiles of the incubated cores (station 2) - background (BG), bicarbonate (Bic) and pyruvate (Pyr) without (left plot) and with nitrification inhibitor (right plot). Top graphs shows the values in SRM area. mL–1 from 0–1 cm depth, graphs in the middle the profiles from 1–2 cm depth and at the bottom the profiles from 6–7 cm depth.

Page 84: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 3

76

Chapter 3 Supplementary Material

MH

DH

HP

H

MH

DH

HP

H

MH

DH

HP

H

AA

+Inh

MH

DH

HP

H

MH

DH

HP

H

MH

DH

HP

H

Glu

+Inh

MH

DH

HP

H

MH

DH

HP

H

MH

DH

HP

H

AA

MH

DH

HP

H

MH

DH

HP

H

MH

DH

HP

H

Glu

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

MH

DH

HP

HM

HD

HH

PH

IPL abundance[SRM area . mL ] -1

IPL abundance[SRM area . mL ] -1

IPL abundance[SRM area . mL ] -1

Pyr +

Inh

Pyr

Bic

+ In

hB

icB

G +

Inh

BG

6-7

cm

1-2

cm

0-1

cm 7 6 5 4

10 10 10 10

810

7 6 5 4

10 10 10 10

810

7 6 5 4

10 10 10 10

810

Hex

ose,

pho

spho

hexo

se-c

rena

rcha

eol

Dih

exos

e-cr

enar

chae

olM

onoh

exos

e-cr

enar

chae

ol

Figu

re S

3: I

PL-c

rena

rcha

eol p

rofil

es o

f th

e in

cuba

ted

core

s (s

tatio

n 3)

- ba

ckgr

ound

(BG

), bi

carb

onat

e (B

ic),

pyru

vate

(Pyr

), gl

ucos

e (G

lu) a

nd a

min

o ac

ids

(AA

) with

out (

left

plot

) and

with

nitr

ifica

tion

inhi

bito

r (rig

ht p

lot).

Top

gra

phs

show

s th

e va

lues

in S

RM a

rea.

mL–1

from

0–1

cm

dep

th, g

raph

s in

th

e m

iddl

e th

e pr

ofile

s fro

m 1

–2 c

m d

epth

and

at t

he b

otto

m th

e pr

ofile

s fro

m 6

–7 c

m d

epth

.

Page 85: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 86: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 87: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

79

4. Abundance and diversity of denitrifying and anammox bacteria in seasonally hypoxic and sulfidic sediments of the saline Lake Grevelingen

Yvonne A. Lipsewers, Ellen C. Hopmans, Filip J.R. Meysman, Jaap S. Sinninghe Damsté, and Laura Villanueva

Denitrifying and anammox bacteria are involved in the nitrogen cycling in marine sediments but the environmental factors that regulate the relative importance of these processes are not well constrained. Here, we evaluated the abundance, diversity and potential activity of denitri-fying, anammox, and sulfide-dependent denitrifying bacteria in the sediments of the seasonally hypoxic saline Lake Grevelingen, known to harbor an active microbial community involved in sulfur oxidation pathways. Depth distributions of 16S rRNA gene, nirS gene of denitrifying and anammox bacteria, aprA gene of sulfur-oxidizing and sulfate-reducing bacteria, and ladderane lipids of anammox bacteria were studied in sediments impacted by seasonally hypoxic bottom waters. Samples were collected down to 5 cm depth (1 cm resolution) at three different locations before (March) and during summer hypoxia (August). The abundance of denitrifying bacteria did not vary despite of differences in oxygen and sulfide availability in the sediments, whereas anammox bacteria were more abundant in the summer hypoxia but in those sediments with lower sulfide concentrations. The potential activity of denitrifying and anammox bacteria as well as of sulfur-oxidizing, including sulfide-dependent denitrifiers and sulfate-reducing bacteria, was potentially inhibited by the competition for nitrate and nitrite with cable and/or Beggiatoa-like bacteria in March and by the accumulation of sulfide in the summer hypoxia. The simultaneous presence and activity of organoheterotrophic denitrifying bacteria, sulfide-dependent denitrifi-ers and anammox bacteria suggests a tight network of bacteria coupling carbon-, nitrogen- and sulfur cycling in Lake Grevelingen sediments.

Frontiers in Microbiology 7, 1661, 2016

Page 88: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

80

4.1. IntroductionNitrogen availability is a major factor controlling primary production in temperate coastal ma-rine environments with high anthropogenic nitrogen input (Herbert, 1999). Denitrification is a key process in the nitrogen cycle of coastal sediments releasing gaseous end products, nitric ox-ide (NO), nitrous oxide (N2O) and dinitrogen gas (N2) to the atmosphere. This nitrogen removal can result in a decrease of nitrogen availability for primary producers and thereby controlling the rate of eutrophication in coastal marine systems (Seitzinger 1988; Herbert 1999). Multiple microbial mediated pathways result in the removal of nitrogen in anoxic sediments (see Devol, 2015 for a detailed review). Here we focus on (1) denitrification, the stepwise conversion of nitrate/nitrite to dinitrogen gas which is mainly performed by facultative organoheterotrophic anaerobic bacteria and some archaea (Zumft, 1997); (2) anaerobic ammonium oxidation (ana-mmox), the oxidation of ammonium with nitrite to dinitrogen gas carried out by anammox bacteria (Kuypers et al., 2003); and (3) sulfide-dependent denitrification, the oxidation of sulfide with nitrate performed by autotrophic members of α-, β-, γ- and ε-proteobacteria, which could contribute to denitrification and to the removal of sulfide in the oxygen transition zone of coast-al marine sediments (Shao et al., 2010).

Numerous environmental factors, such as the availability of nitrogen speciation and concen-tration, temperature, oxygen concentrations, organic matter quality and quantity, bioturbation and other sediment characteristics have been suggested to affect the distribution and abundance of denitrifying and anammox bacteria (Thamdrup and Dalsgaard 2002; Meyer et al., 2005; Jen-sen et al., 2008; Dang et al., 2010; Laverock et al., 2013; Prokopenko et al., 2013; Babbin et al., 2014; Zhang et al., 2014). In this study, we assessed the effects of hypoxia and elevated sulfide concentration on the abundance and activity of denitrifiers and anammox bacteria in marine sediments as these factors have been previously suggested as potential limiting factors in denitri-fication processes (Brunet and Garcia-Gil, 1996; Burgin and Hamilton, 2007; Aelion and Wart-tinger, 2010; Neubacher et al., 2011; 2013; Bowles et al., 2012). Seasonal hypoxia is an increasing phenomenon that occurs in coastal areas causing a decrease in the electron acceptors (O2, NO3

–) in the bottom waters (Diaz & Rosenberg 2008). Besides, sulfide inhibition can decrease denitri-fication rates as the enzyme catalyzing the reduction of nitrous oxide to N2 is sensitive to sulfide (e.g. Porubsky et al., 2009). In addition, several studies have provided putative evidence indicat-ing that sulfide inhibits the anammox reaction. For example, Dalsgaard et al. (2003) reported a decrease in anammox activity in the sulfidic waters of the anoxic basin of Golfo Dulce (Costa Rica). Also Jensen et al. (2008) showed that sulfide had a direct inhibiting effect on the activity of anammox bacteria in the Black Sea.

In this study, we evaluated the impact of environmental factors, such as oxygen and free sulfide concentrations in the diversity, abundance, activity, and spatial distribution of bacteria involved in N2 removal pathways in sediments of Lake Grevelingen (The Netherlands), a sea-sonally hypoxic saline reservoir. Here, we determined the diversity, abundance and potential activity of denitrifying bacteria, anammox bacteria and sulfur-oxidizing bacteria (SOB), includ-ing sulfide-dependent denitrifiers, and sulfate-reducing (SRB), in three different stations within the lake both in March (before hypoxia) and August (i.e. during hypoxia). In order to determine changes in the diversity, abundance and activity of denitrifiers, we targeted the nirS gene, en-coding for the cytochrome cd1 nitrite reductase, catalyzing nitrite reduction to nitric oxide (NO) (Braker et al., 1998; Smith et al., 2007; Huang et al., 2011). Here, we focused on the cytochrome

Page 89: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

81

Introduction/Material and Methods

cd1-containing nitrite reductase (nirS gene), as it has been found to be more widespread in the bacterial communities compared to the copper-containing nitrite reductase (nirK) in various sed-iments (Braker 1998; Priemé et al. 2002; Liu et al. 2003; Throbäck et al. 2004; Tiquia et al. 2006; Oakley et al. 2007; Dang et al. 2009; Huang et al. 2011). For anammox bacteria, the nirS gene was also quantified as it has been recently suggested to be a functional biomarker anammox bacteria (Li et al., 2011). Both, denitrifying and anammox bacteria harbor one copy of the nirS gene, indicating that the nirS gene might be a suitable marker to compare the abundance and distribu-tion of nirS-type denitrifiers and anammox bacteria in coastal sediments. The potential activity of denitrifying and anammox bacteria was also determined by estimating the gene expression of those metabolic genes as in previous studies (Smith et al., 2007; Lam et al., 2009; Bale et al., 2014; Bowen et al., 2014; Zhang et al., 2014; Chapter 2). Besides functional genes, the abundance of anammox bacteria was also determined by the quantification of ladderane lipids, which are specific lipid biomarkers for this microbial group (Sinninghe Damsté et al., 2002; Jaeschke et al., 2009; Russ et al., 2013). Finally, the diversity and abundance of sulfur-oxidizing and sulfate-re-ducing bacteria in marine sediments was estimated by targeting the aprA gene encoding the adenosine-5’-phosphosulfate (APS) reductase (Lenk et al., 2011; Blazejak and Schippers, 2011; Dyksma et al., 2016). The APS reductase is operating in in SOB, oxidizing sulfite to APS, as well as in SRB, operating in reverse direction, converting APS to adenosine monophosphate (AMP) and sulfite (SO3

2–) (Meyer and Kuever, 2007b). Our starting hypothesis is that the abundance and potential activity of organoheterotrophic

denitrifiers and anammox bacteria would decrease upon increase of the sulfide concentration found in the sediments during the summer hypoxia. Moreover, recent studies point out that sulfide-dependent denitrifiers play a relevant role in the sulfide transition zone of intertidal sedi-ments (Dyksma et al., 2016). Therefore, we hypothesize that their abundance and activity would be higher in hypoxic and sulfidic conditions, thus contributing more to the general N2 removal in comparison to organoheterotrophic denitrifying and anammox bacteria and eventually being involved in sulfide detoxification which could promote anammox activity.

4.2. Material and Methods

4.2.1. Study site and sampling Lake Grevelingen is a former estuary within the Scheld- Rhine- Meuse delta, and was formed by the construction of the Grevelingendam on the landside in 1964 and the Brouwersdam on the seaward side in 1971. The lake (surface area: 108 km2, mean water depth 5.3 m) mainly consists of shallow water areas with the exception of the former tidal gullies, that have a water depth of up to 48 m (Keldermann 1984; Nienhuis and De Bree, 1984). After its closure, the lake transformed into a freshwater body, but in 1979, the connection with the North Sea was par-tially re-established, and since then, Lake Grevelingen has a high and relatively constant salinity (29−32). Within Lake Grevelingen, the Den Osse basin forms a deeper basin within the main gully, and experiences a regular seasonal stratification leading to oxygen depletion in the bottom water (summer hypoxia) (Wetstejn, 2011). Due to sediment focusing, the Den Osse basin (max-imum water depth 34 m) also experiences a rapid accumulation of fine-grained, organic rich sediments (sediment accumulation rate ~2 cm yr–1; Donders, 2012).

Sediment cores were collected along a depth gradient in Den Osse basin during two cruises March and August 2012 on board of the RV Luctor. Sampling took place at three different sta-

Page 90: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

82

tions: S1 was located in the deepest point of the basin at 34 m water depth (51.747°N, 3.890°E), S2 at 23 m (51.749°N, 3.897°E) and S3 at 17 m (51.747°N, 3.898°E). Sediment was collected with single core gravity corer (UWITEC) using transparent PVC core liners (6 cm inner diam-eter, 60 cm length). Four sediment cores were collected at each station in March and in August. The cores were sliced with a 1 cm resolution until 5 cm depth, and sediment samples were collected for lipid and DNA/RNA analysis and kept at –80ºC until further processing. In each sampling campaign, a water column depth profile of temperature, salinity and oxygen (O2) con-centration was recorded at S1 using an YSI 6600 CTD instrument (for details see Hagens et al., 2015). Oxygen concentrations recorded by the CTD instrument were calibrated based on dis-crete water samples using an automated Winkler titration procedure (Knap et al., 1996). Bottom water concentrations of ammonium (NH4

+), nitrite (NO2–) and nitrate (NO3

–) were measured colometrically on a SEAL QuAAtro segmented flow nutrient analyzer. Monitoring data at Lake Grevelingen (Wetstejn, 2011) shows that the water column is laterally homogenous over the Den Osse basin scale, which allows estimation of the bottom water parameters at S2 and S3 from corresponding depths in the measured CTD profiles at station S1.

4.2.2. Sediment geochemistry High-resolution depth profiles of O2 and sulfide (H2S) were measured in intact sediment cores to determine the oxygen penetration depth (OPD) and the sulfide appearance depth (SAD), using commercial micro-electrodes (Unisense A.S., Denmark) operated with a motorized micromanip-ulator (for details on the procedure see Malkin 2014). The oxygen penetration depth (OPD) is operationally defined as the depth below which [O2] < 1 µM, while the sulfide appearance depth (SAD) is operationally defined as the depth below which [H2S] > 1 µM (Seitaj et al., 2015).

Sediment cores were sectioned in increments of 0.5 cm from the sediment-water interface to 5 cm depth, and pore water was extracted by centrifugation, and analyzed following the proce-dure of (Sulu-Gambari et al., 2016). After filtration through 0.22 µm cellulose filters (Chromafil Xtra), pore water samples were analyzed for total free sulfide (∑H2S) via spectrophotometry (Cline, 1969; standard deviation ± 0.4 µM), whereas ammonium (NH4

+) was determined by a SEAL QuAAtro segmented flow analyzer (Aminot et al., 2009) after a 25 times dilution with a low nutrient seawater matrix solution (standard deviation ± 3.5%).

The free sulfide and ammonium concentration depth profiles of the sediment pore water were averaged to provide a 1 cm resolution down to 5 cm sediment depth to enable a direct comparison with results of the DNA/RNA and lipid analysis. The total organic carbon (TOC) content of the sediment was determined on sediment samples that were freeze-dried, ground to a fine powder and analyzed by an a Thermo Finnigan Delta plus isotope ratio monitoring mass spectrometer (irmMS) connected to a Flash 2000 elemental analyzer (Thermo Fisher Scientific, Milan). Before the analysis, samples were first acidified with 2N hydrogen chloride (HCl) to re-move the inorganic carbon (Nieuwenhuize et al., 1994). Concentrations of TOC are expressed as mass % of dry sediment.

4.2.3. DNA/RNA extraction DNA and RNA from sediments (previously centrifuged to remove excess of water thus values are given as grams of wet weight; S1, S2 and S3; 0–5 cm sediment depth; 1 cm resolution) were extracted by using the DNA and RNA PowerSoil® Total Isolation Kit, respectively (Mo Bio Laboratories, Inc., Carlsbad, CA). Nucleic acid concentrations were quantified spectrophoto-

Page 91: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

83

Material and Methods

metrically (Nanodrop, Thermo Scientific, Wilmington, DE) and checked by agarose gel electro-phoresis for integrity. Extracts were kept frozen at –80°C. The RNA extracts were treated with RNase-free DNase (DNA-free™, Ambion Inc., Austin, TX), and RNA quality and concentration were estimated by the Experion RNA StdSens Analysis Kit (Bio-Rad Laboratories, Hercules, CA). DNA contamination was checked by PCR using RNA as a template. Reverse transcription was performed as specified in Chapter 2.

4.2.4. PCR amplification and cloning Amplifications of the nirS gene of denitrifying bacteria (S1, S2, S3, March and August, 0–1 cm), and anammox bacteria 16S rRNA gene (S2, March, 1–2 cm), specific nirS genes of Scalindua sp. (S3, August, 0–1 cm) and the aprA gene (S2, August, 0–1 cm) were performed with the primer pairs specified in Table S1. The PCR reaction mixture consisted of (final concentration): Q-solu-tion (PCR additive, Qiagen, Valencia, CA) 1 ×; PCR buffer 1 ×; BSA (200 µg mL–1); dNTPs (20 µM); primers (0.2 pmol µL–1); MgCl2 (1.5 mM); 1.25 U Taq polymerase (Qiagen, Valencia, CA). PCR conditions for these amplifications were: 95°C, 5 min; 35 × [95°C, 1 min; Tm, 1 min; 72°C, 1 min]; final extension 72°C, 5 min. PCR products were gel purified (QIAquick gel purifi-cation kit, Qiagen, Valencia, CA) and cloned in the TOPO-TA cloning® kit (Life Technologies, Carlsbad, CA) and transformed in E. coli TOP10 cells following the manufacturer’s recommen-dations. In addition, in order to test the specificity of the quantitative PCR reaction we repeated the reactions of anammox bacteria 16S rRNA gene (Broc541F - Amx820R) with DNA extract of S2, March, 0–1 cm, nirS gene of heterotrophic denitrifying bacteria (nirS1F - nirS3R) with cDNA of S2, March, 0–1 cm, and aprA gene (Apr1F- Apr5R) with cDNA of S2, August, 0–1 cm, which were then treated to add 3’-A-overhangs and then cloned with the TOPO-TA cloning® kit as indicated above. Recombinant plasmid DNA was sequenced using the M13R primer by Macrogen Inc. (Amsterdam, The Netherlands).

4.2.5. Phylogenetic analysis Sequences were analyzed for the presence of chimeras using the Bellerophon tool at the Green-Genes website (http://greengenes.lbl.gov/). Sequences were aligned with MEGA6 software (Tamura, 2013) by using the alignment method ClustalW. The phylogenetic trees of the nirS and aprA genes were computed with the Neighbour-Joining method (Saitou and Nei, 1987) using the Poisson model with a bootstrap test of 1,000 replicates. The phylogenetic affiliation of the partial anammox bacteria 16S rRNA gene sequences was compared to release 123 of the SILVA NR SSU Ref database (http://www. arb-silva.de/; Quast et al., 2013) using the ARB software package (Ludwig et al., 2004). Sequences were added to the reference tree supplied by the SILVA database using the ARB Parsimony tool. Sequences were deposited in NCBI with the following accession numbers: KP886533–KP886678 for nirS gene sequences of denitrifiers, KP886679–KP8866700 for 16S rRNA gene sequences of anammox bacteria, KP886701–KP886721 for nirS gene sequences of anammox bacteria and KP886722–KP886804 for aprA gene sequences of SOB and SRB.

4.2.6. Quantitative PCR (qPCR) analysis qPCR analyses were performed on a Biorad CFX96TM Real-Time System/C1000 Thermal cycler equipped with the CFX ManagerTM software for sediment DNA/RNA extracts (S1, S2 and S3; 0–5 cm sediment depth; 1 cm resolution). Detailed information about the primers used in this

Page 92: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

84

study are summarized in Table S1. The abundance of denitrifying bacteria specific nirS gene was quantified using the primer set nirS1F/nirS3R as described by Braker et al. (1998). The abun-dance of anammox bacteria 16S rRNA gene was estimated using primers Brod541F/Amx820R as described by Li et al. (2010). Additionally, a fragment of the Scalindua sp. specific nirS gene, which codes for the cytochrome cd1-containing nitrite reductase, was quantified using the primer combination Scnir372F/Scnir845R as described by Lam et al. (2009). The abundance of SOB and SRB including sulfide dependent denitrifiers was estimated by targeting the dissimilatory adenosine-5’-phosphosulfate (APS) reductase (aprA gene) involved in the APS reduction of sul-fur-oxidizing bacteria and in the sulfite oxidation of sulfate-reducing bacteria by using the primer combination Apr-1-FW/Apr-5-RW as described by Meyer and Kuever (2007a) (see Table S1 for details). Gene abundances are expressed as copies g–1 sediment of wet weight.

All qPCR amplifications were performed in triplicate with standard curves ranging from 100 to 107 molecules per microliter. Standard curves and qPCR amplifications were performed as pre-viously described in Chapter 2. Coefficients of determination (R2) for standard curves ≥ 0.998 and qPCR efficiencies (E) ≥ 80% were accepted.

4.2.7. Anammox bacteria ladderane lipid analysis Intact polar lipids (IPLs) were extracted with the Bligh and Dyer extraction method (Bligh 1959; mod. by Pitcher et al., 2011) as described in detail by Bale et al. (2014). Intact ladderane phos-pholipids specific for anammox bacteria, the C20-[3]-monoether ladderane attached to a phos-phatidylcholine (PC) headgroup (PC-monoether ladderane) was analyzed by HPLC-MS/MS following Jaeschke et al. (2009) and quantified using an external standard consisting of isolated PC-monoether ladderane. Sediment samples between 0–5 cm depth (1 cm resolution) were an-alyzed and PC-monoether ladderane lipid concentrations were expressed per nanogram of dry weight sediment (ng g–1). In order to determine the fatty acid composition of the ladderane lip-ids, aliquots of the Bligh and Dyer extracts (BDE) obtained from the 0–1 and 4–5 cm sediment layers were saponified by reflux with aqueous KOH (in 96% MeOH) for 1 h. Fatty acids were obtained by acidifying the saponified samples to a pH of 3 with 1N HCl in MeOH and extracted using dichloromethane (DCM). The fatty acids were converted to their corresponding fatty acid methyl esters (FAMEs) by methylation with diazomethane (CH2N2) as described by Rush et al. (2012a). Polyunsaturated fatty acids (PUFAs) were removed by eluting the sample over a silver nitrate (AgNO3) (5 %) impregnated silica column with DCM and air-dried at room temperature. The fatty acid fractions were dissolved in acetone, filtered through a 0.45 μm polytetrafluoroeth-ylene (PTFE) filters (4 mm diameter), and analyzed by high performance liquid chromatography coupled to positive ion atmospheric pressure chemical ionization tandem mass spectrometry (HPLC/APCI-MS/MS) in selective reaction monitoring (SRM) mode following Hopmans et al. (2006) including the recent modifications described by Rush et al. (2012b). Ladderane lipids were quantified using external calibration curves of three standards of isolated methylated ladderane fatty acids (C20-[3]-ladderane fatty acid, and C20-[5]-ladderane fatty acid) (Hopmans et al., 2006; Rush et al., 2011).

4.3. Results Sediment samples were collected along a depth gradient in Den Osse basin (Lake Grevelingen) during two cruises March (before summer hypoxia) and August (during summer hypoxia) 2012. Sampling took place at three different stations: S1 was located in the deepest point of the basin

Page 93: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

85

Material and Methods/Results

at 34 m water depth, S2 at 23 m and S3 at 17 m.

4.3.1. Environmental conditionsThe temperature of the bottom water at station S1 (Fig. S1) showed a regular seasonal cycle with lowest values in late winter (1.5 °C in February) and highest values in late summer (16.9 °C in September). During the spring campaign (March 2012), the water column was only partially stratified and showed a limited surface-to-bottom temperature gradient. In contrast, during the summer campaign (August 2012), the water column was thermally stratified. The yearly pattern of the bottom water oxygenation at station S1 was inversely correlated the temperature, with greatest oxygenation levels in winter and fall, and lowest concentrations in summer (Fig. S1). In March 2012, the bottom water oxygenation was similar for all three stations, whereas in August 2012, the bottom water at S1 and S2 were anoxic (< 1 µM), while the bottom water at S3 still had 88 µM of O2 (36% air saturation). Bottom water ammonium (NH4

+) concentrations in station S1 ranged from 3 µM in March to 11.5 µM in August; nitrite (NO2

–) concentrations were relatively constant (0.7–1 µM) in March and August. Nitrate concentrations ranged from 28 µM in March to <2 µM in August in station S1 (Fig. S1) whereas in stations S2 and S3 values varied between 28 µM in March to ~10 µM in August.

The OPD in the sediment was seasonally variable and increased from S1 to S3, i.e. in S1 in March OPD was 1.5 mm and in August the sediment was completely anoxic. In S2, OPD was between 1.7–2.5 mm in March and in August ca. 0.5 mm (hypoxic) and in S3, the OPD was be-tween 1.5–2.2 mm in March and ca. 1.0 mm (hypoxic) in August. (Table 1). The sulfide appear-ance depth (SAD) varied between March and August in all stations. The SAD moved towards the sediment surface between March and August in all stations. In March, the SAD in stations S1 and S2 was at 18.4 and 21.3 mm respectively, whereas in S3, the SAD was detected at 41.8 mm sediment depth. However, in August, SAD in stations S1 and S2 was at 0.4 and 0.6 mm, respec-tively. In station S3, the sulfide was detected at 4.2 mm sediment depth (Table 1). At station S2, white mats of Beggiatoa sp.-like microorganisms covered the sediment surface in March. Small polychaetes were observed in the sediment at station S3 in March, suggesting some bioturbation. Sulfide (∑H2S) concentrations in the sediment pore water were low in March in all three stations, i.e. ranging from 0 to 0.007 mM, whereas in comparison, all three stations showed high sulfide concentrations in August, i.e. 0.15 to 1.6 mM (Table 1) (see Seitaj, 2015 for detailed geochemical profiles of station S1). Ammonium concentrations were low in March, i.e. ranging from 0.24 to 0.63 mM on average, in comparison with August when NH4

+ concentrations reached higher values, i.e. 0.7 to 1.2 mM on average (Table 1). The TOC content of the sediments varied slightly between stations and seasons, ranging between 1.8 and 4.4% (Table S2).

4.3.2. Diversity, abundance and potential activity of nirS-type denitrifiers In our study, we focused on heterotrophic and autotrophic bacteria that are able to perform the dissimilatory reduction of nitrite to nitric oxide. This reaction forms an intermediate step in the complete denitrification of nitrate to N2 and is catalyzed by the cytochrome cd1-contain-ing nitrite reductase encoded by the nirS gene (Braker et al., 2000). The diversity of nirS-type denitrifiers was evaluated for the surface sediment layer (0–1 cm) at the three stations in March and August by phylogenetic analysis targeting the nirS gene (Figs. 1 A, B). In general, the nirS sequences obtained (147 sequences in total) were closely related to nirS sequences of uncultured organisms found in coastal marine environments with a high input of organic matter such as

Page 94: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

86

Tabl

e 1:

Sed

imen

t por

ewat

er a

mm

onia

(NH

4+) a

nd su

lfide

(HS– )

conc

entra

tions

(1 c

m re

solu

tion)

, oxy

gen

pene

tratio

n de

pth

(OPD

) and

sul

fide

appe

ar-

ance

dep

th (S

AD

) det

erm

ined

by

mic

ro-s

enso

r pro

filin

g.

Stat

ion

Sedi

men

t dep

th

[cm

]H

S- [μM

]N

H4+

[μM

]O

PD [m

m]

SAD

[mm

]

Mar

chA

ugus

tM

arch

Aug

ust

Mar

chA

ugus

tM

arch

Aug

ust

1

0–1

081

027

965

6

1.5

018

.40.

4

1–2

015

0341

010

71

2–3

016

3963

613

22

3–4

019

6283

315

67

4–5

3320

6397

917

68

2

0–1

011

5716

555

0

1.7

2.5

21.3

0.6

1–2

080

245

583

6

2–3

010

0852

610

27

3–4

212

3859

211

38

4–5

2711

9067

112

28

3

0–1

021

173

537

1.5

2.2

41.8

4.2

1–2

017

715

469

4

2–3

014

623

673

6

3–4

010

933

374

9

4–5

093

408

746

* Dat

a ar

e av

erag

ed to

reac

h a

1 cm

reso

lutio

n; d

ata

prov

ided

by

Fatim

ah S

ulu-

Gam

bari

(201

6; u

npub

lishe

d da

ta),

**da

ta p

rovi

ded

by D

orin

a Se

itaj,

2015

; unp

ublis

hed

data

).

Page 95: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

87

Results

estuarine sediments and eutrophic bay sediments (Braker et al., 2000, Zhang et al., 2014). The phylogenetic analysis of protein sequences of the nirS gene revealed two distinct clusters (Fig. 1A; cluster 1 and 2), where most (ca. 95%) of the sequences clustered in cluster 1 (ca. 95%). Within cluster 1, approximately 90% of the nirS gene sequences were grouped into subcluster 1.1 and 10% into subcluster 1.2 (Figs. 1 A, B; Fig. S2). Within subcluster 1.1 sequences were affiliated to nirS sequences of members of α-, β- and γ-proteobacteria able to perform autotro-phic denitrification coupled to sulfide oxidation (Thiobacillus denitrificans) and heterotrophic de-nitrification (Azospirillum brasilense, Marinobacter hydrocarbonoclasticus, Kangiella aquamirina, Halomonas sp.). Sequences grouped in subcluster 1.2 were affiliated to the nirS gene sequences of divers α- proteobacteria able to perform autotrophic denitrification coupled to sulfide or iron oxida-tion (Paracoccus denitrificans, Sideroxidans lithotrophicus), as well as heterotrophic denitrification (e.g. Aromatoleum aromaticum, Azocarus toluclasticus, Acidovorax delafieldii). Sequences grouped into cluster 2 were affiliated to nirS gene sequences of uncultured bacteria detected in coastal marine and estuarine sediments (Braker et al., 2000; Zhang et al., 2014). In order to determine the specificity of the qPCR assay, sequences of nirS cDNA (complementary DNA of nirS mRNA) generated during the qPCR reaction were cloned and sequenced (28 sequences in total) and also added to the protein-coding nirS sequences phylogenetic tree showing that those sequences were grouped in the clusters described before (Figs. 1 A, B).

The abundance and distribution of nirS-type denitrifiers was estimated through quantification of the nirS gene copy number in the upper 5 cm (1 cm resolution) of the sediments at the three sampling sites in March and August (Fig. 2). The nirS gene abundance was relatively stable with depth with slightly higher values in station S1 (6.4 × 107 copies g–1 on average) compared to sta-tions S2 and S3 (5.7 × 107 and 5.2 × 107 copies g–1 on average, respectively). Overall, nirS gene abundance was slightly higher in March compared to August in S2 and S3, and this difference was especially evident in station S1.

To estimate the potential transcriptional activity of nirS-type denitrifiers, nirS transcripts (mRNA copy numbers) were quantified (data reported in Table S3) and the RNA:DNA ratio was calculated. In station S1, transcripts were only detectable in March in the upper 2 cm of the sediment, 5.6 × 103 copies g–1on average, whereas in August, nirS gene transcripts could only be detected within the 2–3 cm depth layer (102 copies g–1). In station S2, nirS gene transcripts were only detectable in March in the upper 3 cm and numbers varied between 1.9 × 102–1.5 × 103 copies g–1 with the highest value within the 1–2 cm zone. In station S3, the nirS gene transcripts could be detected in March in the upper cm of the sediment (1.5 × 102 copies g–1) and in August for 1–2 cm and for 3–4 cm sediment (1.1 × 102 and 1.8 × 103 copies g–1, respectively ) (Table S3). The ratio of nirS gene and transcript copies (RNA:DNA ratio) was ≤ 0.00013 in all stations in March and August.

4.3.3. Diversity, abundance and potential transcriptional activity of anammox bacteria Most of the anammox bacterial nirS gene sequences obtained in this study (20 sequences out of 21) were part of one cluster (cluster 1, Fig. 3) closely related to nirS sequences of “Candidatus Scalindua profunda” (Van de Vossenberg et al., 2013), and of uncultured bacteria obtained from continental margin sediment of the Arabian Sea (Sokoll et al., 2012) and surface sediments of the South China Sea (Li et al., 2013).

Page 96: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

88

M

ari

no

bacte

r h

yd

rocarb

on

ocla

sti

cu

s;

-Pro

teobact

erium

(Y

P005430950)

ɣ u

ncu

lture

d b

act

eriu

m c

lon

e;

est

ua

rin

e s

ed

ime

nt,

Ch

ina

(E

U2

35

96

7)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

) u

ncu

lture

d b

act

erium

clo

ne; e

stuarine s

edim

ent, C

hin

a (

EU

236093)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

1)

uncu

lture

d b

act

erium

; B

alti

c S

ea s

edim

ent (C

AJ8

7437)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235802)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (6

) u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (E

U235828)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

26)

Azo

sp

irillu

m b

rasilen

se;

-Pro

teobact

erium

(A

F361789)

α S

tati

on

1, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (2

1)

T

hio

bacillu

s d

en

itri

fican

s;

-Pro

teobact

erium

(W

P018076713)

β S

tati

on

2, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

9)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235749)

uncu

lture

d b

act

erium

; cl

one; est

uarine s

edim

ent, C

hin

a (

EU

235857)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

0)

Sta

tio

n 1

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (6

) u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235813)

Sta

tio

n 3

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (5

) u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235897)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

0)

Sta

tio

n 2

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (9

) S

tati

on

3, A

ug

ust;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

4)

uncu

lture

d b

act

erium

; B

alti

c S

ea s

edim

ent (C

AJ8

7433)

Sta

tio

n 3

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (2

6)

uncu

lture

d b

act

erium

; B

alti

c S

ea s

edim

ent (C

AJ8

7431)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

2)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

236072)

uncu

lture

d b

act

erium

; cy

anobact

erial a

ggre

gate

, B

alti

c S

ea (

CA

D29810)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

4)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (3

) S

tati

on

1, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (5

) S

tati

on

3, A

ug

ust;

0-1

cm

Lake G

revelin

gen

sed

imen

t (7

) u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235993)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

ts; cD

NA

(20)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235797)

Sta

tio

n 3

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

1)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235880)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235841)

Kan

gie

lla a

qu

imari

na;

-Pro

teobact

erium

(W

P018624170)

ɣ u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235809)

Sta

tio

n 1

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (5

) u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235873)

Sta

tio

n 3

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

8)

uncu

lture

d b

act

erium

; co

ast

al m

arine s

edim

ent, W

A (

AJ2

48421)

Sta

tio

n 3

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (3

0)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235919)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

4)

un

cultu

red

ba

cte

riu

m c

lon

e;

est

ua

rin

e s

ed

ime

nt,

Ch

ina

(E

U2

35

78

3)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235874)

90 5

6

93

865

6 59

6264

63 0.1

48

se

qu

en

ce

s 0

-1 c

m L

ak

e G

rev

elin

ge

n s

ed

ime

nt:

8 x

Sta

tion

1 M

arc

h 3

x S

tatio

n 1

Au

gu

st 4

x S

tatio

n 2

Ma

rch

7

x S

tatio

n 2

Au

gu

st +

10 x

Sta

tion 2

, M

arc

h, cD

NA

6 x

Sta

tion

3 M

arc

h 1

0 x

Sta

tion

3 A

ug

ust

4

x S

tatio

n 1

Marc

h 5

x S

tatio

n 1

August

19 x

Sta

tion 2

Marc

h 7

x S

tatio

n 2

August +

1

1 x

Sta

tion 2

, 1

6 x

Sta

tion 3

Marc

h 1

5 x

Sta

tion 3

August M

arc

h, cD

NA

77

se

qu

en

ce

s 0

-1 c

m

La

ke

Gre

ve

lin

ge

n

se

d.

C

aen

isp

irillu

m s

alin

aru

m;

-Pro

teobact

erium

(W

P00954211

8)

α H

alo

monas

sp. H

GD

K1 (

AC

V88081)

Kilo

nie

lla lam

inari

ae;

-Pro

teobact

erium

(W

P020593499)

α T

hala

sso

sp

ira x

iam

en

en

sis

; -P

rote

obact

erium

(W

P007089813)

α M

ag

neto

sp

irillu

m m

ag

neti

cu

m;

-Pro

teobact

erium

(W

P011

386514)

α M

agneto

spirill

um

sp. S

O-1

(W

P008620648)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235964)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235827)

B

rad

yrh

izo

biu

m o

lig

otr

op

hic

um

;-P

rote

obact

erium

(W

P015669866)

α u

ncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235956)

Para

co

ccu

s d

en

itri

fican

s;

-Pro

teobact

erium

(W

P011

748767)

α P

ara

cocc

us

denitr

ifica

ns

(AA

B17878)

Para

co

ccu

s p

an

totr

op

hu

s;

-Pro

teobact

erium

(W

P024843009)

α S

tati

on

3, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

2)

No

vis

pir

illu

m ite

rso

nii;

-Pro

teobact

erium

(W

P019646524)

α R

oseo

bacte

r lito

ralis;

-Pro

teobact

erium

(W

P013963004)

α R

ose

obact

er

litora

lis (

AE

I95094)

Th

ala

sso

bacte

r are

nae

; -P

rote

obact

erium

(W

P0211

00911

Rhodobact

er

sp. S

w2 (

WP

008027070)

uncu

lture

d b

act

erium

; m

arine s

edim

ent, W

A (

AJ2

48438)

Sta

tio

n 3

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

6)

R

ueg

eri

a p

om

ero

yi;

-Pro

teobact

erium

(W

P011

241999)

α R

uegeria p

om

ero

yi (

AA

V97354)

L

ab

ren

zia a

gg

reg

ata

; -P

rote

obact

erium

(W

P006938174)

α S

tati

on

1, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

8)

Sta

tio

n 1

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

3)

Aci

dovo

rax

sp. M

R-S

7 (

WP

020228065)

Cupriavi

dus

sp. B

IS7 (

WP

01944911

0)

S

idero

xyd

an

s lit

ho

tro

ph

icu

s;

-Pro

teobact

erium

(W

P013029265)

β

Ac

ido

vo

rax

de

lafi

eld

ii;

-Pro

teo

ba

cte

riu

m (

WP

00

57

93

43

6)

β A

rom

ato

leu

m a

rom

ati

cu

m;

-Pro

teobact

erium

(W

P011

236329)

β A

zoarc

us t

olu

cla

sti

cu

s;

-Pro

teobact

erium

(WP

018988937)

β P

seu

do

mo

nas s

tutz

eri

; -P

rote

obact

erium

(W

P003279945)

ɣ T

hau

era

terp

en

ica;

-Pro

teobact

erium

(W

P021248854)

β u

ncu

lture

d b

act

erium

; m

arine s

edim

ent, W

A (

AJ2

48420)

Sta

tio

n 2

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

6)

Sta

tio

n 3

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (2

9)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

28)

uncu

lture

d b

act

erium

clo

ne; est

uarine s

edim

ent, C

hin

a (

EU

235968)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (2

7)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

27)

Sta

tio

n 1

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (2

7)

Sta

tio

n 3

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t (3

0)

Sta

tio

n 2

, M

arc

h;

0-1

cm

Lake G

revelin

gen

sed

imen

t; c

DN

A (

28)

Sta

tio

n 1

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

3)

Sta

tio

n 1

, Au

gu

st;

0-1

cm

Lake G

revelin

gen

sed

imen

t (1

2)

63

99

98

10

0

10

0

10

0

10

0

10

0

94

10

0

86

10

0

71

85

73 6

5

76

77

75

64

0.1

cluster 1

su

bclu

ste

r 1 (d

eta

ils s

ee

pa

rt B

)

subcluster 1

subcluster 2

cluster 2

12

se

qu

en

ce

s 0

-1 c

m L

ak

e G

rev

eli

ng

en

se

dim

en

t: 5

x S

tatio

n 1

Ma

rch

, 2

x S

tatio

n 2

Ma

rch

+ 2

x (

cDN

A),

2

x S

tatio

n 3

Ma

rch

, 1

x S

tatio

n 3

Au

gu

st

A)

B)

Figu

re 1:

Phy

loge

netic

tree

of

part

ial n

irS g

ene

sequ

ence

s of

deni

trify

ing

bact

eria

retri

eved

in th

is st

udy

(147

DN

A se

quen

ces r

ecov

ered

from

stat

ions

S1,

S2

and

S3

betw

een

0–1

cm se

dim

ent d

epth

, in

Mar

ch a

nd A

ugus

t by

ampl

ifica

tion

(PC

R) a

nd c

loni

ng a

nd 2

8 cD

NA

sequ

ence

s obt

aine

d fr

om st

atio

n S2

be

twee

n 0–

1 cm

sed

imen

t dep

th in

Mar

ch re

cove

red

by a

mpl

ifica

tion

(qPC

R) a

nd c

loni

ng) a

nd c

lose

st re

lativ

es (b

old:

our

seq

uenc

es a

nd c

lose

st k

now

n re

lativ

es);

the

scal

e ba

r ind

icat

es 1

0 %

sequ

ence

div

erge

nce.

Page 97: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

89

Results

The diversity of the 16S rRNA gene sequences of anammox bacteria (Fig. S3) obtained from surface sediments (0–1 cm) revealed that most of the sequences (21 sequences derived from PCR plus cloning, and 23 gene sequences obtained from the qPCR assay and further cloning to check the specificity of the qPCR assay) were closely related to “Candidatus Scalindua” bro-dae and “Candidatus Scalindua marina” obtained from surface sediments of the Gullmar Fjord (Brandsma et al., 2011), and to sequences detected in OMZ waters of the Arabian Sea, Peru and Namibia (Woebken et al., 2008) and Peruvian coastal margin (Henn, unpublished). Two 16S rRNA sequences were more distantly related to the other sequences, and were closely related to the 16S rRNA sequence of “Candidatus Scalindua wagneri” obtained from a bioreactor (Woeb-ken et al., 2008) (Fig. S3).

We also quantified the C20-[3]-ladderane monoether-PC (for 0–5 cm sediment depth) and ladderane core lipids (for 0–1 and 4–5 cm sediment depth) (Table S4 and S5) as markers for the presence of anammox bacteria. Abundance of PC-monoether ladderane ranged between 0.9–20.4 ng g–1 with highest values in station S1 in March (between 3.4–20.4 and 2.4–7 ng g–1, respectively). PC-monoether ladderane values were relatively constant with depth in August in all stations (between 0.9–7 ng g–1; Table S4). On the other hand, PC-monoether ladderane abun-dance was always higher in March in the upper 2 cm of the sediment and decreased four-fold between 2 and 5 cm in all stations (on average from 10.7 to 2.5 ng g–1). The summed concen-tration of the ladderane fatty acids was on average lowest in station S1 (13 ng g–1), with slightly

Figure 2: Depth profiles of anammox bacteria 16S rRNA gene abundance [copies g–1] (circle); anammox bacteria nirS gene abundance [copies g–1] (triangle), denitrifying bacteria nirS gene abundance [copies g–1] (square); SOB and SRB aprA gene abundance [copies g–1] (diamond); A) Station 1; B) Station 2; C) Station 3; black symbol: March; white symbol: August.

Page 98: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

90

uncultured bacterium, Peruvian OMZ seawater (ACR07885)uncultured Scalindua sp., Black Sea waterCandidatus (AFO85870)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (15)

uncultured bacterium, Peruvian OMZ seawater (ACR07884)Kuenenia stuttgartiensisCandidatus (CT573071)

Anaerolinea thermophila (WP013561145)uncultured bacterium, Peruvian OMZ seawater (ACR07901)

uncultured Scalindua sp., Pacific marine water column, USACandidatus (AGM49043)2 sequences subseafloor sediment, South China Sea (ACV84553; ACV84614)2 sequences water column (AFO85813; AFO85819)Black Sea OMZ

uncultured bacterium, Peruvian OMZ seawater (ACR07879)uncultured bacterium, Peruvian OMZ seawater (ACR07891)

uncultured bacterium,Peruvian OMZ seawater(ACR07882)uncultured bacterium, Peruvian OMZ seawater (ACR07881)uncultured bacterium, Peruvian OMZ seawater (ACR07877)

uncultured bacterium, Peruvian OMZ seawater (ACR07883)uncultured bacterium, Peruvian OMZ seawater (ACR07890)

uncultured bacterium, Peruvian OMZ seawater (ACR07889)Uncultured bacterium, Peruvian OMZ seawater (ACR07875)

uncultured Scalindua sp.Candidatus Black Sea water (AFO85831)uncultured Scalindua sp.Candidatus Black Sea water (AFO85849)Candidatus Scalindua (scal02098)profundauncultured bacterium, sediment, South China Sea (ACV84622)uncultured Scalindua sp.Candidatus Black Sea water (AFO85782)uncultured bacterium, surface sediment, South China Sea (AEJ85669)uncultured bacterium, sediment, Arabian Sea (AFZ84060)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (20)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (5)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (7)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (4)Station 3, August; Lake Grevelingen 0-1 cm sediment depth (18)

15 x Station 3, August; Lake Grevelingen 0-1 cm sediment depth

5473

50

65

58

0.25

clu

ste

r 1

Figure 3: Phylogenetic tree of partial nirS gene sequences of anammox bacteria retrieved in this study and closest relatives (S3; 0–1 cm Lake Grevelingen sediment; bold: sequences retrieved in March and se-quences of known relatives); the scale bar indicates 25% sequence divergence.

higher average values in S2 (27 ng g–1), and the highest average values in S3 (41 ng g–1) (Table S5). The summed ladderane fatty acid concentrations were comparable to the abundance of anammox bacteria determined by the 16S rRNA gene quantification in the first centimeter of the sediment obtained in different stations and seasons, whereas the PC-monoether ladderane revealed a different trend compared to the 16S rRNA gene abundance, i.e., was more abundant in March compared to August (Table S4).

The abundance of anammox bacteria was determined by the quantification of the 16S rRNA gene copy number of anammox bacteria as well as of the nirS gene copy number of members of the genus “Candidatus Scalindua” (Figure 2) (Strous et al., 2006; Li et al., 2011). Copy numbers of the anammox bacteria 16S rRNA gene ranged between 9.5 × 105 and 8.1 × 107 copies g–1 with highest values in S3 (between 2 × 106 and 8.1 × 107 copies g–1). The abundance of the ana-mmox bacterial 16S rRNA gene was slightly higher in August compared to March in all stations. The “Candidatus Scalindua” nirS gene abundance followed the same depth trend but values were 2 – 3 orders of magnitude lower (between 1.1 × 104 and 9.9 × 104 copies g–1) compared to the anammox bacteria 16S rRNA gene copy numbers.

The anammox bacterial 16S rRNA transcript, used as a proxy for the potential transcriptional activity, varied between 1.3 × 103 – 5.1 × 106 copies g–1 of sediment (Table S3). The anammox 16S rRNA transcript copy number varied slightly and reached highest values in S2 and S3 in August between 2–4 cm sediment depth (2.7 × 106 copies g–1 on average). The ratio between 16S rRNA gene and transcript (RNA:DNA ratio) was ≤ 0.32 in all stations in March and August.

Page 99: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

91

Results

Anammox bacteria nirS transcript copies were below detection level of the qPCR assay.

4.3.4. Diversity, abundance and potential transcriptional activity of bacteria involved in sulfur cycling Here, we focused on bacteria involved in sulfur cycling, performing dissimilatory sulfur oxida-tion or sulfate reduction. The dissimilatory APS reductase encoded by the aprA gene is operating in SRB, converting APS to adenosine monophosphate (AMP) and sulfite (SO3

2–), as well as in SOB, operating in reverse direction, oxidizing sulfite to APS (Meyer and Kuever, 2007b). The diversity of the microorganisms harboring the aprA gene was evaluated in the first centimeter of the sediment core at S2 in August as the observation of Beggiatoa mats on top indicated the occurrence of sulfide oxidation at this station. The protein sequences (86 sequences PCR + cloning) coded by the aprA gene grouped in two clusters (Fig. 4; cluster I (66%) and II (34%)). Within cluster I, sequences clustered in three distinct subclusters. Sequences of the aprA gene included in subcluster 1.1 (36%) were affiliated to heterotrophic SRB of the δ-proteobacteria class (e.g. Desulfosarcina sp., Desulfofaba gelida, Desulfobulbus propionicus) and to aprA gene sequences of uncultured bacteria found in environments such as in Black Sea sediments and associated with benthic organisms (Blazejak and Schippers, 2011; Ruehland et al., 2008). In subcluster 1.2 (21%), sequences were closely affiliated to β- and γ-proteobacteria involved in autotrophic sul-fur-dependent denitrification (Thiobacillus denitrificans and Sulfuricella denitrificans) or in phototro-phic sulfur oxidation (Lamprocystis purpurea) and to the aprA gene sequence of an uncultured bacterium associated with the sea urchin Asterechinus elegans (Quast et al., 2009). Sequences clus-tering in subcluster 1.3 (9%) were affiliated with aprA gene sequences of bacteria of the phylum Firmicutes (i.e. Desulfotomaculum sp.) known to perform heterotrophic sulfate reduction, and to sequences of uncultured bacteria obtained in various sediments such as hydrothermal seep sedi-ments, Peru margin sediments and salt lake sediments (Blazejak and Schippers, 2011; Meyer and Kuever, 2007c; Kleindienst et al., 2012). Cluster II contained sequences (34%) closely related to obligately chemolithoautotrophic members of the β-proteobacteria class (Thiobacillus thioparus) able to perform nitrate reduction to nitrite with thiocyanate, and to aprA gene sequences of uncultured bacteria retrieved in salt lake sediments or associated with benthic organisms (Becker et al., 2009). In addition, to assess the specificity of the aprA gene qPCR assay, aprA gene tran-scripts (cDNA of aprA mRNA) generated during the qPCR assay (21 sequences in total) were cloned, sequenced and added to the aprA phylogenetic tree, where they were classified in the clusters previously described (Fig. 4).

The abundance and the potential activity of SOB including sulfide-dependent denitrifiers and SRB were determined by the quantification of the aprA gene and its gene transcripts (Fig. 2, Table S3). The copy number of the aprA gene was relatively constant with varying sediment depth and season. Gene copy numbers of the aprA gene were of the same order of magnitude as observed for the denitrifiers nirS (on average 3.1 × 108 copies g–1) and reached highest val-ues in S2 in March and August (on average 3.3 × 108 copies g–1). At station S1, the aprA gene transcript (Table S3) was detectable at 2–3 cm sediment depth in March (5.4 × 102 copies g–1 on average), whereas in August the nirS transcript was detectable at 0–1 cm (3.8 × 103 copies g–1) and 2–4 cm (1.8 × 104 copies g–1 on average) sediment depth. In S2 and S3, aprA gene transcripts were detectable at 1–4 cm sediment depth in March and August (on average 4.7 × 103 copies g–1 in March and 4.4 × 103 copies g–1 in August). The ratio between aprA gene and transcript (RNA:DNA ratio) was ≤ 0.000088 in all stations in March and August.

Page 100: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

92

23 sequences Station 2, August; 0-1 cm Lake Grevelingen sediment (including 3 x cDNA) closely related to sequences of uncultured bacteria and δ-Proteobacteria Desulfatibacillum alkenivorans, Desulfofaba gelida and Desulfobacter curvatus

41 sequences Station 2, August; 0-1 cm Lake Grevelingen sediment (including 12 x cDNA) closely related to sequences of uncultured bacteria and the ɣ-Proteobacterium Thiobacillus thioparus (WP018507768)

Station 2, August; 0-1 cm Lake Grevelingen sediment (45) Olavius algarvensis endosymbiont; marine sediment (CAJ81243)

Station 2, August; 0-1 cm Lake Grevelingen sediment (44) Station 2, August; 0-1 cm Lake Grevelingen sediment (79)

Station 2, August; 0-1 cm Lake Grevelingen sediment (57) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (18)

uncultured bacterium; mud vulcano (AFV48087) Station 2, August; 0-1 cm Lake Grevelingen sediment (56) Station 2, August; 0-1 cm Lake Grevelingen sediment (63) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (9) Station 2, August; 0-1 cm Lake Grevelingen sediment (39) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (19)

Station 2, August; 0-1 cm Lake Grevelingen sediment (59) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (1)

Station 2, August; 0-1 cm Lake Grevelingen sediment (65) uncultured δ-Proteobacterium; surface of crab (ABY89711) Desulfobulbus propionicus; δ-Proteobacterium (YP004194237) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (23)

uncultured Desulfofustis sp.; δ-Proteobacterium (WP023406400)uncultured bacterium; marine sponge (AFK76428)

Station 2, August; 0-1 cm Lake Grevelingen sediment (1) Station 2, August; 0-1 cm Lake Grevelingen sediment (33)

Station 2, August; 0-1 cm Lake Grevelingen sediment (74) uncultured bacterium; marine sediment Gulf of Mexico (CCB84484)

Station 2, August; 0-1 cm Lake Grevelingen sediment (37) Station 2, August; 0-1 cm Lake Grevelingen sediment (67)

uncultured bacterium; Black Sea sediment (CCC58324) Desulfotomaculum alcoholivorax; Firmicutes (WP027366015)

Station 2, August; 0-1 cm Lake Grevelingen sediment (13) Station 2, August; 0-1 cm Lake Grevelingen sediment (52)

Station 2, August; 0-1 cm Lake Grevelingen sediment (50) Station 2, August; 0-1 cm Lake Grevelingen sediment (69)

uncultured bacterium; Peru margin sediment (CCC58328) Desulfotomaculum thermocisternum; Firmicutes (WP027356136)

Station 2, August; 0-1 cm Lake Grevelingen sediment (14) Station 2, August; 0-1 cm Lake Grevelingen sediment (64)

uncultured bacterium; salt lake sediment (ACM47772) δ-Proteobacterium NaphS2 (WP006420559)

Archaeoglobus sulfaticallidus; Archaeoglobi (YP007908062) Station 2, August; 0-1 cm Lake Grevelingen sediment; cDNA (13)

5799

52

99

82

9799

99

98

91

83

5190

8453

69

98 95

99

63

51

93

90

83

54

89

0.05

23 sequences Station 2, August; 0-1 cm Lake Grevelingen sediment (including 5 x cDNA) closely related to sequences of uncultured bacteria and β-ProteobacteriaThiobacillus denitrificans and Sulfuricella denitrificans and ɣ-Proteobacterium Lamprocystis purpurea

clu

ste

r 1

su

bclu

ste

r 1

su

bclu

ste

r 2

su

bclu

ste

r 3

clu

ste

r 2

Figure 4: Phylogenetic tree of partial aprA gene sequences of sulfur-oxidizing and sulfate-reducing bac-teria (SOB and SRB) retrieved in this study (86 DNA sequences recovered by amplification (PCR) and 21 cDNA sequences recovered by amplification (qPCR) and cloning from sediments (0–1 cm) of station S2 in August) and closest relatives (bold: our sequences and closest known relatives); the scale bar indicates 5% sequence divergence.

Page 101: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

93

Discussion

4.4. DiscussionThe phylogenetic analysis of the nirS gene sequences amplified from Lake Grevelingen sed-iments revealed a substantial diversity of autotrophic and heterotrophic denitrifiers (Fig. 1). Most of the sequences obtained in this study were affiliated to α-, β- and γ-proteobacteria able to perform denitrification coupled to the oxidation of reduced sulfur compounds. Sequences of the nirS gene (~30%) and the aprA gene (21%) (Fig. 4) were closely related to sequences of Thiobacillus denitrificans, able to couple the oxidation of inorganic sulfur species such as sulfide with denitrification (Shao et al., 2010). Sulfide-dependent denitrifiers couple nitrogen and sulfur cycling and their presence point to an important denitrification potential in these sediments. In order to determine differences in abundance, and therefore of denitrification potential, be-tween sediments and with depth, in our study we quantified the abundance of the nirS gene. The nirS gene abundance was relatively constant with increasing sediment depth, suggesting a stable community of heterotrophic and autotrophic denitrifying bacteria in Lake Grevelingen sediments. NirS gene abundance (in the order of 107 copies g–1) was slightly higher in March in all three stations, but the difference was especially notable in station S1, most likely explained by the lower free sulfide concentrations in March compared to August (Table 1). Lower sulfide concentrations in March could favor the proliferation of heterotrophic denitrifiers not coupled with sulfide oxidation as sulfide has been seen to impair denitrification by inhibition of NO and N2O reductases (Sørensen et al., 1980). In addition, the nirS gene abundance (Fig. 2) determined in Lake Grevelingen sediments (between 107–108 copies g–1) was one to three orders of mag-nitude higher compared to the values previously detected in other marine sediments such as in sediments of open estuaries with permanently oxygenated bottom water (104–107 copies g–1 on average) (Smith et al., 2007; Zhang et al., 2013). The nirS gene abundance has previously been shown to correlate positively with potential denitrification rates (Mosier and Francis, 2010), and therefore, we can conclude that the sediments of the marine Lake Grevelingen harbor a stable community of heterotrophic and autotrophic denitrifiers, which can be actively involved in the denitrification process independent of season and sediment depth.

Besides quantifying the nirS gene as an indicator of the denitrification potential of these sedi-ments, we also estimated the presence of nirS gene transcripts as a measure of potential activity (Table S3). Unfortunately, there is a lack of in situ studies correlating nirS gene expression with denitrification rates, and some studies have even shown a lack of direct relationship between nirS gene expression and modeled rates of denitrification in marine surface sediments (Bowen et al., 2014). On the other hand, other studies have observed a correlation between a decrease of nirS gene transcripts and reduction of denitrification rates together with declining concentrations of nitrate in estuarine sediment (Dong et al., 2000, 2002; Smith et al., 2007), which would justify the use of nirS gene transcripts as a proxy for potential denitrification performed by nirS-type denitrifiers. In this study, the copy number of denitrifier nirS transcripts was two to three orders of magnitude lower compared to values previously reported for estuarine sediments (Smith et al., 2007), suggesting a low denitrification rate in surface sediments of all stations in March and in deeper sediment layers in August. Gene expression of the nirS gene was detected in surface sediments in March (all stations) and in deeper sediment layers in August (stations S1 and S3), which is most likely explained by the availability of nitrate/nitrite in the water-sediment interface (bottom water).

Page 102: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

94

Apart from nirS-containing denitrifiers, also anammox bacteria were detected in Lake Grev-elingen sediments by applying both DNA and lipid-based biomarkers. The phylogenetic analysis of the anammox 16S rRNA gene amplified from the sediments pointed to an anammox com-munity dominated by “Candidatus Scalindua” (Fig. S3). In order to estimate the abundance of anammox bacteria in the sediments and their potential role in the nitrogen cycle in this system, we estimated the abundance of the “Candidatus Scalindua” nirS gene. In contrast to nirS-type denitrifiers, anammox bacteria showed a clear seasonal contrast in their nirS gene abundance with higher copy numbers in August (summer hypoxia) compared to March (Fig. 2). However, this seasonal trend was not reflected in the abundance of the anammox bacteria lipid biomarker PC-monoether ladderane lipid (Table S4). Previous studies have suggested that this biomarker lipid could be partly preserved in the sediment due to a relatively low turnover rate (Brandsma et al., 2011; Bale et al., 2014; Chapter 2). On the other hand, the concentration of ladderane lipid fatty acids (Table S5) correlated with the anammox bacteria abundance given by 16S rRNA gene quantification, i.e. with the lowest values in station S1 and highest values in station S3. This suggests that the abundance of ladderane lipid fatty acids could be interpreted as a proxy for anammox bacteria abundance together with specific gene quantification. It is also worth notic-ing that a difference of three to four orders of magnitude was detected between abundances of anammox 16S rRNA gene and the “Candidatus Scalindua” nirS gene (Fig. 2). This discordance between anammox 16S rRNA genes and functional genes has been previously observed and is possibly linked to primer biases attributed to the anammox bacteria 16S rRNA gene primers that would amplify 16S rRNA gene fragments of bacteria other than anammox bacteria (Li et al., 2010; Harhangi et al., 2012; Bale et al., 2014; Chapter 2). However, in our study we have ruled out the possibility of the anammox bacteria 16S rRNA gene primers amplifying bacteria other than anammox by cloning and sequencing of the product generated during the qPCR reaction as shown in Fig. S3. Therefore, further studies should address the causes of this discordance.

Factors other than hypoxia and sulfide concentration could have also contributed to the dif-ferences in abundance and activity of anammox bacteria. For example, both the anammox bac-teria 16S rRNA and nirS gene abundances were higher in August in comparison to March (Fig. 2). This may indicate that anammox bacteria are more abundant at higher temperatures (15°C), which is the anammox bacteria temperature optimum in temperate shelf sediments (Dalsgaard et al., 2005). This seasonality of anammox bacteria abundance has been reported before in sandy and as well in muddy, organic rich sediments of the North Sea (Bale et al., 2014; Chapter 2). Ad-ditionally, oxygen penetration depth in the sediments decreased and the sediment even became entirely anoxic during summer stratification (Table 1), which would also favor the anaerobic metabolism of anammox bacteria (Dalsgaard and Thamdrup, 2002). Anammox bacterial abun-dance was highest in station S3, which could also be related to elevated bioturbation activity that was observed in this station, as bioturbation and mixing extends the area of nitrate reduction, which might fuel the anammox process (Meyer et al., 2005; Laverock et al., 2013). Another factor determining anammox and denitrification pathways in sediments is the organic carbon content. Anammox and denitrification have been suggested to be more important nitrogen removal path-ways in sediments of low carbon input compared to sulfide-dependent denitrification which seem to be more important in sulfidic sediments with high carbon input (Burgin and Hamilton, 2007). The TOC content in Lake Grevelingen sediments is high, i.e. in the order of 2.5–4.5% (Malkin et al., 2014), which is hence consistent with the low activity of anammox bacteria. More-

Page 103: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

95

Discussion

over in station S1, fresh organic-rich sediment rapidly accumulates (> 2 cm yr–1; Malkin et al., 2014), which might inhibit the activity of anammox bacteria and denitrifying bacteria.

Overall, the nirS gene abundance of denitrifiers was three to four orders of magnitude higher compared to the anammox bacteria ‘Scalindua’ nirS gene values, suggesting that nirS-type denitri-fiers have a more prominent role in the overall denitrification activity in Lake Grevelingen sed-iments in comparison with anammox bacteria (Fig. 2). This is also supported by the anammox bacteria nirS gene transcript abundance, which remained below detection limit. On the other hand, anammox bacteria 16S rRNA gene transcripts were detected throughout the sediment (Table S3). A recent study by Bale et al. (2014) observed a good correlation between the 16S rRNA gene abundance, 16S rRNA gene transcript abundance and anammox rates in North Sea sediments, which suggests that the transcriptional activity of anammox bacteria 16S rRNA gene is a suitable proxy of anammox bacteria activity. However, the anammox bacteria 16S rRNA gene RNA:DNA ratio was low (0.01–1) compared to the values previously reported in sediments of the southern North Sea (1.6–34.6) (Chapter 2), further supporting a low anammox activity in the Lake Grevelingen sediments.

Although denitrification was more relevant than anammox process as suggested by the gene abundance and potential activity (Fig. 2, Table S3), the measured denitrification rates (between 36–96 µM m–2 d–1; D. Seitaj, personal communication) in Lake Grevelingen are low in compar-ison with other marine sediments. For example, denitrification rates were reported to vary be-tween 30–270 µM m–2 d–1 in marine Arctic sediments (Rysgaard et al., 2004), while denitrification rates of 270±30 µmol N2 m

−2 d−1 were measured in sediments of the Lower St. Lawrence Estu-ary (Crowe et al., 2012). The lower denitrification and anammox potential in Lake Grevelingen sediments could be explained by the effect of high concentrations of sulfide (potentially inhib-iting both heterotrophic denitrification and anammox). In our study, the abundance of nirS-type denitrifiers did not vary substantially for the different stations however the highest anammox bacterial abundance was observed in station S3 where lowest sulfide concentration was reported (Fig. 2, Table 1). Besides, these sediments harbor an important population of sulfide-dependent denitrifiers (as indicated by aprA gene values comparable to those found in sediments of the Black Sea; Blazejak and Schippers, 2011), which could alleviate the inhibitory effect of sulfide on other microbial groups such as anammox bacteria. For example, a study by Russ et al. (2014) observed the coexistence and interaction of sulfide-dependent denitrifying and anammox bacte-ria in a co-culture in which anammox bacteria remained active. Also a study by Wenk et al (2013) provided evidence for the coexistence of anammox bacteria and sulfide-dependent denitrifiers in the stratified water column of Lake Lugano, and reported that the addition of sulfide in incu-bation studies enhanced both processes. They speculated that anammox bacteria in this system would rely on nitrite released as intermediate during sulfide-dependent denitrification and that they would overcome inhibiting or toxic effects of sulfide by creating a sulfide-free microenvi-ronments in aggregates as previously proposed by Wenk et al (2013). However, in the case of Lake Grevelingen sediments the potential detoxification of sulfide and a source of nitrite by sulfide-dependent denitrifiers did not translate in significant anammox potential. Another expla-nation for the low denitrification potential of these sediments can be found in the interactions of anammox and heterotrophic denitrifiers with sulfur oxidizers also involved in the nitrogen

Page 104: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

96

cycle. For example in Lake Grevelingen, cable bacteria have been detected in stations S1 and S3 in March, whereas station S2 was dominated by Beggiatoaceae down to ~3 cm, but both groups were hardly detectable in subsurface sediments (0.5 cm downwards) during summer hypoxia (Seitaj et al., 2015; Seitaj, unpublished data). In fact, the analysis of aprA gene of SOB/SRB in our study showed that up to 36% of the aprA gene sequences were closely related to the aprA se-quences of Desulfobulbus propionicus (Fig. 4), which has been identified as closest known relative of cable bacteria (Pfeffer et al., 2012). Cable bacteria perform a novel ‘electrogenic’ form of sulfur oxidation, in which the oxidation of the electron donor and the reduction of the electron accep-tor are separated over centimeter-scale distances, and the necessary redox coupling is ensured by long-distance electron transport (Nielsen et al., 2010; Pfeffer et al., 2012). Cable bacteria use ox-ygen as terminal electron acceptor (Nielsen et al., 2010; Meysman et al., 2015) and recent studies also indicate that nitrate (Marzocchi et al., 2014) and nitrite (Risgaard-Petersen et al., 2014) can be utilized, suggesting that cable bacteria can also play a role in bioavailable nitrogen removal. In addition, marine Beggiatoa spp. couple the oxidation of sulfide to nitrate reduction resulting in N2 and/or NH4

+ (Mußmann et al., 2003). Previous studies have observed that both denitrification and anammox in anoxic sediments can be supported by intracellular nitrate transport performed by sulfide-oxidizing bacteria like Thioploca and Beggiatoa (Prokopenko et al., 2013; Muβmann et al., 2007; Jørgensen et al., 2010) down to deeper sediment layers and possibly supplying anammox bacteria with nitrite and/or ammonia produced by DNRA (Teske and Nelson, 2006, Prokopen-ko et al., 2011).The nrfA gene, encoding a nitrite reductase catalyzing the conversion of nitrite to ammonia (Smith et al., 2007), could not be detected in our sediment samples (data not shown) by using general nrfA primers (Mohan et al., 2004), however we cannot rule out completely the presence of microorganisms performing DNRA as the primers used could be not the most ap-propriate ones for this system. Therefore, the presence of sulfide-dependent denitrifiers (such as Thiobacillus denitrificans), cable bacteria and Beggiatoa-like bacteria present in Lake Grevelingen sediments could potentially support denitrification and anammox processes in March. However, the nitrate and nitrite concentrations in the bottom water in March were reported to be relatively low (28 and 0.7 μM on average, respectively, Fig. S1), which is expected to induce a strong com-petition for nitrate, nitrite and sulfide with Beggiatoaceae, cable bacteria and other nitrate-reduc-ing bacteria (e.g. Thiobacillus thioparus), explaining the limited denitrification potential observed in the Lake Grevelingen sediments.

4.5. Conclusion Our study has unraveled the coexistence and potential activity of heterotrophic and autotrophic denitrifiers, anammox bacteria as well as SOB/SRB in seasonally hypoxic and sulfidic sediments of Lake Grevelingen. Our starting hypothesis was that the abundance and activity of denitrifiers and anammox bacteria would decrease upon increase of the sulfide concentration found in the sediments during the summer hypoxia. However, nirS-type heterotrophic denitrifiers were a sta-ble community regardless of changes in oxygen and sulfide concentrations in different seasons and with a similar abundance to that detected in other sediments not exposed to high sulfide concentrations.

Besides, nirS-type denitrifiers outnumbered anammox bacteria leading to the conclusion that anammox does not contribute significantly to the N2 removal process in Lake Grevelingen sed-iments. Apart from that, the anammox bacteria population seemed to be affected by the phys-

Page 105: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

97

Discussion/Conclusion

icochemical changes between seasons. For example, their abundance and activity was higher in lower sulfide concentrations and low carbon input, also supporting a possible inhibition of anammox bacteria by sulfide. The sulfide-dependent denitrifiers in Lake Grevelingen sediments have proven to be abundant and expected to contribute significantly to the N2-removal in these sediments. Their activity of sulfide oxidation is intuitively expected to reduce the concentration of sulfide, which is in turn toxic for organoheterotrophic denitrifiers and anammox bacteria. However, in this system the detoxification mediated by sulfide-dependent denitrifiers is either not sufficient or other factors are contributing to the low denitrification potential observed in the sediments of Lake Grevelingen.

Recent studies have also reported the presence of cable bacteria and sulfide-oxidizers of the Beggiatoaceae family in the Lake Grevelingen sediments. Both denitrifiers and anammox bacte-ria activity could be inhibited by the competition with cable bacteria and Beggiatoaceae for elec-tron donors and acceptors (such as sulfide, nitrate and nitrite) before summer hypoxia (March). During summer hypoxia (August), sulfide inhibition and low nitrate and nitrite concentrations seem to limit the activity of heterotrophic and autotrophic (sulfide-dependent) denitrifiers and anammox bacteria. Further studies also involving denitrification rate determinations will be re-quired to further assess the effects of hypoxia and high sulfide concentrations in the sediments of Lake Grevelingen.

Acknowledgements

We thank the captain and crew of the R/V Luctor (Peter Coomans and Marcel Kristalijn) for support during sampling. We thank Eric Boschker for support and discussions during the sam-pling campaigns, and Diana Vasquez, Dorina Seitaj, Fatimah Sulu-Gambari and Anton Tramper for assistance with the collection and analysis of water column and pore water data. We thank Pieter van Rijswijk, Silvia Hidalgo Martinez, Marcel van der Meer and Sandra Heinzelmann for assistance during sediment sampling. Analytical support was provided by Anchelique Mets, De-nise Dorhout, Irene Rijpstra and Elda Panoto. We thank Prof. Stefan Schouten for constructive comments on the manuscript. This work was supported by grants from the Darwin Center for Biogeosciences (grant numbers 3062 and 142.16.3092) and ERC Grant 306933 to FM. This Re-search was supported by the SIAM Gravitation Grant 024.002.002 from the Dutch Ministry of Education, Culture, and Science (OCW).

Page 106: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

98

Figure S1: Bottom water temperature [°C] (red square) and concentrations of oxygen [μM] (blue square), ammonium [μM] (dark green triangle), nitrite [μM] (pink triangle) and nitrate [μM] (light green triangle).

Figure S2: A) Relative percentage of nirS sequences of denitrifying bacteria in cluster 1 and 2 and B) in subcluster 1.1 and 1.2 according to the three stations in March and August.

Page 107: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

99

Supplementary Material

0.1

uncultured bacterium, ocean sediment, 5306 m water depth (FJ746210)uncultured bacterium, Manantial del Toro hypersaline groundwater (JF747691)

17 x Station 2, March, 0-1cm Lake Grevelingen sediment (qPCR)8 x Station 2, March, 1-2 cm Lake Grevelingen sediment (cloning)

Candidatus Scalindua brodae, anaerobic ammonium oxidizer enrichment reactor (EU142948)Candidatus Scalindua brodae (JRYO01000256)Candidatus Scalindua brodae, Black Sea (AY257181) uncultured bacterium, seafloor lavas from the East Pacific Rise (EU491842)uncultured planctomycete, Barrow Canyon sediment, Alaskan margin (DQ869385)uncultured bacterium, sediment from the Kings Bay, Svalbard, Arctic (EU050867)

uncultured planctomycete, Saanich Inlet, 200 m depth (HQ163732)uncultured planctomycete, seawater (DQ534740)uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m

uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU478646)uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU478653)uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU478687)uncultured planctomycete, seawater (DQ534739)uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU478684)uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU478669)uncultured planctomycete, seawater (DQ534742)

uncultured Planctomycetales bacterium, Peruvian OMZ; Station 4; 35 m (EU47869)uncultured Planctomycetales bacterium, Black Sea, 100 m (EU478626)uncultured Planctomycetales bacterium, Namibian OMZ; Station 202; 52 m (EU478632)uncultured Planctomycetales bacterium, Namibian OMZ; Station 202; 52 m (EU478631)uncultured planctomycete, Saanich Inlet, 120 m depth (GQ349291)uncultured Planctomycetales bacterium, Namibian OMZ; Station 202; 52 m (EU478645)uncultured Planctomycetales bacterium, Namibian OMZ; Station 202; 52 m (EU478634)uncultured Planctomycetales bacterium, Namibian OMZ; Station 202; 52 m (EU478633)uncultured planctomycete, Saanich Inlet, 120 m depth (GQ349285) uncultured planctomycete, Saanich Inlet, 120 m depth (GQ349247)uncultured planctomycete, Barrow Canyon sediment, Alaskan margin (DQ869384)

uncultured Planctomycetales bacterium, low temperature hydrothermal oxides (South West Indian Ridge) (JN860309)Candidatus Scalindua marina, surface sediment under a water column of 118 m (EF602039)Candidatus Scalindua marina, surface sediment under a water column of 118 m (EF602038)uncultured planctomycete, sea water from Gullmarsfjorden (DQ386151)Candidatus Scalindua sp. anaerobic ammonium oxidizer enrichment reactor (EU142947)uncultured Planctomycetales bacterium (AB573102)

uncultured Candidatus Scalindua sp., estuarine sediment, 0−1 cm depth (FM992882) uncultured bacterium, Yellow Sea continental shelf sediment (GQ143789)uncultured bacterium, Manantial del Toro hypersaline groundwater (JF747652)

Station 2, March, 1-2 cm Lake Grevelingen sediment (1) Station 2, March, 1-2 cm Lake Grevelingen sediment (8)

Candidatus Scalindua wagneri, rotating biological contactor (AY25488)uncultured Candidatus Scalindua sp., estuarine sediment, 0−1 cm depth (FM992881)

uncultured planctomycete, intertidal sand bank (AB300486)uncultured Candidatus Scalindua sp., laboratory−scale bioreactor (AB822932) uncultured Candidatus Scalindua sp., enrichment culture from marine sediment (AB811947) anaerobic ammonium−oxidizing planctomycete JMK−2, anammox bioreactor (AB281489)

uncultured planctomycete, marine sediment ca. 50 m (EU373086) uncultured bacterium, Manantial del Toro hypersaline groundwater (JF747633)

uncultured bacterium, subseafloor sediment of the South China Sea (EU386010)uncultured bacterium, deep−sea sediments from different depths (AB015552)

uncultured Planctomycetales bacteria, Peruvian OMZ; Station 4; 35 m

uncultured bacteria Arabian Sea; Xisha Trough (South China Sea); OMZ of Chile; hypersaline grondwater

6 x Station 2, March, 0-1 cm Lake Grevelingen sediment (qPCR) 11 x Station 2, March, 1-2 cm Lake Grevelingen sediment (cloning)

uncultured bacteria Xisha Trough (South China Sea); Manantial del Toro hypersaline grondwater

Can

did

atu

s S

calin

du

a

Figure S3: Phylogenetic tree of anammox bacteria partial 16S rRNA gene sequences retrieved in this study (21 DNA sequences recovered by amplification (PCR) and cloning and 21 DNA sequences recov-ered by amplification (qPCR) and cloning from sediments (0–1 cm) of station S2 in March) and closest relatives (bold: our sequences and closest known relatives); the scale bar indicates 10% sequence diver-gence.

Page 108: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

100

Tabl

e S1

: Prim

er p

airs

des

crib

ed in

the

text

, PC

R co

nditi

ons a

nd a

mpl

icon

size

use

d in

this

stud

y.

Ass

ayTa

rget

Prim

er p

air

Tm

[°C

]A

mpl

icon

si

ze [b

]R

ef.

qPC

R +

PC

R/cl

onin

gA

nam

mox

bac

teria

16

S rR

NA

gen

e

Brod

541F

(5’-G

AG

CA

CG

TAG

GTG

GG

TTTG

T-3’

)qP

CR

59/

PCR

5827

9Li

et a

l.,

2010

Am

x820

R (5

’-AA

AA

CC

CC

TCTA

CTT

AG

TGC

CC

-3’)

qPC

R +

PC

R/cl

onin

gSc

alind

ua sp

. nirS

gen

e

Scni

r372

F (5

’-TG

TAG

CC

AG

CA

TTG

TAG

CG

T-3’

)qP

CR/

PCR

5747

4La

m e

t al.,

20

09Sc

nir8

45R

(5’-T

CA

AG

CC

AG

AC

CC

ATT

TGC

T-3’

)

qPC

R +

PC

R/cl

onin

gD

enitr

ifyin

g ba

cter

ia

nirS

gen

e

nirS

1F (5

’-CC

TAY

TGG

CC

GC

CRC

ART

-3’)

qPC

R/PC

R 60

222

Brak

er e

t al

., 19

98ni

rS3R

(5’-G

CC

GC

CG

TCRT

GVA

GG

AA

-3’)

qPC

R+cl

onin

gSu

lfide

oxi

dizi

ng

bact

eria

apr

A g

ene

Apr

1F (5

’TG

GC

AG

ATC

ATG

ATY

MAY

GG

-3’)

qPC

R/PC

R 56

359

Mey

er &

K

ueve

r, 20

07A

pr5R

(5’-G

CG

CC

AA

CYG

GRC

CRT

A-3

’)

PCR

cond

ition

s: 95

°C 5

min

; 40×

[95°

C 1

min

, Tm

40

s, 72

°C 1

min

]; 72

°C 5

min

. qPC

R co

nditi

ons:

95°C

4 m

in; 4

0× [9

5°C

30

s, Tm

40

s, 72

°C 3

0 s]

; 80

°C 2

5 s.

Page 109: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

101

Supplementary Material

Table S2: Total organic carbon (TOC) after acidification of freeze dried sediment.

Station Sediment depth [cm]

TOC [%]

March August

1

0–1 2.86 1.81

1–2 3.67 2.47

2–3 3.83 3.52

3–4 3.98 3.11

4–5 4.28 3.66

2

0–1 3.11 2.38

1–2 3.62 3.29

2–3 3.77 3.24

3–4 3.49 3.5

4–5 3.85 3.52

3

0–1 2.87 3.04

1–2 4.02 3.9

2–3 3.89 3.56

3–4 4.01 4.39

4–5 3.63 3.73

Page 110: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

102

Tabl

e S3

: Tra

nscr

ipt (

cDN

A; m

RNA

cop

y) a

bund

ance

det

ecte

d in

sedi

men

ts o

f al

l sta

tions

in M

arch

and

Aug

ust.

Stat

ion

Sedi

men

t de

pth

[cm

]

nirS

gen

ede

nitr

ifyin

g ba

cter

ia16

S rR

NA

gen

ean

amm

ox b

acte

riaap

rA g

ene

SOB

and

SR

B

cDN

A [c

opy

g-1]

cDN

A [c

opy

g-1]

cDN

A [c

opy

g-1]

Mar

chA

ugus

tM

arch

Aug

ust

Mar

chA

ugus

t

1

0–1

1.8

× 1

02n.

d.1.

9 ×

105

8.1

× 1

05n.

d.3.

8 ×

103

1–2

1.1

× 1

04n.

d.1.

8 ×

105

1.3

× 1

03n.

d.n.

d.

2–3

n.d.

1.0

× 1

021.

0 ×

105

2.8

× 1

055.

4 ×

102

2.5

× 1

04

3–4

n.d.

n.d.

6.5

× 1

046.

9 ×

105

n.d.

1.1

× 1

04

4–5

n.d.

n.d.

2.5

× 1

042.

3 ×

105

n.d.

n.d.

2

0–1

1.9

× 1

02n.

d.8.

5 ×

104

3.0

× 1

05n.

d.n.

d.

1–2

1.5

× 1

03n.

d.6.

0 ×

105

1.1

× 1

067.

0 ×

103

7.4

× 1

02

2–3

6.7

× 1

02n.

d.9.

6 ×

105

1.7

× 1

061.

4 ×

104

3.4

× 1

03

3–4

n.d.

n.d.

1.9

× 1

051.

3 ×

105

1.6

× 1

035.

9 ×

102

4–5

n.d.

n.d.

4.8

× 1

045.

5 ×

104

n.d.

n.d.

3

0–1

1.5

× 1

02n.

d.4.

9 ×

106

2.9

× 1

05n.

d.n.

d.

1–2

n.d.

1.1

× 1

023.

1 ×

106

5.1

× 1

063.

5 ×

103

1.8

× 1

04

2–3

n.d.

n.d.

4.1

× 1

053.

1 ×

106

5.0

× 1

024.

3 ×

102

3–4

n.d.

1.8

× 1

032.

2 ×

105

1.3

× 1

051.

4 ×

103

3.7

× 1

03

4–5

n.d.

n.d.

8.5

× 1

043.

7 ×

104

n.d.

n.d.

qPC

R co

nditi

ons:

95°C

4 m

in; 4

0 ×

[95°

C 3

0 s,

Tm 4

0 s,

72°C

30

s]; 8

0°C

25

s. nd

: non

det

ecta

ble,

i.e. v

alue

s wer

e un

der t

he d

etec

tion

limit

of th

e qP

CR

appr

oach

.

Page 111: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

103

Supplementary Material

Table S4: Results of anammox bacteria PC-monoether ladderane lipid analysis at all stations in March and August (0−5 cm sediment depth).

PC-monoether ladderane lipid [ng g−1]Station Sediment depth [cm] March August

1

0–1 20.4 2.4

1–2 13.3 3.4

2–3 6.1 2.8

3–4 3.4 7

4–5 5 7

2

0–1 10.8 6.2

1–2 4.7 1.9

2–3 2 1.9

3–4 1 1.4

4–5 1 1.4

3

0–1 8.6 2.4

1–2 6.1 1.2

2–3 1.6 1.2

3–4 1.1 0.9

4–5 1.4 1.4

Page 112: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 4

104

Tabl

e S5

: Res

ults

of

anam

mox

bac

teria

FA

ME

ana

lysis

of

all s

tatio

ns in

Mar

ch a

nd A

ugus

t (0−

1 an

d 4−

5 cm

sedi

men

t dep

th).

long

cha

in F

AM

Es

[ng

g-1]

sum

Stat

ion

mon

thde

pth

C20

-[3]

C20

-[5]

C18

-[3]

C18

-[5]

C20

-[3]

–C18

-[5]

Stat

ion

1

Mar

ch0.

53.

63.

84.

05.

516

.8

Mar

ch4.

53.

55.

42.

05.

015

.8

Aug

ust

0.5

2.5

1.5

5.1

3.9

13.1

Aug

ust

4.5

1.8

1.8

1.4

2.8

7.8

Stat

ion

2

Mar

ch0.

56.

14.

76.

97.

825

.4

Mar

ch4.

54.

89.

14.

44.

422

.7

Aug

ust

0.5

5.3

5.7

19.3

8.4

38.7

Aug

ust

4.5

3.1

8.2

4.4

4.1

19.8

Stat

ion

3

Mar

ch0.

518

.916

.332

.139

.310

6.5

Mar

ch4.

53.

67.

94.

04.

019

.5

Aug

ust

0.5

4.7

4.1

8.9

9.7

27.4

Aug

ust

4.5

2.8

3.2

1.6

3.0

10.6

Page 113: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 114: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 115: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

107

5. Impact of seasonal hypoxia on activity and community structure of chemolithoautotrophic bacteria in a coastal sediment

Yvonne A. Lipsewers, Diana Vasquez-Cardenas (equal contributors), Dorina Seitaj, Regina Schauer, Silvia Hidalgo Martinez, Jaap S. Sinninghe Damsté,

Filip J.R. Meysman, Laura Villanueva, and Henricus T.S. Boschker

Seasonal hypoxia in coastal systems drastically changes the availability of electron acceptors in bottom water, which alters the sedimentary reoxidation of reduced compounds. However, the effect of seasonal hypoxia on the chemolithoautotrophic community that catalyze these reoxi-dation reactions, is rarely studied. Here we examine the changes in activity and structure of the sedimentary chemolithoautotrophic bacterial community of a seasonally hypoxic saline basin under oxic (spring) and hypoxic (summer) conditions. Combined 16S rRNA gene amplicon sequencing and analysis of phospholipid derived fatty acids indicated a major temporal shift in community structure. Aerobic sulfur-oxidizing Gammaproteobacteria (Thiotrichales) and Epsi-lonproteobacteria (Campylobacterales) were prevalent during spring, whereas Deltaproteobac-teria (Desulfobacterales) related to sulfate reducing bacteria prevailed during summer hypoxia. Chemolithoautotrophy rates in the surface sediment were three times higher in spring compared to summer. The depth distribution of chemolithoautotrophy was linked to the distinct sulfur oxidation mechanisms identified through microsensor profiling, i.e., canonical sulfur oxidation, electrogenic sulfur oxidation by cable bacteria, and sulfide oxidation coupled to nitrate reduction by Beggiatoaceae. The metabolic diversity of the sulfur-oxidizing bacterial community suggests a complex niche partitioning within the sediment probably driven by the availability of reduced sulfur compounds (H2S, S0, S2O3

2−) and electron acceptors (O2, NO3−) regulated by seasonal

hypoxia. Chemolithoautotrophic microbes in the seafloor are dependent on electron acceptors like ox-ygen and nitrate that diffuse from the overlying water. Seasonal hypoxia however drastically changes the availability of these electron acceptors in the bottom water, and hence, one expects a strong impact of seasonal hypoxia on sedimentary chemolithoautotrophy. A multidisciplinary in-vestigation of the sediments in a seasonally hypoxic coastal basin confirms this hypothesis. Our data show that bacterial community structure and the chemolithoautotrophic activity varied with the seasonal depletion of oxygen. Unexpectedly, the dark carbon fixation was also dependent on the dominant microbial pathway of sulfur oxidation occurring in the sediment (i.e., canonical sulfur oxidation, electrogenic sulfur oxidation by cable bacteria, and sulfide oxidation coupled to nitrate reduction by Beggiatoaceae). These results suggest that a complex niche partitioning within the sulfur-oxidizing bacterial community additionally affects the chemolithoautotrophic community of seasonally hypoxic sediments.

Applied and Environmental Microbiology 83, e03517-16, 2017

Page 116: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

108

5.1. Introduction The reoxidation of reduced intermediates formed during anaerobic mineralization of organic matter is a key process in the biogeochemistry of coastal sediments (Soetaert et al., 1996, Jør-gensen and Nelson, 2004). Many of the microorganisms involved in the reoxidation of reduced compounds are chemolithoautotrophs, which fix inorganic carbon using the chemical energy derived from reoxidation reactions (dark CO2 fixation). In coastal sediments, sulfate reduction forms the main respiration pathway, accounting for 50% to 90% of the organic matter min-eralization (Soetaert et al., 1996). The reoxidation of the pool of reduced sulfur compounds produced during anaerobic mineralization (dissolved free sulfide, thiosulfate, elemental sulfur, iron monosulfides and pyrite) hence forms the most important pathway sustaining chemolitho-autotrophy in coastal sediments (Jørgensen and Nelson, 2004, Howarth, 1984).

Various lineages from the Alpha-, Gamma-, Delta- and Epsilonproteobacteria, including re-cently identified groups, such as particle-associated Gammaproteobacteria and large sulfur bac-teria, couple dark CO2 fixation to the oxidation of reduced sulfur compounds in oxygen defi-cient marine waters and sediments, in coastal marine sediments and in lake sediments (Brüchert et al., 2003; Lavik et al., 2009; Sorokin et al., 2011; Grote et al., 2012; Jessen et al., 2016; Ye et al., 2016). Both chemolithoautotrophic sulfur-oxidizing Gammaproteobacteria and sulfur dis-proportionating Deltaproteobacteria have been identified to play a major role in the sulfur and carbon cycling in diverse intertidal sediments (Lenk et al., 2010; Boschker et al., 2014; Dyksma et al., 2016). Hence, chemolithoautotrophic sulfur-oxidizing communities vary between sediment environments, but it is presently not clear as to which environmental factors are actually deter-mining the chemolithoautotrophic community composition at a given site.

Seasonal hypoxia is a natural phenomenon that occurs in coastal areas around the world (Diaz and Rosenberg, 2008) and provides an opportunity to study the environmental factors con-trolling sedimentary chemolithoautotrophy. Hypoxia occurs when bottom waters become de-pleted of oxygen (< 63 µmol O2 L

–1), and has a large impact on the biogeochemical cycling and ecological functioning of the underlying sediments (Diaz and Rosenberg, 2008). The reduced availability or even absence of suitable soluble electron acceptors (O2, NO3

–) in the bottom water during part of the year should in principle result in a reduction or complete inhibition of sedimentary sulfur reoxidation, and hence, limit chemolithoautotrophy. At present, the preva-lence and temporal variations of chemolithoautotrophy have not been investigated in coastal sediments of seasonal hypoxic basins.

Likely, the availability of soluble electron acceptors (O2, NO3–) in the bottom water is not

the only determining factor of chemolithoautotrophy. A recent study conducted in a seasonally hypoxic saline basin (Lake Grevelingen, The Netherlands) indicated that the intrinsic structure and composition of the sulfur oxidizing microbial community also determined the biogeochem-istry of the sediment (Seitaj et al., 2015). In this study, three distinct microbial sulfur oxidation mechanisms were observed throughout a seasonal cycle: (1) electrogenic sulfur oxidation by heterotrophic cable bacteria (Desulfobulbaceae); (2) canonical aerobic oxidation of free sulfide at the oxygen-sulfide interface and, (3) sulfide oxidation coupled to nitrate reduction by filamen-tous members of the Beggiatoaceae family that store nitrate intracellularly. The consequences of these three mechanisms on the chemolithoautotrophic community have however not been studied. The first sulfur oxidizing mechanism has been shown to affect the chemolithoauto-trophic community in sediments only under laboratory conditions (Vasquez-Cardenas et al.,

Page 117: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

109

Introduction/Material and Methods

2015) while the third mechanisms may directly involve chemolithoautotrophic Beggiatoaceae as some species are known to grow autotrophically (Nelson and Jannasch, 1983). Accordingly, we hypothesize that the presence of these sulfur oxidation regimes as well as the depletion of O2 and NO3

– will result in a strong seasonality in both the chemolithoautotrophic activity and community structure under natural conditions.

To examine the above-mentioned hypothesis, we conducted a multidisciplinary study with intact sediments of Lake Grevelingen, involving both geochemistry and microbiology. Field sampling was conducted during spring (oxygenated bottom waters) and summer (oxygen deplet-ed bottom waters). The dominant sulfur oxidation mechanism was geochemically characterized by sediment microsensor profiling (O2, H2S, pH) whereas the abundance of cable bacteria and Beggiatoaceae was performed with fluorescence in situ hybridization (FISH). General bacterial diversity was assessed by 16S rRNA gene amplicon sequencing and the analysis of phospholipid derived fatty acids (PLFA). PLFA analysis combined with 13C stable isotope probing (PLFA-SIP) provided the activity and community composition of chemolithoautotrophs in the sediment. This approach was complemented by the analysis of genes involved in dark CO2 fixation, i.e., characterizing diversity and the abundance of the genes cbbL and aclB that code for key enzymes in Calvin-Benson Bassham (CBB) and reductive tricarboxylic acid (rTCA) carbon fixation path-ways. This multidisciplinary research showed strong temporal and spatial shifts of the chemo-lithoautotrophic composition and activity in relation to the seasonal hypoxia and the main sulfur oxidation mechanisms present in the sediments of the marine Lake Grevelingen.

5.2. Material and methods

5.2.1. Study site and sediment sampling Lake Grevelingen is a former estuary located within the Rhine-Meuse-Scheldt delta area of the Netherlands, which became a closed saline reservoir (salinity ~30) by dam construction at both the land side and sea side in the early 1970s. Due to an absence of tides and strong currents, Lake Grevelingen experiences a seasonal stratification of the water column, which in turn, leads to a gradual depletion of the oxygen in the bottom waters (Hagens et al., 2015). Bottom water oxygen at the deepest stations typically starts to decline in April, reaches hypoxic conditions by end of May (O2 < 63 µM), further decreasing to anoxia in August (O2 <0.1 µM), while the re-ox-ygenation of the bottom water takes place in September (Seitaj et al., 2015).

To study the effects of the bottom water oxygenation on the benthic chemolithoautotrophic community we performed a field sampling campaign on March 13th, 2012 (before the start of the annual O2 depletion) and August 20th, 2012 (at the height of the annual O2 depletion). Detailed water column, pore water and solid sediment chemistry of Lake Grevelingen over the year 2012 have been previously reported (Seitaj et al., 2015; Hagens et al., 2015; Sulu-Gambari et al., 2016). Sediments were recovered at three stations along a depth gradient within the Den Osse basin, one of the deeper basins in Marine Lake Grevelingen: Station 1 (S1) was located in the deepest point (34 m) of the basin (51.747°N, 3.890°E), Station 2 (S2) at 23 m depth (51.749°N, 3.897°E) and Station 3 (S3) at 17 m depth (51.747°N, 3.898°E). Intact sediment cores were retrieved with a single core gravity corer (UWITEC) using PVC core liners (6 cm inner diameter, 60 cm length). All cores were inspected upon retrieval and only cores with a visually undisturbed surface were used for further analysis.

Page 118: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

110

Thirteen sediment cores for microbial analysis were collected per station and per time point: two cores for phospholipid-derived fatty acid analysis combined with stable isotope probing (PLFA-SIP), two cores for nucleic acid analysis, four cores for 13C-bicarbonate labeling, three cores for microelectrode profiling, one core for quantification of cable bacteria (see supple-ment), and one core for quantification of Beggiatoaceae (see supplement). Sediment cores for PLFA extractions were sliced manually on board the ship (5 sediment layers; sectioning at 0.5, 1, 2, 4, and 6 cm depth). Sediment slices were collected in Petri dishes, and replicate depths were pooled and thoroughly mixed. Homogenized sediments were immediately transferred to centrifuge tubes (50 mL) and placed in dry ice until further analysis. Surface sediments in August consisted of a highly porous “fluffy” layer that was first left to settle after core retrieval. After-wards, the top 1 cm thick layer was recovered through suction (rather than slicing). Sediment for nucleic acid analysis was collected by slicing manually at a resolution of 1 cm up to 5 cm depth. Sediment samples were frozen in dry ice, transported to the laboratory within a few hour and placed at −80°C until further analysis.

5.2.2. Microsensor profiling Oxygen depth profiles were recorded with commercial microelectrodes (Unisense, Denmark; tip size: 50 µm) at 25−50 µm resolution. For H2S and pH (tip size: 50 and 200 µm), depth profiles were recorded at 200 µm resolution in the oxic zone, and at 400 or 600 µm depth resolution be-low. Calibrations for O2, pH and H2S were performed as previously described (Seitaj et al., 2015; Malkin et al., 2014). ΣH2S was calculated from H2S based on pH measured at the same depth using the R package AquaEnv (Hofmann et al., 2010). The oxygen penetration depth (OPD) is operationally defined as the depth below which [O2] < 1 µM, while the sulfide appearance depth (SAD) is operationally defined as the depth below which [H2S] > 1 µM. The diffusive oxygen uptake (DOU) was calculated from the oxygen depth profiles as previously described in detail (Vasquez-Cardenas et al., 2015).

5.2.3. DNA extraction and 16S rRNA gene amplicon sequencing DNA from 0 cm to 5 cm sediment depth (in 1 cm resolution) was extracted using the DNA PowerSoil® Total Isolation Kit (Mo Bio Laboratories, Inc., Carlsbad, CA). Nucleic acid con-centrations were quantified spectrophotometrically (Nanodrop, Thermo Scientific, Wilmington, DE) and checked by agarose gel electrophoresis for integrity. DNA extracts were kept frozen at –80°C.

Sequencing of 16S rRNA gene amplicons was performed on the first cm of the sediment (0–1 cm depth) in all stations in March and August as described before (Moore et al., 2015). Further details are provided in the SI. The 16S rRNA gene amplicon reads (raw data) have been deposited in the NCBI Sequence Read Archive (SRA) under BioProject number PRJNA293286. The phylogenetic affiliation of the 16S rRNA gene sequences was compared to release 119 of the SILVA NR SSU Ref database (http://www. arb-silva.de/; Quast et al., 2013) using the ARB software package (Ludwig et al., 2004). Sequences were added to a reference tree generated from the Silva database using the ARB Parsimony tool.

5.2.4. Sediment incubations and PLFA-SIP analysis Sediment cores were labeled with 13C-bicarbonate to determine the chemolithoautotrophic activ-ity and associated bacterial community by tracing the incorporation of 13C into bacterial PLFA.

Page 119: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

111

Material and Methods

To this end, four intact cores were sub-cored with smaller core liners (4.5 cm inner diameter; 20 cm height). In situ bottom water was kept over the sediment and no gas headspace was pres-ent. Cores were kept inside a closed cooling box during transport to the laboratory.

Stock solutions of 80 mM of 13C-bicarbonate (99% 13C; Cambridge Isotope Laboratories, Andover, Ma, USA) were prepared as previously described (Boschker et al., 2014). 13C-bicarbon-ate was added to the sediment from 3 cm above the surface to 8 cm deep in the sediment cores in aliquots of 100 μL through vertically aligned side ports (0.5 cm apart) with the line injection method (Jørgensen, 1978). March sediments were incubated at 17±1°C in the dark for 24 h and continuously aerated with 13C-saturated air to maintain 100% saturated O2 conditions as those found in situ, but avoiding the stripping of labeled CO2 from the overlying water (Vasquez-Carde-nas et al., 2015). In August, sediments were incubated at 17±1°C in the dark for 40 h to ensure sufficient labeling as a lower activity was expected under low oxygen concentrations. In August, oxygen concentrations in the overlying water were maintained near in situ O2 levels measured in the bottom water (S1: 0−4% saturation, S2: 20−26%, S3: 35−80%). A detailed description of aeriation procedures can be found in the SI.

At the end of the incubation period, sediment cores were sliced in five depth intervals (as described for PLFA sediment cores sliced on board) thus obtaining two replicate slices per sedi-ment depth and per station. Sediment layers were collected in centrifuge tubes (50 mL) and wet volume and weight were noted. Pore water was obtained by centrifugation (4500 rpm for 5 min) for dissolved inorganic carbon (DIC) analysis, and sediments were lyophilized for PLFA analysis. Biomarker extractions were performed on freeze dried sediment as described before (Guckert et al., 1985) and 13C-incorporation into PLFA was analyzed as previously reported (Boschker et al., 2014). Nomenclature of PLFA can be found in the SI. Detailed description of the PLFA calcu-lations can be found in the literature (Vasquez-Cardenas et al., 2015; Boschker and Middelburg, 2002). Total dark CO2 fixation rates (mmol C m–2 d–1) are the depth-integrated rates obtained from the 0−6 cm sediment interval.

5.2.5. Quantitative PCR To determine the abundance of chemolithoautotrophs we quantified the genes coding for two enzymes involved in dark CO2 fixation pathways: the large subunit of the RubisCO enzyme (ribulose 1,5-bisphosphate carboxylase/oxygenase) cbbL gene, and the ATP citrate lyase aclB gene. Abundance of the RubisCO cbbL gene was estimated by using primers K2f/V2r, specific for the forms IA and IC of the RuBisCO form I large subunit gene cbbL which is present in obligately and facultatively lithotrophic bacteria (Nanba et al., 2004; Nigro and King, 2007). The abundance of ATP citrate lyase aclB gene was quantified by using the primer set aclB_F/aclB_R, which is based on the primers 892F/1204R (Campbell et al., 2003), specific for the ATP citrate lyase β gene of chemoautotrophic bacteria using the rTCA pathway, with several nucleotide differences introduced after aligning n=100 sequences of aclB gene fragments affiliated to Epsi-lonproteobacteria (See Table S1 for details).

The quantification of cbbL and aclB genes via quantitative PCR (qPCR) was performed in 1 cm resolution for the sediment interval between 0–5 cm depth in all stations in both March and August. qPCR analyses were performed on a Biorad CFX96TM Real-Time System/C1000 Thermal cycler equipped with CFX ManagerTM Software. All qPCR reactions were performed in triplicate with standard curves from 100 to 107 molecules per microliter. Standard curves and

Page 120: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

112

qPCR reactions were performed as previously described (Chapter 2). Melting temperatures (Tm) are listed in Table S1, qPCR efficiencies (E) for aclB gene and cbbL gene amplifications were 70 and 82%, respectively. Correlation coefficients for standard curves were ≥ 0.994 for aclB gene and ≥ 0.988 for cbbL gene amplification.

5.2.6. PCR amplification and cloning Amplifications of the RubisCO cbbL gene and the ATP citrate lyase aclB gene were performed with the primer pairs specified in Table S1. The PCR reaction mixture was the following (final concentration): Q-solution (PCR additive, Qiagen, Valencia, CA) 1 ×; PCR buffer 1 ×; BSA (200 µg mL–1); dNTPs (20 µM); primers (0.2 pmol µL–1); MgCl2 (1.5 mM); 1.25 U Taq poly-merase (Qiagen, Valencia, CA). PCR conditions for these amplifications were the following: 95°C, 5 min; 35 × [95°C, 1 min; Tm, 1 min; 72°C, 1 min]; final extension 72°C, 5 min. PCR products were gel purified (QIAquick gel purification kit, Qiagen, Valencia, CA) and cloned in the TOPO-TA cloning® kit (Life Technologies, Carlsbad, CA) and transformed in E. coli TOP10 cells following the manufacturer’s recommendations. Recombinant plasmid DNA was sequenced using M13R primer by BASECLEAR (Leiden, The Netherlands).

Sequences were aligned with MEGA5 software (Tamura et al., 2011) by using the alignment method ClustalW. The phylogenetic trees of the cbbL and aclB genes were computed with the Neighbour-Joining method (Saitou and Nei, 1987). The evolutionary distances were estimated using the Jukes-Cantor method (Jukes and Cantor, 1969) for DNA sequences and with the Poisson correction method for protein sequences (Zuckerkandl and Pauling, 1965) with a boot-strap test of 1,000 replicates. Sequences were deposited in NCBI with the following accession numbers: KT328918–KT328956 for cbbL gene sequences and KT328957–KT329097 for aclB gene sequences. Further details on experimental procedures and methods are found in the Sup-plementary Information.

5.3. Results

5.3.1. Geochemical characterization The seasonal variation of the bottom water oxygen concentration in Lake Grevelingen strongly influenced the porewater concentrations of O2 and H2S. In March, bottom waters were fully oxygenated at all stations (299–307 μM), oxygen penetrated 1.8–2.6 mm deep in the sediment, and no free sulfide was recorded in the first few centimeters (Table 1). The width of the suboxic zone, operationally defined as the sediment layer located between the oxygen penetration depth (OPD) and the sulfide appearance depth (SAD), varied between 16–39 mm across the three stations in March 2012. In contrast, in August, oxygen was strongly depleted in the bottom waters at S1 (< 0.1 μM) and S2 (11 μM), and no O2 could be confidently detected by microsen-sor profiling in the surface sediment at these two stations. At station S3, the bottom water O2 remained higher (88 μM), and oxygen penetrated down to 1.1 mm. In August, free sulfide was present near the sediment-water interface at all three stations, and the accumulation of sulfide in the pore water increased with water depth (Fig. 1a). Depth pH profiles showed much larger variation between stations in March compared to August (Fig. 1a). The pH profiles in S1 and S3 in March were similarly characterized by highest values in the oxic zone and low pH values (pH < 6.5) in the suboxic zone. The pH depth profile in S2 showed an inverse pH profile with a pH minimum in the oxic layer and a subsurface maximum below. The pH profiles at S1 and S3

Page 121: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

113

Material and Methods/Results

in August 2012 showed a gradual decline of pH with depth, while the pH profile at S2 in August was more or less constant with depth.

5.3.2. Bacterial diversity by 16S rRNA gene amplicon sequencingThe general diversity of bacteria was assessed by 16S rRNA gene amplicon sequencing anal-ysis, which was applied to the surface sediments (0−1 cm) of all stations in both March and August. Approximately 50% of the reads were assigned to three main clades: Gamma, Delta-, and Epsilonproteobacteria (Fig. 2). The remaining reads were distributed amongst the orders Bacteroidetes (14%), Planctomycetes (6%), Alphaproteobacteria (3%), other orders (20%), the candidate phylum WS3 (2%) and unassigned (5%) (given as the average of the three stations and both seasons).

Reads classified within the Gammaproteobacteria were more abundant in March during ox-ygenated bottom water conditions than in August (Fig. 2). The majority of these reads were assigned to the orders Alteromonadales, Chromatiales and Thiotrichales. The first order includes chemoheterotrophic bacteria that are either strict aerobes or facultative anaerobes (Bowman and McMeekin, 2005). Phylogenetic comparison revealed that the reads assigned to the Chromatiales group were closely related to the Granulosicoccaceae, Ectothiorhodospiraceae, and Chroma-tiaceae families (Fig. S1). Reads falling in the Thiotrichales group were closely related to sul-fur-oxidizing bacteria from the Thiotrichaceae family with 30% of the sequences related to the genera “Candidatus Isobeggiatoa”, “Ca. Parabeggiatoa” and Thiomargarita (Fig. S2). It has been recently proposed that the genera Beggiatoa, Thiomargarita and Thioploca should be reclassified into the originally published monophyletic family of Beggiatoaceae (Salman et al., 2011, 2013), and so here, these genera will be further referred to as Beggiatoaceae. Most of the reads assigned to the Beggiatoaceae came from station S2 in spring, whereas the percentage of reads assigned to sulfur oxidizers from the order Thiotrichales decreased in August when oxygen concentrations in the bottom water were low.

Within the Deltaproteobacteria, reads were assigned to the orders Desulfarculales and Desul-fobacterales (Fig. S3). Reads within the order of Desulfobacterales were mainly assigned to the families Desulfobacteraceae (between 10−20%) and Desulfobulbaceae (~5%) (Fig. 2). Addition-al phylogenetic comparison revealed that within the Desulfobulbaceae, 60% of reads, obtained from the three stations in both seasons, clustered within the genus Desulfobulbus, of which 30% were related to the electrogenic sulfur-oxidizing cable bacteria (Fig. S4). Overall, the relative abundance of reads assigned to the Desulfobulbaceae family was similar between the two sea-sons in S2 and S3, whereas in S1 the percentage of reads was approximately 4.5-fold higher in March compared to August (Fig. 2). In contrast, the relative abundance of reads assigned to the Desulfarculales and Desulfobacteraceae increased in August during hypoxia (Fig. 2), and phylogenetic comparison revealed typical sulfate reducer genera Desulfococcus, Desulfosarcina, and Desulfobacterium, indicative of anaerobic metabolism (Figs. S5a–c).

All Epsilonproteobacteria reads were assigned to the order Campylobacterales, such as the Campylobacteraceae and Helicobacteraceae families. Phylogenetic comparison showed that the reads were closely related to the genera Sulfurovum, Sulfurimonas, Sulfurospirillum, Arcobacter (Figs. S6a–d), all capable of sulfur oxidation with oxygen or nitrate (Campbell et al., 2006). The per-centage of reads assigned to the Epsilonproteobacteria in August was lower than those in March, with highest numbers present in S2 in March (~6%), (Fig. 2).

Page 122: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

114

Tabl

e 1:

Geo

chem

ical

cha

ract

eriz

atio

n, c

hem

oaut

otro

phy

rate

s, an

d qu

antifi

catio

n of

cab

le b

acte

ria a

nd B

eggi

atoa

ceae

in th

e th

ree

stat

ions

in L

ake

Gre

velin

gen

for s

prin

g (M

arch

) and

sum

mer

(Aug

ust).

Mon

th/

Stat

ion

Bot

tom

wat

er *

con

c.

[μM

]D

OU

OPD

SA

DSu

boxi

c zo

ne**

pH

sign

atur

e***

Che

mol

itho-

auto

trop

hy

[mm

ol C

m-2

d-1]

Cab

le

biov

olum

e

[mm

3 cm

-2]

Beg

giat

oa

biov

olum

e

[mm

3 cm

-2]

°CO

2 N

O3–

[mm

ol O

2 m

-2 d

-1]

[mm

][m

m]

[mm

]

M

15

299

(oxi

c)28

.218

.2±

1.7

1.8±

0.04

17.5

±0.

716

e-SO

x3.

1±0.

52.

550.

02

25

301

(oxi

c)27

.915

.8±

3.1

2.6±

0.65

21.3

±2.

519

Nitr

ate-

stor

ing

Begg

iato

a1.

9±0.

1N

D0.

11

35

307

(oxi

c)27

.717

.1±

5.7

2.4±

0.4

41.8

±8.

639

e-SO

x1.

4±0.

32.

080.

05

A

117

0(a

noxi

c)1.

70

00.

9±1.

10.

9

Sulfa

te

redu

ctio

n/

cano

nica

l su

lfur

oxid

atio

n

0.2±

0.1

0.11

0.00

1

217

12(h

ypox

ic)

11.6

00

0.6±

00.

60.

8±0.

3N

D3.

24

319

88(h

ypox

ic)

10.6

13.9

±2.

11.

1±0.

14.

2±2.

73

1.1±

0.5

0.00

30.

004

M -

Mar

ch, A

- A

ugus

t, D

OU

: diss

olve

d O

2 up

take

; OPD

: O2

pene

tratio

n de

pth;

SA

D: ∑

H2S

app

eara

nce

dept

h; e

-SO

x: e

lect

roge

nic

sulfu

r oxi

datio

n;

ND

: not

det

erm

ined

. *Bo

ttom

wat

er is

cla

ssifi

ed a

s ano

xic

with

O2

conc

entra

tion

belo

w 1

μM

and

hyp

oxic

bel

ow 6

3 μM

. **T

he th

ickn

ess o

f th

e su

boxi

c zo

ne is

defi

ned

as t

he a

vera

ge S

AD

min

us t

he a

vera

ge O

PD. *

**pH

sig

natu

re s

erve

s to

indi

cate

the

sul

fur

oxid

izin

g m

echa

nism

tha

t do

min

ates

the

po

rew

ater

che

mist

ry a

s des

crib

ed b

y Se

itaj e

t al.

(201

5). C

able

- ca

ble

bact

eria

.

Page 123: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

115

Results

Figu

re 1

: Geo

chem

ical

fing

erpr

int,

(b) c

hem

oaut

otro

phy

dept

h pr

ofile

s, (c

) bio

volu

me

of fi

lam

ento

us B

eggi

atoa

ceae

(bla

ck) a

nd c

able

bac

teria

(blu

e)

in se

dim

ent o

f La

ke G

reve

linge

n fo

r Mar

ch a

nd A

ugus

t in

all t

hree

stat

ions

(not

e ch

ange

in sc

ale

for B

eggi

atoa

ceae

bet

wee

n M

arch

and

Aug

ust).

S1:

st

atio

n 1

(34

m),

S2: s

tatio

n 2

(23

m),

S3: s

tatio

n 3

(17

m).

Page 124: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

116

5.3.3. Bacterial community structure by phospholipid derived fatty acid analysis The relative concentrations of phospholipid derived fatty acid (PLFA) were determined to a depth of 6 cm and were analyzed by principal component analysis (PCA) to determine differ-ences in bacterial community structure (Fig. 3). A total of 22 individual PLFA, each contributing more than 0.1% to the total PLFA biomass, were included in this analysis. Samples with low total PLFA biomass (less than one standard deviation below the mean of all samples) were excluded. The PCA analysis indicated that 73% of the variation within the dataset was explained by the first two principal components (PC). While PC1 correlated with sediment depth (particularly for the March samples), PC2 clearly exposed differences in the bacterial community structure between seasons (Fig. 3a). Surface sediments (0–1 cm) were characterized by high relative con-centrations of C16 and C18 monounsaturated PLFA. In contrast, ai15:0, i17:1ω7c, and 10Me16:0 were more abundant in deeper sediments (Fig. S7a). The surface sediment showed an increased contribution of iso, anteiso and branched PLFAs in August relative to March, and this was more similar to the deeper sediments from March (Fig. 3a). Nonetheless, August sediments also showed a higher abundance of 16:0, 14:0, and 18:1ω9c compared to the deeper sediment layers in March (Fig. S7a).

5.3.4. Chemolithoautotrophic activity and community by 13C-phospholipid fatty acids analysisIncorporation of 13C-labeled dissolved inorganic carbon was found in bacterial PLFAs after 24 to 40 hours of incubation (Fig. S7b). Depth integrated dark CO2 fixation rates based on 13C-in-corporation (Table 1) showed a significant difference between seasons and stations (p = 0.0005). In March, chemolithoautotrophy increased with water depth, while in August, the opposite trend was observed. The dark CO2 fixation rate in March for S1 was the highest across all seasons and stations, but dropped by one order of magnitude in August. In contrast, at the intermediate station (S2), the dark CO2 fixation rate was only two times higher in March compared to August, while in the shallowest station (S3) the depth integrated rates were not significantly different between seasons (p = 0.56).

The depth distribution of the chemolithoautotrophic activity also differed between seasons (p = 0.02; Fig. 1b). In August, the chemolithoautotrophic activity was restricted to the upper cm of the sediment at all stations while in March, chemolithoautotrophic activity was measured deep into the sediment (up to 4 cm). In stations S1 and S3 during March, the activity depth profile showed high activities down to 4 cm, whereas at S2 chemolithoautotrophy only extended down to 2 cm with highest activity in the top 1 cm. This distinction between the depth distri-butions of chemolithoautotrophy at S2 versus S1/S3 in March correlated with the distinct pH profiles observed for S2 versus S1/S3 (Fig. 1c).

The PLFA 13C-fingerprints were analyzed by PCA to identify differences in chemolithoauto-trophic community. Only PLFA that contributed more than 0.1% to the total 13C-incorporation and sediment layers that showed chemolithoautotrophy rates higher than 0.01 μmol C cm–3 d–1 were taken into account in this analysis. Because chemolithoautotrophy rates were low in Au-gust, the PCA analysis mainly analyzed the chemolithoautotrophic communities in March (Fig. 3b). Within the dataset, 64% of the variation was explained by two principal components. PC1 revealed a clear differentiation between station S2, and stations S1 and S3 (Fig. 3b) in agreement with the distinct pH profiles described for March (Fig. 1a). S2 sediment horizons showed a

Page 125: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

117

Results

Figure 2: Percentages of total bacterial 16S rRNA gene reads in stations S1 (yellow), S2 (blue) and S3 (red) in March (filled) and August (hatched). Classified bacterial phyla, classes and orders > 3% of the total bacteria reads (in March or August) are reported (exception: family Thiotrichaceae < 3%), *including cable bacteria, **including Beggiatoaceae.

Page 126: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

118

higher contribution of monounsaturated C16 and C18 fatty acids, whereas sediments from S1 and S3 revealed an increased 13C-incorporation into fatty acids with iso and anteiso C15 and saturated C14 PLFA (Fig. S7c), together these results indicate distinct chemolithoautotrophic microbial assemblages between S2 and S1/S3. On the contrary, the three sediment samples from August (that had sufficient 13C-incorporation for the analysis) did not cluster in the PCA analysis and exhibited divergent PLFA profiles, thus the chemolithoautotrophic bacterial community for Au-gust could not be characterized further based on PLFA analysis.

5.3.5. Chemolithoautotrophic carbon fixation pathways Various CO2 fixation pathways are used by autotrophic bacteria (for detailed reviews see Berg, 2011, Hügler and Sievert, 2011). The CBB pathway is utilized by cyanobacteria and many aerobic or facultative aerobic proteobacteria of the alpha, beta and gamma subgroups whereas the rTCA pathway operates in anaerobic or microaerobic members of phyla such as Chlorobi, Aquificae, proteobacteria of the Delta- and Epsilon- subgroups and Nitrospirae (Berg, 2011). The spatial and temporal distribution of bacteria possessing these two autotrophic carbon fixation pathways was studied by quantifying the abundance of cbbL gene (CBB pathway) (Nigro and King, 2007) and aclB gene (rTCA pathway) (Campbell et l., 2006, modified in this study) by quantitative PCR down to 5 cm sediment depth (Fig. 4).The diversity of cbbL and aclB gene sequences obtained from the surface sediments (0–1 cm) was analyzed (Figs. 5a, b).

Significant differences were found in the abundance of cbbL and aclB genes (p = 5.7 × 10–7) and between season (p = 0.002), but not between stations (Fig. 4). The abundance of the cbbL gene copies was at least 2-fold higher than that of the aclB gene in March and August in all sta-tions (p = 5×10–5). In March, the depth profiles of cbbL gene showed a similar trend in stations

Figure 3: Principal component analysis (PCA) of relative PLFA concentrations (a) and of 13C-incorpora-tion into PLFA (b) in sediments from the three stations in March (no border) and August (black border). The percentage of variability explained by the first two principal components (PC) is indicated on each axis. Red: station S1, Yellow: S2, Blue S3. Symbols indicate sediment depth as seen in plot.

Page 127: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

119

Results

S1 and S3 with a decreasing gene abundance from the surface towards deeper layers, but depth integrated abundance of the cbbL gene was more than two-fold higher in S1 than in S3 (Fig. 4a). The depth distribution of the cbbL gene in the deepest station (S1) differed between seasons (p = 0.004), with lower gene copy abundances in the top 4 cm in August (Fig. 4a). In contrast, in the shallow station (S3), which experiences less seasonal fluctuations in bottom water O2, the cbbL gene copy number did not differ significantly between seasons (p = 0.4), except for a sharp decrease in the top cm of the sediment. In station S2, abundance and depth distribution of cbbL gene copies was similar between the two seasons (p = 0.3). All detected cbbL gene sequences clustered with uncultured Gammaproteobacteria clones (Fig. 5a), assigned to the orders Chro-matiales and Thiotrichales (both cbbL1 and cbbL2 clusters) reported in intertidal sediment from Lowes Cove, Maine (Nigro and King, 2007).

The abundance of the aclB gene significantly differed between months (p = 0.0004) and sta-tions (p = 0.04). Similar depth profiles of the aclB gene were detected in stations S1 and S3, with subsurface maxima in deeper sediments (at 3–4 cm and 2–3 cm deep for S1 and S3 respectively, Fig. 4b). Station S2 showed the highest aclB gene abundance in March, which remained constant with sediment depth (p = 0.86). In August, aclB gene abundance decreased substantially in S2 (p < 0.001). All stations showed a uniform distribution of the aclB gene abundance with depth in August. Bacterial aclB gene sequences in surface sediments (0−1 cm) of all stations were predominantly related to aclB sequences of Epsilonproteobacteria (Fig. 5b). Within the Epsi-lonproteobacteria, sequences clustered in six different subclusters and were mainly affiliated to bacteria in the order of Campylobacterales, i.e., to the genera Sulfuricurvum, Sulfurimonas, Thiovu-

Figure 4: Abundance [copy g−1] of two carbon fixation pathway genes with sediment depth [cm] for S1 (left), S2 (middle) and S3 (right) (a) RuBisCO cbbL gene (b) ATP citrate lyase aclB gene (black: March, white: August). The error bars represent the standard deviations of qPCR replicates.

Page 128: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

120

Station 1 August (seq19); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq19); 0-1 cm Lake Grevelingen sediment

uncultured proteobacterium clone LCS147; intertidal marine surface sediment (DQ683601)Station 2 August (seq4); 0-1 cm Lake Grevelingen sediment

Station 3 August (seq18); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq16); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCB16; intertidal marine sediment, burrow (DQ683650)

uncultured proteobacterium clone LCS69; intertidal marine surface sediment, (DQ683590)uncultured proteobacterium clone LCS120; intertidal marine surface sediment, (DQ683598)

Station 1 March (seq4); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq27); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq20); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq22); 0-1 cm Lake Grevelingen sediment

uncultured proteobacterium clone LCS15; intertidal marine surface sediment, (DQ683585)uncultured proteobacterium clone LCSS223; intertidal marine subsurface sediment, (DQ683644)

Station 3 August (seq13); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq15); 0-1 cm Lake Grevelingen sediment

uncultured proteobacterium clone LCSS; intertidal marine subsurface sediment, (DQ683627)uncultured proteobacterium clone LCSS21; intertidal marine subsurface sediment, (DQ683641)

Station 3 August (seq20); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCB173; intertidal marine sediment, burrow (DQ683661)Station 1 August (seq16); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq23); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq23); 0-1 cm Lake Grevelingen sediment

Station 2 August (seq24); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq17); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq26); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCSS9 (intertidal marine subsurface sediment (DQ683619)Station 2 March (seq18); 0-1 cm Lake Grevelingen sediment

Station 1 August (seq17); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCS155 (intertidal marine surface sediment (DQ683603)

Station 3 August (seq14); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCS42 (intertidal marine surface sediment (DQ683588)

Station 3 August (seq15); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq19); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq21); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq26); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq17); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq19); 0-1 cm Lake Grevelingen sediment

strain ALJD (AY914807)Thioalkalivibrio denitrificansstrain Hl17 (EF199949)Thioalkalivibrio halophilus

sp. K90mix (CP001905)Thioalkalivibriostrain Arh2 (AY914804)Thioalkalivibrio thiocyanoxidans

strain Al2 (AY914800)Thioalkalivibrio versutusThioploca ingrica (AP014633)

sp. PS (ABBZ01000395)Beggiatoastrain HL 5 cbbL-1 gene (DQ390451)Thiomicrospira halophila

sp. Lilliput-1 (AM404076)ThiomicrospiraXCL-2 (CP000109)Thiomicrospira crunogena

cbbL-1 gene (DQ272534)Thiomicrospira crunogenacbbL-1 gene (DQ272531)Thiomicrospira kuenenii

cbbL-1 gene (D43621)Hydrogenovibrio marinusHydrogenovibrio marinus (AB122069)

Station 2 March (seq25); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq24); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq16); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCSS153; intertidal marine subsurface sediment (DQ683632)

Station 1 August (seq18); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq28); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq24); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq24); 0-1 cm Lake Grevelingen sediment

Station 1 August (seq15); 0-1 cm Lake Grevelingen sedimentuncultured proteobacterium clone LCS90 (intertidal marine surface sediment (DQ683592)Station 3 March (seq21); 0-1 cm Lake Grevelingen sediment Station 3 March (seq25); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq27); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq18); 0-1 cm Lake Grevelingen sediment

strain HL 5 cbbL-2 gene (DQ390452)Thiomicrospira halophilastrain Arh1 (AY914806)Thioalkalivibrio paradoxus

XCL-2 (CP000109)Thiomicrospira crunogenacbbL-2 gene (DQ272533)Thiomicrospira crunogena

cbbL-2 gene (DQ272532)Thiomicrospira kueneniisp. Lilliput-1 (AM404077)Thiomicrospira

cbbL-2 gene (AB122070)Hydrogenovibrio marinuscbbL-2 gene (D43622)Hydrogenovibrio marinus

Thiomicrospira pelophila (DQ272530)sp. V2501 cbbL-2 gene (AB634593)Thiomicrospira

Al3 (CP007030)Thioalkalimicrobium aerophilumALM1 (CP002776)Thioalkalimicrobium cyclicum

Thioalkalimicrobium sibiricum (DQ272529)Thioalkalimicrobium aerophilum (DQ272527)

Thioalkalimicrobium cyclicum (DQ272528)

68

96

95

95

95

59

70

77

99

94

83

80

8886

6056

86

96

99

66

53

98

62

67

83

70

65

75

50

54

94

79

53

97

53

68

93

60

51

73

57

5768

0.02

-Proteobacteria, ChromatialesT

hio

tric

hale

s

cbbL

1

-P

rote

obact

eria,

cbbL

2

-Proteobacteria, Thiotrichales

Figure 5a: Phylogenetic tree (amino acid-based) of cbbL gene sequences retrieved in this study and closest relatives. The scale bar indicates 2% sequence divergence.

Page 129: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

121

Results

Station 2 March (seq14); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq26); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq3); 0-1 cm Lake Grevelingen sediment

uncultured bacterium; epibiota hydrothermal vent shrimp (CBJ18443) Station 2 March (seq21); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq4); 0-1 cm Lake Grevelingen sediment

Station 3 March (seq10); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq7); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq11); 0-1 cm Lake Grevelingen sedimentuncultured bacterium; epibiota hydrothermal vent shrimp (CBJ18444)

Station 2 March (seq9); 0-1 cm Lake Grevelingen sedimentuncultured bacterium; hydrothermal vent (CBH30829)

uncultured bacterium; epibiota hydrothermal vent shrimp (CBJ19314)uncultured bacterium;epibiota hydrothermal vent shrimp (CBJ19314)

Sulfuricurvum kujiense (YP004059532)Station 2 August (seq9); 0-1 cm Lake Grevelingen sediment

Sulfurimonas denitrificans (WP011372203)Thiovulum sp. ES (WP008352685)

Station 1 March (seq9); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq3); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq13); 0-1 cm Lake Grevelingen sedimentuncultured bacterium; hydrothermal vent (CBH30846)

Arcobacter sulfidicus (AAQ76339)Candidatusuncultured bacterium; hydrothermal vent (CBH30845)Station 1 March (seq3); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq25); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq11); 0-1 cm Lake Grevelingen sedimentuncultured bacterium; deep sea hydrothermal vent (CBH30818)uncultured episymbiont of Alvinella pompejana (AAQ76330) Station 2 March (seq28); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq17); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq3); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq8); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq5); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq5); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq12); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq10); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq14); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq6); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq21); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq24); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq3); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq13); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq7); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq16); 0-1 cm Lake Grevelingen sediment

Station 3 March (seq15); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq4); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq25); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq8); 0-1 cm Lake Grevelingen sediment

Sulfurimonas gotlandica (WP008336361)Station 2 August (seq2); 0-1 cm Lake Grevelingen sediment

Station 1 August (seq18); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq15); 0-1 cm Lake Grevelingen sediment

Station 2 August (seq6); 0-1 cm Lake Grevelingen sedimentStation 3 August (seq11); 0-1 cm Lake Grevelingen sediment1

Station 1 August (seq22); 0-1 cm Lake Grevelingen sedimentSulfurimonas autotrophica (WP013327378)

Station 3 August (seq7); 0-1 cm Lake Grevelingen sedimentuncultured bacterium; deep sea hydrothermal vent (AAS01105)

sp. AR (WP008245184)SulfurovumStation 3 March (seq7); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq15); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq16); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq27); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq9); 0-1 cm Lake Grevelingen sediment

Station 2 March (seq1); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq8); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq16); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq3); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq2); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq29); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq22); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq4); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq20); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq17); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq12); 0-1 cm Lake Grevelingen sediment

Station 1 March (seq15); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq2); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq14); 0-1 cm Lake Grevelingen sedimentStation 1 March (seq19); 0-1 cm Lake Grevelingen sedimentStation 3 March (seq16); 0-1 cm Lake Grevelingen sediment

Station 3 August (seq15); 0-1 cm Lake Grevelingen sedimentuncultured episymbiont of Alvinella pompejana (AAQ76323)uncultured episymbiont of Alvinella pompejana (AAQ76331)uncultured episymbiont of Alvinella pompejana (AAQ76305)

Nautilia profundicola (WP012663645)Caminibacter mediatlanticus (WP007475383)

uncultured episymbiont of Alvinella pompejana (AAQ76296)uncultured episymbiont of Alvinella pompejana (AAQ76291)

Station 1 March (seq8); 0-1 cm Lake Grevelingen sedimentStation 2 March (seq5); 0-1 cm Lake Grevelingen sediment

Nitratifractor salsuginis (WP013553540)Station 3 March (seq22); 0-1 cm Lake Grevelingen sediment

sp. TC5-1 (WP022847029)Desulfurobacteriumuncultured sp. (ABW04211)Sulfurihydrogenibium

Persephonella marina (WP012676068)Desulfurobacterium thermolithotrophum (WP013638724)Desulfurobacterium crinifex (ABI50077)Thermovibrio ammonificans (WP013537775)Thermovibrio ruber (ABI50082)

Station 2 August (seq23); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq24); 0-1 cm Lake Grevelingen sediment

Station 2 August (seq22); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq25); 0-1 cm Lake Grevelingen sediment

Station 1 August (seq13); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq14); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq9); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq20); 0-1 cm Lake Grevelingen sediment

Station 1 August (seq21); 0-1 cm Lake Grevelingen sedimentStation 1 August (seq8); 0-1 cm Lake Grevelingen sediment

Thermoplasmatales archaeon (WP008441970)Station 3 August (seq8); 0-1 cm Lake Grevelingen sediment

Methanosaeta harundinacea (WP014587073)Station 3 March (seq20); 0-1 cm Lake Grevelingen sediment

Station 2 August (seq21); 0-1 cm Lake Grevelingen sedimentStation 2 August (seq18); 0-1 cm Lake Grevelingen sediment

Station 2 August (seq20); 0-1 cm Lake Grevelingen sediment

100

99

84

100

100

66

89

60

65

100

98

94

99

66

89

51

52

59

63

79

98

54

64

88

51

80

6780

94

50

86

50

97

82

52

65

71

52

73

57

58

62

55

5765

67

0.05

10x Station 1 March 16x Station 2 March9x Station 3 March

8x Station 1 August7x Station 2 August8x Station 3 August

Phylum Aquificea

Ph

ylu

m E

ury

arc

haeo

taE

ps

ilo

np

rote

ob

ac

teri

a

clust

er

1cl

ust

er

2cl

ust

er

3cl

ust

er

4cl

ust

er

5cl

ust

er

6

Figure 5b: Phylogenetic tree (amino acid-based) of aclB gene sequences retrieved in this study and closest relatives. The scale bar indicates 5% sequence divergence.

Page 130: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

122

lum, Arcobacter, and macrofaunal endosymbionts. As observed for the cbbL gene, no clustering of sequences was observed according to station or season (Fig. 5b).

5.3.6. Quantification of cable bacteria and filamentous Beggiatoaceae We performed a detailed microscopy-based quantification of the biovolume of sulfur oxidizing cable bacteria and filamentous Beggiatoaceae because both groups have been reported to govern the sediment geochemistry and sulfur cycling in sediments of Lake Grevelingen (Seitaj et al., 2015), and thus are likely to influence the chemolithoautotrophic community. Biovolume data of both filamentous bacteria for S1 have been reported before (Seitaj et al., 2015) whereas data for S2 and S3 are novel results from the 2012 campaign. In March, high biovolumes of cable bacteria were detected in S1 and S3 (Table 1) with filaments present throughout the suboxic zone until a maximum depth of 4 cm (Fig. 1c). At the same time, cable bacteria were absent in S2 (Fig. 1c). In August, cable bacteria were only detected between 1 and 2 cm deep at the deepest station (S1), albeit at abundances that were close to the detection limit of the technique. Beggiatoaceae were found in all three stations in March, although the biovolume at S2 was one order of magni-tude higher than in the other two stations (Table 1). In station S2, Beggiatoaceae were uniformly distributed up to the sulfide appearance depth (Fig. 1c). In August, Beggiatoaceae were no longer detectable in stations S1 and S3 (Table 1), while at S2, filaments were no longer found in deeper sediment, but formed a thick mat at the sediment-water interface (Fig. 1c).

5.4. Discussion

5.4.1. Temporal shifts of chemolithoautotrophy and the associated bacterial assemblages Together, our geochemical and microbiological characterization of the sediments of Lake Grev-elingen indicates that the availability of electron acceptors (O2, NO3

–) constitutes the main en-vironmental factor controlling the activity of the chemolithoautotrophic bacterial community. In the deeper basins of Lake Grevelingen (below water depth of 15 m), the electron acceptor availability changes on a seasonal basis due to the establishment of summer hypoxia (Seitaj et al., 2015; Sulu-Gambari et al., 2016). In March, when high oxygen levels are found in the bottom waters, the chemolithoautotrophy rates were substantially higher than in August when oxygen in the bottom water was depleted to low levels (Fig. 1b; Table 1). Moreover in August, the chemo-lithoautotrophy rates showed a clear decrease with water depth (from S3 to S1) in line with the decrease of the bottom water oxygen concentration over depth.

The 16S rRNA gene amplicon sequencing analysis in our study reveals that the response of the chemolithoautotrophic bacterial community in the surface sediments of Lake Grevelingen is associated with the seasonal changes in bottom water oxygen (Fig. 2). Generally, Lake Grevelin-gen surface sediments harbor a distinct microbial community compared to other surface sedi-ments such as coastal marine (North Sea), estuarine and Black Sea sediments (Jessen et al., 2016; Ye et al., 2016; Dyksma et al., 2016) studied by 16S rRNA gene amplicon sequencing analysis.

In March, with oxic bottom waters, the microbial community in the top centimeter of the sediment at all three stations was characterized by high abundances of Epsilon- and Gammapro-teobacteria, which are known chemolithoautotrophic sulfur-oxidizers (e.g. capable of oxidizing sulfide, thiosulfate, elemental sulfur and polythionates) using oxygen or nitrate as electron ac-ceptor (Campbell et al., 2006; Sorokin et al., 2001; Takai et al., 2006; Labrenz et al., 2013). In

Page 131: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

123

Results/Discussion

August, the lack of electron acceptors (O2, NO3–) in the bottom water was accompanied by a

decrease in the relative abundance of these Gammaproteobacteria and Epsilonproteobacteria. At the same time, the numbers of reads related to the Desulfobacteraceae family increased in the top centimeter of the sediment, and a shift in the microbial community structure towards sul-fate reducing bacteria related to the genera Desulfococcus and Desulfosarcina was evident. In coastal sediments, these genera are characteristic in deeper sediment layers experiencing anaerobic min-eralization (Llobet-Brossa et al., 2002; Miyatake, 2011; Muyzer and Stams, 2008), and thus, they are not unexpected in surface sediments during strongly hypoxic (S2) and anoxic (S1) conditions encountered in August.

Shifts in PLFA patterns are in agreement with the temporal difference in the bacterial com-munity (Fig. 3a), with more PLFAs found in Gammaproteobacteria and Epsilonproteobacteria in March (i.e., 16:1ω7c and 18:1 ω7c; Vestal and White, 1989) as opposed to more PLFAs found in sulfate reducing bacteria in the Deltaproteobacteria in August (i.e., ai15:0, i17:1ω7c, 10Me16:0; Boschker and Middelburg, 2002; Taylor and Parkes, 1983, Edlund et al., 1985). However, sedi-ments below the OPD in March have PLFA patterns that are more similar to those of surface sediments in August, in agreement with the observation that anaerobic metabolism such as sul-fate reduction prevail in deeper anoxic sediments also in March.

5.4.2. Spring: Activity and diversity of chemolithoautotrophic bacteriaAlthough the availability of soluble electron acceptors (O2, NO3

–) in the bottom water is import-ant, our data show that it cannot be the only structuring factor of the chemolithoautotrophic communities. In March 2012, all the three stations examined experienced similar bottom water O2 and NO3

– concentrations (Table 1), but still substantial differences were observed in the composition of the chemolithoautotrophic communities as determined by PLFA-SIP (Fig. 3b), as well as in the depth distribution of the chemolithoautotrophy rates in the sediment (Fig. 1b). We attribute these differences to the presence of specific sulfur oxidation mechanisms that are active in the sediments of Lake Grevelingen (Seitaj et al., 2015). In March 2012, based on FISH counts and microsensor profiles two separate sulfur oxidation regimes were active at the sites investigated: sites S1 and S3 were impacted by electrogenic sulfur oxidation by cable bacteria, while site S2 was dominated by sulfur oxidation via nitrate-storing, filamentous Beggiatoaceae. Each of these two regimes is characterized by a specific sulfur-oxidizing microbial community and a particular depth distribution of the chemolithoautotrophy. We now discuss these two re-gimes separately in more detail.

Electrogenic sulfur oxidationIn March, stations S1 and S3 (Fig. 1a) showed the geochemical fingerprint of electrogenic

sulfur oxidation (e-SOx) consisting of a centimeter-wide suboxic zone that is characterized by acidic pore waters (pH < 7) (Nielsen et al., 2010; Meysman et al., 2015). Electrogenic sulfur oxidation is attributed to the metabolic activity of cable bacteria (Pfeffer et al., 2012), which are long filamentous bacteria related to the sulfate reducing genus Desulfobulbus that extend centi-meters deep into the sediment (Schauer et al., 2014). The observed depth-distribution of cable bacteria at S1 and S3 (Fig. 1c) was congruent with the geochemical fingerprint of e-SOx. Cable bacteria couple the oxidation of sulfide in deeper layers to the reduction of oxygen near the sediment-water interface, by channeling electron along their longitudinal axis (long-distance elec-tron transport). Note that stations S1 and S3 also contained some Beggiatoaceae. However, they

Page 132: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

124

attained low biovolumes and were only found at certain depths, suggesting that they did not play a significant role in sulfur oxidation.

When e-SOx was present in the sediment (sites S1 and S3 in March), the chemolithoautotro-phic activity penetrated deeply into the sediment and was evenly distributed throughout the su-boxic zone (Fig. 1b). These field observations confirm previous laboratory incubations in which cable bacteria were induced under oxic conditions in homogenized sediments, and a highly simi-lar depth pattern of deep chemolithoautotrophy was noted (Vasquez-Cardenas et al., 2015). This deep dark CO2 fixation is unexpected in two ways. Firstly, cable bacteria are likely not responsible for the deep CO2 fixation although they do perform sulfur oxidation, as cable bacteria from Lake Grevelingen have been shown to incorporate organic carbon rather than inorganic carbon (Vasquez-Cardenas et al., 2015). Secondly, chemolithoautotrophy is generally dependent on re-oxidation reactions, but there is no transport of oxygen or nitrate to centimeters depth in these cohesive sediments, and so the question is: how can chemolithoautotrophic reoxidation occur in the absence of suitable electron acceptors?

To reconcile these observations, it was proposed that in incubated electro-active sediments from Lake Grevelingen, heterotrophic cable bacteria can form a sulfur-oxidizing consortium with chemolithoautotrophic Gamma- and Epsilonproteobacteria throughout the suboxic zone

Figure 6: Schematic representation of the main sulfur oxidation mechanisms by chemoautotrophic bac-teria under oxic conditions in spring (left) and hypoxic conditions in summer (right) in marine Lake Grevelingen (single-celled and filamentous sulfur-oxidizing bacteria are shown as follows, Epsilonproteo-bacteria (yellow), Gammaproteobacteria (green), Deltaproteobacteria (red), grey arrows represent possible reoxidation processes by chemoautotrophic bacteria and the color of the reaction indicates the bacterial group involved, blue arrow: trajectory of filamentous Beggiatoaceae (thick green lines) from the surface to the sulfide horizon and back up to the surface, black arrow: long-distance electron transport by cable bacteria (red broken lines).

Page 133: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

125

Discussion

(Vasquez-Cardenas et al., 2015). The results obtained in our study provide various lines of evi-dence that support the existence of such a consortium in intact sediment cores. The PLFA-SIP patterns in S1 and S3 (Fig. 3b) showed major 13C-incorporation in PLFA which are present in sulfur-oxidizing Gamma- and Epsilonproteobacteria (Takai et al., 2006; Inagaki et al., 2003; Donachie et al., 2005; Zhang et al., 2005), and this occurred throughout the top 5 cm of the sediment corresponding to the zone of e-SOx activity. In addition, the depth profiles of genes involved in dark CO2 fixation (Fig. 4) revealed chemolithoautotrophic Gammaproteobacteria using the CBB cycle as well as Epsilonproteobacteria using the rTCA cycle, which confirms the potential dark CO2 fixation by both bacterial groups deep in the sediment. The higher abun-dance of cbbL genes in S1 compared to S3 indicates a greater chemolithoautotrophic potential of Gammaproteobacteria, which may explain the two-fold higher total chemolithoautotrophy rate encountered in S1. Moreover, the peak in aclB gene abundance found in deeper sediment suggests that Epsilonproteobacteria could play an important role in sulfur oxidation in the deep-er suboxic zone in both stations. Clearly, a consortium could sustain the high rates of chemolith-oautotrophy throughout the suboxic zone. It has been speculated that chemolithoautotrophs use the cable bacteria as an electron sink in the absence of an electron acceptor like O2 or NO3

– in centimeter deep sediments (Vasquez-Cardenas et al., 2015). However the question still remains as to how the Gamma- and Epsilonproteobacteria are metabolically linked to the cable bacteria (Fig. 6).

In March, stations S1 and S3 also showed a higher contribution of fatty acids occurring in the sulfate reducing Deltaproteobacteria (Boschker et al., 2014; Taylor and Parkes, 1983; Webster et al., 2006) compared to the PLFA-SIP profiles in station S2 (Fig. 3b). Deltaproteobacteria such as Desulfobacterium autotrophicum and Desulfocapsa sp., as identified by 16S rRNA gene sequencing, are known to grow as chemolithoautotrophs by performing H2 oxidation or S0-disproportion-ation (Finster et al., 1998, Böttcher et al., 2005; Fig. 6), and are important contributors to the chemolithoautotrophic activity in coastal sediments (Boschker et al., 2014; Thomsen and Kris-tensen, 1997). However, a further identification of the chemolithoautotrophic Deltaproteobac-teria through the functional genes related to carbon fixation pathways was not performed in this study. Although it is known that autotrophic Deltaproteobacteria mainly use the reverse TCA cycle or the reductive acetyl-CoA pathway (Hügler and Sievert, 2011), further development of the functional gene approach, by designing specific primers is necessary to determine and clarify the diversity of Deltaproteobacteria involved in the chemolithoautotrophic activity.

Overall, we hypothesize that such a diverse assemblage of chemolithoautotrophic bacteria in the presence of cable bacteria is indicative of a complex niche partitioning between these sulfur-oxidizers (Gamma-, Epsilon-, and Deltaproteobacteria). In sulfidic marine sediments found in tidal and deep sea habitats, complex S0 niche partitioning have been proposed where uncultured sulfur-oxidizing Gammaproteobacteria mainly thrive on free sulfide, the epsilonproteobacterial Sulfurimonas/Sulfurovum group oxidizes elemental sulfur, and members of the deltaproteobacterial Desulfobulbaceae family may perform S0-disproportionation (Pjevac et al., 2014).

Nitrate-storing filamentous Beggiatoaceae Nitrate-storing filamentous Beggiatoaceae glide through the sediment transporting their electron acceptor (intracellular vacuoles filled with high concentrations of NO3

–) into deeper sediment

Page 134: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

126

and electron donor (intracellular granules of elemental sulfur) up to the surface, and in doing so, they oxidize free sulfide to sulfate in a two-step process that creates a wide suboxic zone (Seitaj et al., 2015; Mußmann et al., 2003; Sayama et al., 2005; Fig. 6). In March, the microsensor depth profiles at S2 (Fig. 1a) revealed a cm-thick suboxic zone with a subsurface pH minimum at the OPD followed by a pH maximum at SAD, which is the characteristic geochemical fingerprint of sulfur oxidation by nitrate-storing Beggiatoaceae (Seitaj et al., 2015; Sayama et al., 2005). At the same time, microscopy revealed high biovolumes of Beggiatoaceae that were uniformly distrib-uted throughout the suboxic zone (Fig. 1c), and more than half the filaments found were thicker than 15 μm indicating potential nitrate storage. Together these results corroborate the indica-tions of the dominant sulfur-oxidation mechanism suggested by the geochemical fingerprint.

The chemolithoautotrophy depth profile at S2 in March (Fig. 1b) recorded higher activities in the top 1 cm and chemolithoautotrophic activity penetrated down only to 2 cm (at the sulfide appearance depth). The similar depth distribution of Beggiatoaceae and chemolithoautotrophic activity suggests that dark CO2 fixation was primarily carried out by the nitrate-storing Beggia-toaceae. Beggiatoaceae can indeed grow as obligate or facultative chemolithoautotrophs de-pending on the strain (Hagen and Nelson DC, 1996). The PLFA-SIP analysis further supported chemolithoautotrophy by Beggiatoaceae as the PLFA patterns obtained at S2 resembled those of Beggiatoa mats encountered in sediments associated with gas hydrates (Zhang et al., 2005).

However, the CO2 fixation by Beggiatoaceae could be complemented by the activity of other chemolithoautotrophs. Beggiatoaceae have previously been reported to co-occur with chemo-lithoautotrophic nitrate-reducing and sulfur-oxidizing Epsilonproteobacteria (Sulfurovum and Sul-furimonas) in the deep sea Guyamas basin (Bowles et al., 2012). Interestingly, station S2 in March showed the highest abundance of aclB gene copies of all stations and seasons (Fig. 4b), and in addition, had the highest relative percentage of 16S rRNA gene read sequences assigned to the Epsilonproteobacteria (Fig. 2). Therefore, the dark CO2 fixation at station S2 is likely caused by both the motile, vacuolated Beggiatoaceae as well as the sulfur-oxidizing Epsilonproteobacteria. Yet, as was the case of the cable bacteria above, the metabolic link between Beggiatoaceae and the Epsilonproteobacteria remains currently unknown. However, it seems possible that internal-ly stored NO3

– is released into the sediment once Beggiatoa filaments lyse, allowing sulfur-oxida-tion with NO3

– by Epsilonproteobacteria.

5.4.3. Summer: Activity and diversity of chemolithoautotrophic bacteria In August, when the O2 levels in the bottom water decreased substantially, a third geochemi-cal regime was observed at all stations, which was different from the regimes associated with cable bacteria or nitrate-storing Beggiatoaceae. The microsensor profiling revealed an upward diffusive transport of free sulfide to the top millimeters of the sediment, which was produced in deeper sediment horizons through sulfate reduction (Fig. 1a). The uniform decrease in pH with depth is consistent with sediments dominated by sulfate reduction (Meysman et al., 2015). As noted above, the chemolithoautotrophy rates strongly decreased compared to March and were restricted to the very top of the sediment. Chemolithoautotrophic activity showed a close relation with the bottom water oxygen concentration. The number of gene copies of the two carbon fixation pathways investigated (CBB and rTCA) also decreased under hypoxic conditions (Fig. 4), suggesting a strong dependence of subsurface chemolithoautotrophy on the availability of oxygen in bottom waters.

Page 135: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

127

Discussion/Conclusion

The highest chemolithoautotrophy rate was recorded at the shallower station (S3), which was the only station in August where O2 was found to diffusive into the sediment (Table 1), thus supporting aerobic sulfur oxidation at a shallow oxic-sulfidic interface in the sediment (Meysman et al., 2015). Several studies have shown that chemolithoautotrophy ceases in the absence of oxygen in the overlying water column (Vasquez-Cardenas et al., 2015; Miyatake, 2011; Thomsen and Kristensen, 1997). Such full anoxia occurred in the deepest station (S1), but still limited chemolithoautotrophic activity was recorded in the top layer of the sediment. A possible ex-planation is that some residual aerobic sulfur oxidizing bacteria were supported by the very low oxygen levels. Alternatively, cable bacteria activity at S1 in spring leads to a buildup of iron (hydr)oxides (FeOOH) in the top of the sediments, which prevents sulfide diffusion to the bottom water during summer anoxia (3H2S + 2FeOOH → 2FeS + S0 + 4H2O) (Seitaj et al., 2015). The elemental sulfur formed in this reaction may support chemolithoautotrophic sulfur dispropor-tionating bacteria (Finster et al., 1998). At station S2, Beggiatoaceae were found forming a mat at the sediment surface in S2 (Fig. 1b) and chemolithoautotrophy rates were limited to this mat (Fig. 6). Beggiatoaceae can survive hypoxic periods by using stored nitrate as electron acceptor (Schulz and Jørgensen, 2001), which is thought to be a competitive advantage that leads to their proliferation in autumn in S1 (Seitaj et al., 2015). Such survival strategies may be used to produce energy for maintenance rather than growth under hypoxic conditions which would in combina-tion with the low oxygen concentrations explain the low chemolithoautotrophy to biovolume ratio observed in S2 in August compared to March.

5.5. Conclusion Coastal sediments harbor a great potential for chemolithoautotrophic activity given the high anaerobic mineralization, which produce a large pool of reduced sulfur in these organic rich sediments. In this study, two environmental factors were identified to regulate the chemolith-oautotrophic activity in coastal sediments: seasonal hypoxia and the dominant sulfur oxidation mechanism. In sediments where oxygen, the main electron acceptor for chemolithoautotrophs, is depleted because of seasonal hypoxia, chemolithoautotrophy was strongly inhibited. In addi-tion, it is clear that the different sulfur oxidation mechanism (e.g. canonical sulfur oxidation, elec-trogenic sulfur oxidation, or sulfur oxidation mediated by vacuolated Beggiatoaceae) observed in the sediment also determine the magnitude and depth distribution of the dark CO2 fixation, as well as the chemolithoautotrophic bacterial community structure. The seasonal variations in electron acceptors and potentially reduced sulfur species suggest complex niche partitioning in the sediment by the sulfur-oxidizing bacterial community. An in depth study on the availability of different sulfur species in the sediment could shed light on the sulfur preferences by the dif-ferent bacterial groups. Likewise, the potential mechanism used to metabolically link filamentous sulfur-oxidizers and chemolithoautotrophic bacteria remains presently unresolved and requires further study. These tight metabolic relationships may ultimately regulate the cycling of sulfur, carbon and even nitrogen in coastal sediments, including (but not limited to) sediments affected by seasonal hypoxia.

Acknowledgments

We thankfully acknowledge the crew of the R/V Luctor (Peter Coomans and Marcel Kristalijn) and Pieter van Rijswijk for their help in the field during sediment collection, Marcel van der

Page 136: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

128

Meer and Sandra Heinzelmann for support onboard and assistance with incubations, Peter van Breugel and Marco Houtekamer for their assistance with stable isotope analysis, Elda Panoto for technical support and Alexandra Vasquez Cardenas for designing Figure 6. This work was finan-cially supported by several grants from Darwin Centre for Biogeosciences to HTSB (grant no: 3061), LV (grant no: 3062) and FJRM (grant no: 3092), FJRM received funding from the Euro-pean Research Council under the European Union’s Seventh Framework Program (FP7/2007–2013) via ERC grant agreement n°[306933], and RS received support from Danish Council for Independent Research-Natural Sciences.

Page 137: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

129

Supplementary Material

Supplementary information

Experimental procedures

Aeration of 13C- incubations

To maintain the low oxygen levels found in August at S2 and S3, 30 ml of in situ overlying water was removed and cores were closed with rubber stoppers. Headspace was then flushed with N2 through a needle inserted between the core liner and the rubber stopper without disturbing the water surface. After several minutes, N2 flushing was stopped and 10 and 35% of the headspace was replaced with air using a syringe for S2 and S3, respectively. No air was added to S1 cores. All cores were gently mixed (0.2 rpm) on a shacking plate to homogenize oxygen concentration in headspace with those in overlying water. At the end of the experiment oxygen saturation in overlying water were verified with oxygen optodes (PreSens Fibox 3 LCD) obtaining anoxic conditions for S1 (0−4% air saturation), hypoxic for S2 (20−26%) and S3 (35−80%).

PLFA nomenclature The shorthand nomenclature used for phospholipid derived fatty acids is as follows. The num-ber before the colon indicates the number of carbon atoms, while the number after represents the number of double carbon bonds in the fatty acids chain. The position of the initial double bond is then indicated by the last number after the number of carbons from the methyl end (ω). The double bond geometry is designated by cis (c) or trans (t). Methyl branching is described as being in the second carbon iso (i), third carbon anteiso (a) or a number followed by Me is used to indicate the position relative to the carboxyl end (e.g. 10Me16:0). Further details can be found in the literature (Vestal and White, 1989).

PCR 16S rRNA gene amplicon library preparation and analysis PCR reactions were performed with the universal (Bacteria and Archaea) primers S-D-Arch-0519-a-S-15 (5’-CAG CMG CCG CGG TAA-3’) and S-D-Bact-0785-a-A-21 (5’-GAC TAC HVG GGT ATC TAA TCC-3’) (Klindworth et al., 2013) adapted for pyrosequencing by the ad-dition of sequencing adapters and multiplex identifier (MID) sequences. To minimize bias three independent PCR reactions were performed containing: 16.3 µL H2O, 6 µL HF Phusion buffer, 2.4 µL dNTP (25 mM), 1.5 µL forward and reverse primer (10 µM; each containing an unique MID tail), 0.5 µL Phusion Taq and 2 µL DNA (6 ng/µL). The PCR conditions were following: 98ºC, 30 s; 25 × [98ºC, 10 s; 53 ºC, 20 s; 72ºC, 30 s]; 72 ºC, 7 min and 4ºC, 5 min.

The PCR products were loaded on a 1% agarose gel and stained with SYBR® Safe (Life technologies, Netherlands). Bands were excised with a sterile scalpel and purified with Qiaquick Gel Extraction Kit (QIAGEN, Valencia, CA) following the manufacturer’s instructions. PCR purified products were quantified with Quant-iTTM PicoGreen® dsDNA Assay Kit (Life Tech-nologies, Netherlands). Equimolar concentrations of the barcoded PCR products were pooled and sequenced on GS FLX Titanium platform (454 Life Sciences) by Macrogen Inc. Korea.

Samples were analyzed using the QIIME pipeline (Caporaso et al., 2010). Raw sequences were demultiplexed and then quality-filtered with a minimum quality score of 25, length between 250−350, and allowing maximum two errors in the barcode sequence. Sequences were then clustered into operational taxonomic units (OTUs, 97% similarity) with UCLUST (Edgar, 2010). Reads were aligned to the Greengenes Core reference alignment (DeSantis et al., 2006) using

Page 138: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

130

the PyNAST algorithm (Caporaso et al., 2010). Taxonomy was assigned based on the Green-genes taxonomy and a Greengenes reference database (version 12_10) (McDonald et al., 2012; Werner et al., 2012). Representative OTU sequences assigned to the specific taxonomic groups were extracted through classify.seqs and get.lineage in Mothur (Schloss et al., 2009) by using the greengenes reference and taxonomy files.

In order to determine a more accurate taxonomic classification of the bacterial groups with high percentage of reads and known to contain members with chemolithoautotrophic metabo-lism, sequence reads of the order Chromatiales and Thiotrichales (Fig. S1, S2), reads of the order Desulfarculales (Fig. S3), the family of Desulfobulbaceae (Fig. S4a) and the genus Desulfubulbus (Fig. S4b), reads of the family Desulfobacteraceae (Fig. S5a–c), and additionally reads of the or-der Campylobacterales (Fig. S6a–d) were extracted from the dataset and added to a phylogenetic tree as described in the supplementary experimental procedures.

Quantification of cable bacteria and Beggiatoaceae Microscopic identification of cable bacteria was achieved by fluorescence in situ hybridization (FISH), using a Desulfobulbaceae-specific oligonucleotide probe (DSB706; 5′-ACC CGT ATT CCT CCC GAT-3′), according to (Schauer et al., 2014). Cable bacteria biovolume per unit of sediment volume (mm3 cm−3) was calculated based on measured filament length and diameter. The areal biovolume of cable bacteria (mm3 cm−2) was obtained by depth integration over all sediment layers analyzed.

The minimum limits of quantification via FISH for single cells were 1.5 × 106 cells cm−3, taken as the FISH count with the negative control probe NON338; a minimum of 1000 DAPI (4,6-diamidino-2-phenylindole) stained cells was evaluated for this count. The FISH detection limit for cable bacteria was lower than for single cells, and was calculated to be 10 cm filament cm−3 (corresponding to <1 filament in 0.1 mL of sediment).

Filamentous Beggiatoaceae were identified via inverted light microscopy (Olympus IM) within 24 h of sediment retrieval. Intact sediment cores were sectioned at 5 mm intervals over the top 4 cm from which subsamples (20−30 mg) were used to count living Beggiatoaceae (Seitaj et al., 2015). The biovolume was determined by measuring length and width of all filaments found in the subsample, following the counting procedure described in (Jørgensen et al., 2010).

Statistical analysis All statistical analyses were performed using the CRAN: stats package in the open source soft-ware R. A two-way ANOVA (aov) was used to test the effect of station, season, and sediment depth on bacterial biomass, chemolithoautotrophic activity, and gene abundances. PLFA con-centrations and 13C-incorporation values were expressed as a fraction of the total bacterial bio-mass and 13C-incorporation (respectively) per sediment sample. These relative PLFA values were first log-transformed (log (x+1)) and subsequently analyzed with Principal Component Analysis (PCA: prcomp) to distinguish different bacterial and chemolithoautotrophic communities in the sediment.

Page 139: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

131

Supplementary Material

Tabl

e S1

: Prim

er p

airs

des

crib

ed in

the

text

, PC

R co

nditi

ons a

nd a

mpl

icon

size

use

d in

this

stud

y.

Ass

ayTa

rget

Prim

er p

air

Tm

[°C

]A

mpl

icon

si

ze [b

]R

efer

ence

qPC

R +

PC

R/cl

onin

gcb

bL g

ene

K2f

(5’-A

CC

AYC

AA

GC

CSA

AG

CTS

GG

-3’)

qPC

R 57

PCR

6249

2-49

5N

igro

and

K

ing,

200

7

V2r

(5’-G

CC

TTC

SAG

CTT

GC

CSA

CC

RC-3

’)

qPC

R +

PC

R/cl

onin

gac

lB g

ene

aclB

F (5

’-TG

GA

CH

ATG

GTD

GC

HG

GK

GG

T-3’

)qP

CR

54PC

R 53

312

Cam

pbel

l et

al.,

200

3ac

lB R

(5’-A

TART

TKG

GBC

CA

CC

KC

KTC

-3’)

PCR

cond

ition

s: 95

°C 5

min

; 40

× [9

5°C

1 m

in, T

m 4

0 s,

72°C

1 m

in];

72°C

5 m

in. q

PCR

cond

ition

s: 95

°C 4

min

; 40

× [9

5°C

30

s, Tm

40

s, 72

°C

30 s]

; 80°

C 2

5 s.

Page 140: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

132

AB681900, Granulosicoccus antarcticusEF495228, Granulosicoccus antarcticus, seawater

KC160719, Granulosicoccus sp. SS10.26, Antarctic sea sedimentFJ535355, Granulosicoccus coccoides, leaves of sea grass in Troitza Bay, Gulf of Peter the Great, Sea of Japan

GU451354, uncultured bacterium, macroalgal surfaceJF344155, uncultured gamma proteobacterium, petroleum spot over marine sediments from figueiras beach

ARB_EFBD0A28, denovo1856 GI2.1_972JX559163, uncultured bacterium, antarctic aerosol

FJ905714, uncultured bacterium, iron oxide sediments, Volcano 1, Tonga ArcGQ119472, uncultured proteobacterium

ARB_E6BF6C65, denovo119 GI1.1_2739Hq153883, uncultured bacterium, shallow hydrothermal vent 189.1 m

EU246808, uncultured gamma proteobacterium, Calcinus obscurus abdominal floraARB_5B178E07, denovo1043 GI1.1_895

JF344255, uncultured gamma proteobacterium, petroleum spot over marine sediments from figueiras beachEU050785, uncultured gamma proteobacterium, sediment from the Kings Bay, Svalbard, Arctic

GU234740, uncultured marine bacterium, Antarctic sea water collected from 5mARB_62B26A19, denovo3300 GII3.1_6302

AB099937, uncultured bacterium, inactive deep−sea hydrothermal vent chimneysDQ269113, uncultured gamma proteobacterium, surface of marine macro−alga

GQ274231, uncultured gamma proteobacterium, heat exchange system using biocides to inhibit biofilm formationARB_BB7219C4, denovo505 GII2.1_2624ARB_97A6CE1, denovo309 GI2.1_1401

EU491486, uncultured bacterium, seafloor lavas from Hawai’i South Point X4JN606989, uncultured bacterium, white microbial mat from lava tube wall

EU491392, uncultured bacterium, seafloor lavas from Hawai’i South Point X3JX226828, uncultured bacterium, nodule collected from station WS0904 in the Clarion−Clipperton Fracture Zone

EU287407, uncultured bacterium, arctic surface sedimentKC861163, uncultured bacteriumEU652530, uncultured bacterium, Yellow Sea sediment

ARB_6609E7D9, denovo1802 GI1.1_2332JF266091, uncultured bacterium, white microbial mat lava tube wall

HQ397072, uncultured gamma proteobacterium, haloalkaline soilJF344181, uncultured gamma proteobacterium, petroleum spot over marine sediments from figueiras beach

AB476263, uncultured bacterium, ventral setae of Shinkaia crosnieriARB_BBCD0639, denovo969 GI1.1_2370

Hq153957, uncultured bacterium, white filamentous microbial mat in the photic zone of a shallow hydrothermal ventEF092205, uncultured gamma proteobacterium

FR744557, uncultured bacterium, nitrate−amended injection water from the Halfdan oil fieldEU491601, uncultured bacterium, seafloor lavas from the East Pacific Rise

HQ379738, gamma proteobacterium S−1(2010)JX391088, uncultured bacterium, marine sediment

JF344304, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachARB_9C00EB59, denovo1294 GII1.1_8935ARB_9C00EB59, denovo2723 GI1.1_3081

JF344362, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachJF344419, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

JF344330, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachJF344428, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

FM242303, uncultured gamma proteobacterium, sedimentAM176842, uncultured bacterium, mangrove sediment

HQ916536, uncultured bacterium, Lei−Gong−Huo mud volcanoAB278144, uncultured bacterium, obtained from Mariana Trough

HF558592, uncultured bacterium, tailing material

Chromatiales_

Granulosico

ccace

ae_Granulosico

ccus

0.05

DQ498828, Thiocapsa imhoffiiJX521203, uncultured bacterium, terrestrial sulfidic spring

AJ223235, Lamprocystis purpureaAJ006063, Lamprocystis roseopersicina

Y12366, Lamprocystis purpureaAb478670, uncultured Chromatiaceae bacterium, submerged biofilm mats wetland

DQ676347, uncultured gamma proteobacterium, suboxic freshwater−pond bacterioplanktonARBC01000182, Lamprocystis purpurea DSM 4197AJ006057, phototrophic bacterium

DQ676296, uncultured gamma proteobacterium, suboxic freshwater−pond bacterioplanktonFN396696, uncultured marine bacterium, Arctic marine surface sediment

ARB_DB9D716A, denovo836 GI3.1_1527Am882544, uncultured gamma proteobacterium, sediment from oil polluted water

ARB_3ED72159, denovo2717 GI2.1_4916ARB_4D38CEFC, denovo3368 GII2.1_10913

ARB_2E607D92, denovo759 GII1.1_1362

Lamprocystis

AJ831752, uncultured bacterium, chemoclineJX521262, uncultured bacterium, terrestrial sulfidic spring

AJ224430, Isochromatium buderi

uncultured 16

ARB_A2F14DB, denovo674 GII3.1_6959Jn479885, uncultured organism, Guerrero Negro Hypersaline Mat

Jn461133, uncultured organism, Guerrero Negro Hypersaline Mat ARB_528E7DEA, denovo59 GI1.1_4285

Jn450999, uncultured organism, Guerrero Negro Hypersaline Mat Jn458696, uncultured organism, Guerrero Negro Hypersaline Mat

Jn459399, uncultured organism, Guerrero Negro Hypersaline Mat DQ330804, uncultured proteobacterium, hypersaline microbial mat: Guerrero Negro

Jn469059, uncultured organism, Guerrero Negro Hypersaline Mat Am882535, uncultured gamma proteobacterium, sediment from oil polluted water

Am882566, uncultured gamma proteobacterium, sediment from oil polluted water DQ330805, uncultured proteobacterium, hypersaline microbial mat: Guerrero Negro

HQ877095, Lamprobacter modestohalophilusX93472, Halochromatium glycolicumX98597, Halochromatium salexigensAm283535, Halochromatium roseum

Haloch

romatium

KF799054, uncultured bacterium, gutARB_C9137E14, denovo4029 GII3.1_8926

KF513141, uncultured Thiohalocapsa sp., pink berry consortia from Sippewissett salt marshAVFP01004244, microbial mat metagenome, pink berry consortia of the Sippewissett salt marsh

Am491592, Thiohalocapsa marina

Thiohalocapsa

JQ738932, uncultured bacterium, Lonar sediment surface rocks

Chromatiales_Chromatiaceae

0.05

JX138882, uncultured bacterium, marine sedimentsJF809744, uncultured bacterium, Medea hypersaline basin, Mediterranean Sea

EU373841, uncultured gamma proteobacterium, canyon and slope sedimentARB_38DFA738, denovo886 GI3.1_648

JX138600, uncultured bacterium, marine sedimentsEU491667, uncultured bacterium, seafloor lavas from the East Pacific Rise

KC471275, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughDQ289918, uncultured gamma proteobacterium, South Atlantic Bight Permeable Shelf Sediment

ARB_7A35AA84, denovo1280 GI1.1_178HE817849, uncultured bacterium, tissue of sponge Vaceletia crypta

ARB_D868D57F, denovo435 GI3.1_117GU119047, uncultured bacterium

FJ205275, uncultured gamma proteobacterium, active hydrothermal field sediments, depth:1922mJQ347472, uncultured bacterium, Da Ya Bay

ARB_C8883AA2, denovo18 GII3.1_3781ARB_120B20A, denovo2400 GII3.1_10467

ARB_B82136A6, denovo1971 GII2.1_2927ARB_639358B4, denovo3330 GII3.1_10833

ARB_B3D72EEA, denovo473 GI2.1_3560JQ062714, uncultured bacterium, sponge tissue

ARB_57D58DCA, denovo51 GI2.1_107ARB_83FE435B, denovo2355 GI3.1_549

FJ937838, uncultured gamma proteobacterium, coastal waters of Lingshui Bay of Hainan Island in the South China SeaFJ937834, uncultured bacterium, coastal waters of Lingshui Bay of Hainan Island in the South China Sea

JF344578, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachARB_66261C8C, denovo2664 GII3.1_1805

EF208706, uncultured bacterium, sandy carbonate sedimentEU652575, uncultured bacterium, Yellow Sea sediment

KF741570, uncultured bacterium, Talbert salt marsh sedimentJF727679, uncultured gamma proteobacterium, petroleum−contaminated saline−alkali soilsJQ347371, uncultured bacterium, Da Ya Bay

ARB_C634BA31, denovo1128 GI3.1_3820ARB_9441BCE4, denovo1555 GII2.1_57GU118309, uncultured bacterium

JX280354, uncultured bacterium, marine spongeDQ513032, uncultured bacterium, ridge flank crustal fluid

FJ535071, uncultured gamma proteobacterium, cave wall biofilmAF317768, unidentified bacterium wb1_P19

EF125399, uncultured bacterium, naked tidal flat sediment near mangroveARB_37CE9830, denovo1458 GI1.1_565

FJ905747, uncultured bacterium, iron oxide sediments, Volcano 1, Tonga ArcFJ205285, uncultured gamma proteobacterium, active hydrothermal field sediments, depth:1922m

JF272102, uncultured bacterium, biofilmsGU061300, uncultured gamma proteobacterium, Yellow Sea intertidal beach seawater at 20 m depth

ARB_78A3E9B6, denovo818 GI1.1_1073JF344260, uncultured gamma proteobacterium, petroleum spot over marine sediments from figueiras beachHQ191094, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden Sea

ARB_8F2AAD3A, denovo1907 GI3.1_2023JQ579854, uncultured gamma proteobacterium, sediments from Figueiras BeachJF344662, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

GQ246358, uncultured bacterium, North Yellow Sea sedimentsDQ867047, uncultured bacterium, natural gas field

M96398, Nitrosococcus oceaniAY690336, Nitrosococcus oceani

AF508987, Nitrosococcus oceaniAF363287, Nitrosococcus oceani

CP000127, Nitrosococcus oceani ATCC 19707CP000127, Nitrosococcus oceani ATCC 19707AF508994, Nitrosococcus oceaniCP002086, Nitrosococcus watsonii C−113CP002086, Nitrosococcus watsonii C−113

CP001798, Nitrosococcus halophilus Nc 4AJ298727, Nitrosococcus oceaniCP001798, Nitrosococcus halophilus Nc 4

AJ298748, Nitrosococcus halophilusJX240761, uncultured proteobacterium, coastal soil

HF558562, uncultured bacterium, tailing material

Chromatia

les_

Chromatia

ceae_Nitroso

cocc

us

AM882518, uncultured gamma proteobacterium, sediment from oil polluted water from petrochemical treatment plantAF170421, sulfur−oxidizing bacterium OBII5

AJ810552, uncultured gamma proteobacterium, superficial sedimentsARB_334A7441, denovo4022 GII3.1_923

ARB_BB6DFFA9, denovo421 GI1.1_601ARB_516EEF6C, denovo342 GII3.1_5432

uncultured

DQ351800, uncultured gamma proteobacterium, marine sedimentsHQ191097, uncultured sediment bacterium, Dangast muddy intertidal sediment; German Wadden Sea

GQ246364, uncultured bacterium, North Yellow Sea sedimentsAF223297, uncultured gamma proteobacterium B2M28ARB_AB5216C3, denovo3247 GII2.1_612

AF223299, uncultured gamma proteobacterium B2M60JQ586305, uncultured bacterium, arctic marine sediment

HQ191062, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaHQ191090, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden Sea

KF624106, uncultured bacterium, anoxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden SeaAY193138, uncultured proteobacterium, marine sedimentDQ394966, uncultured gamma proteobacterium, harbor sediment

ARB_57253D3A, denovo1419 GI1.1_2310ARB_1DF4361C, denovo2934 GII2.1_7000

FJ545494, uncultured bacterium, North Yellow Sea sedimentGU061176, uncultured gamma proteobacterium, Yellow Sea intertidal beach surface seawater

EU491687, uncultured bacterium, seafloor lavas from the East Pacific RiseEU491530, uncultured bacterium, seafloor lavas from the East Pacific Rise

HQ191065, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaFJ712412, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean Sea

AM889168, uncultured gamma proteobacterium, SedimentsFM211768, uncultured gamma proteobacterium, sediment of Bizerte lagoon

HQ191095, uncultured sediment bacterium, Dangast muddy intertidal sediment; German Wadden SeaKF799091, uncultured bacterium, gutHQ190975, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaHQ190997, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden Sea

HF922366, uncultured bacterium, high pressure reactorHF922334, uncultured bacterium, high pressure reactor

KF624415, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaEU491232, uncultured bacterium, seafloor lavas from the Loi’hi Seamount Pisces Peak X2

Gu369920, uncultured epsilon proteobacterium, white microbial film shallow hydrothermal vent ARB_97962E3E, denovo2892 GI3.1_431

DQ351779, uncultured gamma proteobacterium, marine sedimentsGU061213, uncultured gamma proteobacterium, Yellow Sea intertidal beach seawater at 5 m depth

JF344014, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from rodas beachDQ351744, uncultured gamma proteobacterium, marine sediments

ARB_9A36BE2D, denovo2926 GII3.1_4245Gu302450, uncultured bacterium, marine sediments (900 m water depth, Gulf of Mexico

EU491337, uncultured bacterium, seafloor lavas from Hawai’i South Point X3KC470886, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

FJ416116, uncultured bacterium, sediment from station NEC5 in the Northern Bering SeaFN396751, uncultured marine bacterium, Arctic marine surface sediment

DQ351776, uncultured gamma proteobacterium, marine sedimentsJQ580264, uncultured gamma proteobacterium, sediments from Rodas Beach polluted with crude oil

FJ264585, uncultured bacterium, methane seep sedimentFJ358865, uncultured gamma proteobacterium, marine reef sandy sediment

ARB_B77AF6B7, denovo2435 GI1.1_1585JX504392, uncultured bacterium, oolitic sand

JF344659, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachJX193397, uncultured bacterium, intertidal surface sediment

JF344391, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachDQ395021, uncultured gamma proteobacterium, harbor sediment

ARB_781E2AD4, denovo2106 GI2.1_4886ARB_C0B550B5, denovo551 GII1.1_195ARB_C0B550B5, denovo1651 GI1.1_51

HQ203959, uncultured bacterium, seawaterFM211764, uncultured gamma proteobacterium, sediment of Bizerte lagoon

DQ351781, uncultured gamma proteobacterium, marine sedimentsARB_AD855E3B, denovo1086 GII3.1_4940ARB_42D76043, denovo2199 GI2.1_788

ARB_6192704D, denovo1940 GI2.1_2314AB530205, uncultured bacterium, marine sediment

GQ500735, uncultured bacterium, Charon’s Cascade, sandy clastic sedimentsHm598244, uncultured bacterium, surface sediment Xisha Trough, China Sea

ARB_754734E8, denovo507 GI2.1_2211EF999373, uncultured bacterium, Pearl River Estuary sediments at 22cm depth

DQ395015, uncultured gamma proteobacterium, harbor sedimentAM889166, uncultured gamma proteobacterium, Sediments

DQ395051, uncultured gamma proteobacterium, harbor sedimentJN977193, uncultured bacterium, Jiaozhao Bay sediment

JF344480, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachDQ394926, uncultured gamma proteobacterium, harbor sediment

FJ545485, uncultured bacterium, North Yellow Sea sedimentGQ246283, uncultured bacterium, North Yellow Sea sedimentsAB530211, uncultured bacterium, marine sedimentGU230357, uncultured gamma proteobacterium, coastal sediment

JF344616, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachAM889165, uncultured gamma proteobacterium, Sediments

JN977291, uncultured bacterium, Jiaozhao Bay sedimentAM889164, uncultured gamma proteobacterium, Sediments

DQ394995, uncultured gamma proteobacterium, harbor sedimentDQ395062, uncultured gamma proteobacterium, harbor sediment

KF463825, uncultured gamma proteobacterium, Marine coastal sedimentARB_800DD24B, denovo2123 GI1.1_1928GQ246321, uncultured bacterium, North Yellow Sea sediments

GU230358, uncultured gamma proteobacterium, coastal sedimentGU061255, uncultured gamma proteobacterium, Yellow Sea intertidal beach seawater at 5 m depth

GU061303, uncultured gamma proteobacterium, Yellow Sea intertidal beach seawater at 20 m depthJN038904, uncultured gamma proteobacterium, Chongxi wetland soil

FJ786140, uncultured bacterium, prawn farm sedimentARB_BA77D3F2, denovo1553 GII2.1_8765

ARB_E4ED143B, denovo942 GII2.1_8839ARB_136300C8, denovo2859 GI1.1_837

ARB_E7A49636, denovo132 GII2.1_1684ARB_5E12BDAE, denovo3289 GII3.1_10675

ARB_47DF8362, denovo147 GI2.1_4282ARB_429AAAE0, denovo1745 GI2.1_3679Am882524, uncultured gamma proteobacterium, sediment from oil polluted water

FJ416094, uncultured bacterium, sediment from station NEC5 in the Northern Bering SeaARB_2EDF208B, denovo2396 GII3.1_9109

ARB_87D422D8, denovo2963 GII3.1_3051ARB_BB3E52D1, denovo907 GI2.1_2633

ARB_65268B1C, denovo1635 GII2.1_5704AB722093, uncultured bacterium, subseafloor massive sulfide depositFn553610, uncultured sediment bacterium, Logatchev hydrothermal vent depth = 3224 m

ARB_82B8D5B4, denovo910 GI2.1_3635Fn553597, uncultured sediment bacterium, Logatchev hydrothermal vent 3224 m

EU491112, uncultured bacterium, seafloor lavas from the Loi’hi Seamount Pisces Peak X2FJ497654, uncultured gamma proteobacterium, Vailulu’u SeamountAB722101, uncultured bacterium, subseafloor massive sulfide deposit

ARB_F3CD46EB, denovo3941 GII2.1_7119AB294934, uncultured gamma proteobacterium, microbial mat at a shallow submarine hot spring

Fr670375, uncultured gamma proteobacterium, microbial mat from Eiffel Tower edifice in Lucky Strike EU652533, uncultured bacterium, Yellow Sea sediment

JN977309, uncultured bacterium, Jiaozhao Bay sedimentARB_6DE0F30B, denovo2739 GI3.1_126

EU330397, uncultured gamma proteobacterium, mangrove forest mudARB_910196BD, denovo1551 GI3.1_1932

Hq153920, uncultured bacterium, white microbial film shallow hydrothermal ventEU491810, uncultured bacterium, seafloor lavas from the East Pacific Rise

DQ431906, uncultured gamma proteobacterium, Gulf of Mexico sedimentHQ191077, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden Sea

GU061216, uncultured gamma proteobacterium, Yellow Sea intertidal beach seawater at 5 m depthARB_8B4C06BD, denovo3718 GII1.1_5285ARB_8B4C06BD, denovo1901 GI1.1_3043HQ588368, uncultured bacterium, Amsterdam mud volcano sedimentGQ246407, uncultured bacterium, North Yellow Sea sedimentsFJ497635, uncultured bacterium, Vailulu’u Seamount

Fj712561, uncultured bacterium, Kazan Mud VolcanoKc682690, uncultured bacterium, hydrothermal sulfide chimney, Lau Basin

EU491101, uncultured bacterium, seafloor lavas from the Loi’hi Seamount Pisces Peak X2Fj640811, uncultured bacterium, hydrothermal vent chimney Juan de Fuca Ridge 2267m

GQ903353, uncultured bacterium, ODP Hole 896A basalt crust formation fluidEF999348, uncultured bacterium, Pearl River Estuary sediments at 6cm depth

HQ003527, uncultured gamma proteobacterium, Carrizo shallow lake

Chromatia

les_

Ectothiorhodosp

irace

ae

0.05

Figure S1: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Gammaproteobacteria, order: Chromatiales (black bold: sequences retrieved in this study and closest relatives).

Page 141: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

133

Supplementary Material

AM176844, uncultured bacterium, mangrove sedimentHQ203915, uncultured bacterium, seawater

FJ497639, uncultured gamma proteobacterium, Vailulu’u SeamountARB_3553A2D9, denovo1695 GII2.1_1026

Fr670408, uncultured gamma proteobacterium, microbial mat from Eiffel Tower Lucky Strike vent area Ef687244, uncultured bacterium, iron−oxidizing mat, Chefren mud volcano

KC682859, uncultured bacterium, surface of a basalt collected from Kilo Moana vent field, Lau BasinAB476269, uncultured bacterium, ventral setae of Shinkaia crosnieriHq153952, uncultured bacterium, white filamentous microbial mat in the photic zone of a shallow hydrothermal vent

FJ787300, uncultured bacterium, gas hydrate sediments of the Kazan Mud Volcano, East Mediterranean SeaKF624162, uncultured bacterium, oxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden Sea

HQ191073, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaARB_9F2BCD1A, denovo2124 GI1.1_1929AB015250, uncultured gamma proteobacterium, deepest cold−seep area of the Japan Trench

ARB_27D9B048, denovo3381 GII3.1_4329Fn553471, uncultured sediment bacterium, Logatchev hydrothermal vent field,oceanic sediment

AB476207, uncultured bacterium, ventral setae of Shinkaia crosnieriAB476206, uncultured bacterium, ventral setae of Shinkaia crosnieriAB530202, uncultured bacterium, marine sediment

HQ191082, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaAM882552, uncultured gamma proteobacterium, sediment from oil polluted water from petrochemical treatment plantKF624159, uncultured bacterium, oxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden Sea

ARB_17F65DD9, denovo2929 GII1.1_4344ARB_17F65DD9, denovo1483 GI1.1_747

KF624231, uncultured bacterium, suboxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden SeaJQ287244, uncultured bacterium, East Pacific RiseJQ287289, uncultured bacterium, East Pacific Rise

ARB_585E8CC4, denovo2756 GI3.1_2429FJ905708, uncultured bacterium, iron oxide sediments, Volcano 1, Tonga Arc

KF624156, uncultured bacterium, oxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden SeaARB_83A518EC, denovo2486 GI1.1_92ARB_83A518EC, denovo1329 GII3.1_629

FJ205286, uncultured gamma proteobacterium, active hydrothermal field sediments, depth:1922mEf687482, uncultured bacterium, sediment underneath an iron−oxidizing mat, Chefren mud volcano

ARB_8F6B21AE, denovo413 GI1.1_285AM882574, uncultured gamma proteobacterium, sediment from oil polluted water from petrochemical treatment plant

FJ712538, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean SeaFn397820, uncultured gamma proteobacterium, GeoB 9072−1 anoxic sub−surface mud volcano

AJ535246, uncultured gamma proteobacterium, marine sediment above hydrate ridgeAM745138, uncultured gamma proteobacterium, marine sediments

Fn553607, uncultured sediment bacterium, Logatchev hydrothermal vent sediment at 3224 mAF154089, uncultured hydrocarbon seep bacterium BPC036FJ712588, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean Sea

ARB_71C0FE2D, denovo2398 GII3.1_2072ARB_DFAE6D84, denovo669 GI2.1_13

FJ264668, uncultured bacterium, methane seep sedimentJN977166, uncultured bacterium, Jiaozhao Bay sediment

EU652548, uncultured bacterium, Yellow Sea sedimentARB_1103BE8F, denovo1699 GI2.1_3700

ARB_2C6B48A0, denovo576 GI1.1_851ARB_40FFFC75, denovo472 GI1.1_1416

Hq153917, uncultured bacterium, white microbial film shallow hydrothermal ventHq153955, uncultured bacterium, white filamentous microbial mat shallow hydrothermal vent

ARB_97098AEA, denovo2230 GI2.1_1696HQ191071, uncultured sediment bacterium, Janssand intertidal sediment; German Wadden SeaEF644802, uncultured bacterium, MAR Logatchev hydrothermal vent system, 3000m water depth, sulfide sample

Jf344197, uncultured gamma proteobacterium, petroleum spot over marine sediments ARB_6C6887C4, denovo1420 GI1.1_961

KF268708, uncultured bacterium, marine sedimentJf344186, uncultured gamma proteobacterium, petroleum spot over marine sediments

AB271125, Oligobrachia mashikoi endosymbiont GAB271122, Oligobrachia mashikoi endosymbiont D

AB271124, Oligobrachia mashikoi endosymbiont FAB271120, Oligobrachia mashikoi endosymbiont B

AB252051, Oligobrachia mashikoi endosymbiont AEU491817, uncultured bacterium, seafloor lavas from the East Pacific Rise

JF344283, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachJQ580274, uncultured gamma proteobacterium, sediments from Rodas Beach polluted with crude oil

Fn397817, uncultured gamma proteobacterium, GeoB 9072−1 anoxic sub−surface mud volcanoDQ925898, uncultured bacterium, hydrothermal vent chimneys of Guaymas Basin

AB476202, uncultured bacterium, ventral setae of Shinkaia crosnieriFn397818, uncultured gamma proteobacterium, GeoB 9072−1 anoxic sub−surface mud volcano

FJ497576, uncultured gamma proteobacterium, Vailulu’u SeamountFJ497611, uncultured gamma proteobacterium, Vailulu’u Seamount

FJ497565, uncultured gamma proteobacterium, Vailulu’u SeamountFJ497622, uncultured gamma proteobacterium, Vailulu’u Seamount

ARB_F1FB6017, denovo1531 GI2.1_1865ARB_BF291963, denovo1853 GII2.1_888

KF624239, uncultured bacterium, suboxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden SeaARB_2B841FB1, denovo642 GI3.1_474

FJ712548, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean SeaFJ264584, uncultured bacterium, methane seep sediment

JF344299, uncultured gamma proteobacterium, hydrocarbon polluted marine sediments from figueiras beachFJ905702, uncultured bacterium, iron oxide sediments, Volcano 1, Tonga Arc

ARB_D102CB76, denovo2637 GI1.1_1578ARB_8C35E7C7, denovo309 GII3.1_11763

EU491119, uncultured bacterium, seafloor lavas from the Loi’hi Seamount Pisces Peak X2Fn553445, uncultured sediment bacterium, Logatchev hydrothermal vent field,oceanic sediment

JQ579946, uncultured gamma proteobacterium, sediments from Figueiras Beach

uncu

lture

d

JX559252, uncultured bacterium, antarctic aerosolARB_D18874D7, denovo1457 GI2.1_998ARB_D18874D7, denovo3229 GII2.1_12387

JX570601, endosymbiont of Ridgeia piscesaeKC682827, uncultured bacterium, surface of a basalt collected from Kilo Moana vent field, Lau Basin

Fr670405, uncultured gamma proteobacterium, microbial mat from Eiffel Tower Lucky Strike JQ200201, uncultured bacterium, seawater; next to dolphin Y

AB495251, Cocleimonas flava, marine molluskJn662081, uncultured gamma proteobacterium, particulate detritus

Eu107481, uncultured Leucothrix sp., Okinawa Trough at depth of 1309 metersAB476220, uncultured bacterium, ventral setae of Shinkaia crosnieri

EU555124, uncultured bacterium, Dudley hydrothermal vent

Cocle

imonas

Thio

tric

hale

s_T

hio

tric

hace

ae

0.05

AF532775, Candidatus Isobeggiatoa divolgata, marine sedimentAF532769, Candidatus Isobeggiatoa divolgata, Wadden Sea sedimentARB_D11B76E1, denovo3026 GII2.1_578ARB_D11B76E1, denovo976 GI2.1_627

AB108786, Candidatus Isobeggiatoa divolgata, marine sedimentARB_D764416C, denovo1446 GI2.1_519FJ875195, uncultured Beggiatoa sp., marine sediment underlying a salmon farm

FN561862, Candidatus Isobeggiatoa divolgata, marine sedimentAF532771, uncultured Beggiatoa sp., marine sediment

ARB_6BD95FAD, denovo1168 GI2.1_4524

Candidatus Isobeggiatoa

ARB_9D6DFEEE, denovo3057 GII2.1_5949AF532773, Candidatus Parabeggiatoa communis, marine sediment

AF532772, Candidatus Parabeggiatoa communis, marine sedimentARB_C4375616, denovo1726 GI1.1_3844

JN793555, uncultured Beggiatoa sp., hydrothermal seep at Guaymas BasinFJ875199, uncultured Beggiatoa sp., marine sediment underlying a salmon farm

Candidatus Parabeggiatoa

FR847872, Candidatus Halobeggiatoa sp. HMW−W562, white microbial matFr847873, Candidatus Halobeggiatoa sp. HMW−W572, white microbial matFR847870, Candidatus Halobeggiatoa sp. HMW−W520, white microbial mat

FR847871, Candidatus Halobeggiatoa sp. HMW−S2528, white microbial matARB_9D825299, denovo578 GI1.1_2396

FN811663, Thiomargarita namibiensis, sediment surfaceFr690927, Candidatus Thiomargarita nelsonii, sedimentFR690929, Candidatus Thiomargarita nelsonii, sediment

FR690966, Candidatus Thiomargarita nelsonii, sedimentFR690953, Candidatus Thiomargarita nelsonii, sediment T

hio

marg

arita

ARB_4BFED579, denovo1545 GII2.1_7825FR690945, Candidatus Thiomargarita nelsonii, sediment

FR690946, Candidatus Thiomargarita nelsonii, sedimentFR690961, Candidatus Thiomargarita nelsonii, sediment

ARB_330D7B34, denovo283 GII2.1_1879ARB_330D7B34, denovo347 GI2.1_475

FR690894, Thiomargarita namibiensis, sedimentFR827866, Thiomargarita namibiensis, intertidal mud flat

JN793554, uncultured Beggiatoa sp., hydrothermal seep at Guaymas BasinARB_9FCD4774, denovo879 GI3.1_3565

Thio

tric

hale

s_T

hio

tric

hace

ae

0.05Thio

marg

arita

Figure S2: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Gammaproteobacte-ria, order: Thiotrichales, family: Thiotrichaceae (black bold: sequences retrieved in this study and closest relatives).

Page 142: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

134

EU592480, uncultured bacterium, hypersaline sedimentARB_7060483C, denovo2020 GI3.1_5100

GQ500708, uncultured bacterium, Charon’s Cascade, sandy clastic sedimentsFq658834, uncultured soil bacterium, PAH−contaminated soil

DQ137968, uncultured bacterium, wetlandARB_10DE197D, denovo40 GI2.1_3162FJ516989, uncultured Desulfobacteraceae bacterium, upper sediment

GU208294, uncultured prokaryote, Dongping Lake sedimentHM243855, uncultured bacterium, middle sediment from Honghu Lake

HQ003578, uncultured delta proteobacterium, Carrizo shallow lakeKF268888, uncultured bacterium, marine sediment

ARB_7004B897, denovo93 GII2.1_1855ARB_7004B897, denovo1972 GI3.1_386AB530221, uncultured bacterium, marine sediment

AB694469, uncultured bacterium, Deep−sea sediment at a depth 7111m.Jn511815, uncultured organism, Guerrero Negro Hypersaline Mat

Jn537626, uncultured organism, Guerrero Negro Hypersaline MatGQ246439, uncultured bacterium, North Yellow Sea sediments

ARB_6181BE0B, denovo348 GI1.1_2389FJ716464, uncultured bacterium, Frasassi cave system, anoxic lake water

ARB_8D6EA0F6, denovo3241 GII3.1_2770JQ925081, uncultured bacterium, cold seep sediment from oxygen mimimum zone in Pakistan Margin

JQ580439, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilJN207182, uncultured delta proteobacterium, Namako−ike surface sediment (0−1 cm)

JQ036264, uncultured bacterium, Eel River Basin sediment 2688Gu302479, uncultured bacterium, marine sediments Mississippi CanyonHQ691991, uncultured delta proteobacterium, stratified lagoon

AF029045, benzene mineralizing consortium clone SB−21DQ394937, uncultured Desulfobacterium sp., harbor sediment

GQ246425, uncultured bacterium, North Yellow Sea sedimentsEF125430, uncultured bacterium, mangrove soilDQ811822, uncultured delta proteobacterium, mangrove soil

ARB_5BA65306, denovo2937 GI3.1_931ARB_7BD0700F, denovo1902 GII1.1_5946

FN396676, uncultured marine bacterium, Arctic marine surface sedimentARB_666D9846, denovo1315 GI2.1_3389

ARB_D6FD159, denovo3965 GII2.1_38FJ485073, uncultured delta proteobacterium, biomat in El Zacaton at 114m depthGu127072, uncultured delta proteobacterium, sediment; anoxic zone

DQ831550, uncultured delta proteobacterium, marine sedimentARB_D788D99F, denovo2296 GII3.1_7171

ARB_3535E826, denovo1384 GI2.1_3008HQ178897, uncultured bacterium, sediment treated with Mr. Frosty slow freeze method

Gu302445, uncultured bacterium, marine sediments Mississippi Canyon Jf344708, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

GU362957, uncultured bacterium, marine sediment from the South China SeaARB_AB57ED2D, denovo1051 GII3.1_1604

JQ816986, uncultured bacterium, subseafloor sediment at the Good Weather RidgeDQ811813, uncultured delta proteobacterium, mangrove soil

ARB_8F5D807B, denovo60 GII3.1_7337KC470953, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

Jn518898, uncultured organism, Guerrero Negro Hypersaline Mat DQ811811, uncultured delta proteobacterium, mangrove soil

EU385682, uncultured bacterium, subseafloor sediment of the South China SeaFJ628273, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 50

AJ241002, uncultured delta proteobacterium Sva0516Jf344210, uncultured delta proteobacterium, petroleum spot over marine sediments

ARB_E20D5C9D, denovo188 GI1.1_278FJ437826, uncultured bacterium, Green Lake chemocline at 20.5 m water depth

AY216442, uncultured bacterium, temperate estuarine mudJQ580448, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_4EA6FD94, denovo1367 GII1.1_529AJ430774, delta proteobacterium EbS7, sediment

ARB_A61B58E9, denovo3408 GII2.1_14147FJ437866, uncultured bacterium, Green Lake at 25 m water depth

ARB_A26C8AD9, denovo3152 GII2.1_5152HQ003580, uncultured delta proteobacterium, Carrizo shallow lake

HQ691995, uncultured delta proteobacterium, stratified lagoonJQ580482, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_39E47D34, denovo2286 GI1.1_833Hm598151, uncultured bacterium, 1400 m depth surface sediment Xisha Trough, China Sea

ARB_F317CBC0, denovo1489 GII3.1_3131ARB_B01D0B7D, denovo988 GI3.1_3020

JF495321, uncultured bacterium, sediment from anoxic fjordAF154102, uncultured hydrocarbon seep bacterium GCA017

JQ580472, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilARB_5B4D5DA9, denovo2879 GI1.1_2108

AY542227, uncultured delta proteobacterium, Gulf of Mexico seafloor sedimentsJQ580206, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_56A14516, denovo843 GII2.1_2149ARB_56A14516, denovo558 GI2.1_1507

Jn534265, uncultured organism, Guerrero Negro Hypersaline MatJf344192, uncultured delta proteobacterium, petroleum spot over marine sediments

ARB_85297555, denovo2583 GI2.1_230Jn515097, uncultured organism, Guerrero Negro Hypersaline Mat

FJ484510, uncultured delta proteobacterium, wall biomat sample in El Zacaton at 53m depthARB_EA0946DA, denovo2630 GII2.1_2303

JQ817357, uncultured bacterium, subseafloor sediment at the Formosa RidgeJQ925098, uncultured bacterium, cold seep sediment from oxygen mimimum zone in Pakistan Margin

ARB_170270B2, denovo244 GII1.1_367HQ588498, uncultured bacterium, Amsterdam mud volcano sedimentJN977161, uncultured bacterium, Jiaozhao Bay sediment

HM231166, uncultured bacterium, coconut husk retting soilEU592498, uncultured bacterium, hypersaline sediment

HQ174924, uncultured bacterium, Cherokee Road Extension Blue HoleJQ816148, uncultured bacterium, subseafloor sediment at the Deep Basin

AM745154, uncultured delta proteobacterium, marine sedimentsJQ580310, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilARB_68DB4272, denovo228 GII2.1_2567Jn518931, uncultured organism, Guerrero Negro Hypersaline Mat

FJ437756, uncultured bacterium, Green Lake chemocline at 20.5 m water depthAB630766, uncultured bacterium, aquatic moss pillarsFJ716322, uncultured bacterium, Sawmill Sink water column, at 10.3 m water depth

ARB_9AE7AD4E, denovo2234 GI1.1_2318ARB_364C0C47, denovo3678 GII1.1_6861ARB_5D22E99, denovo730 GI1.1_3588

AJ237601, Desulfobacterium aniliniDQ394914, uncultured Desulfobacterium sp., harbor sediment

AF121886, bacterium 2BP−6Ab763347, Desulfobacterium sp. DS, a river sediment contaminated chlorinated volatile organic compounds

GQ354966, uncultured delta proteobacterium, springFJ517129, uncultured Desulfobacteraceae bacterium, water

FJ516988, uncultured Desulfobacteraceae bacterium, upper sedimentHM066266, uncultured bacterium, Texas state well #DX 68−23−616A

JQ245553, uncultured bacterium, mud volcanoEU245563, uncultured organism, hypersaline microbial mat

Jn497298, uncultured organism, Guerrero Negro Hypersaline Mat ARB_1537C093, denovo32 GII3.1_1946

Jn529954, uncultured organism, Guerrero Negro Hypersaline Mat ARB_9B500F68, denovo614 GII3.1_2698

FR682646, uncultured delta proteobacterium, hydrothermal sedimentsFJ485020, uncultured delta proteobacterium, biomat in El Zacaton at 114m depthFJ485340, uncultured delta proteobacterium, wall biomat sample in El Zacaton at 273m depthJn532040, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_E828CF31, denovo3306 GII3.1_6873FM164951, uncultured Desulfobacterium sp., hot spring

AF026998, unidentified delta proteobacterium OPB16EF205578, uncultured delta proteobacterium, geothermal spring mat

FR682638, uncultured delta proteobacterium, hydrothermal sediments

uncu

lture

d

DQ415831, uncultured bacterium, Frasassi sulfidic cave stream biofilmJN397733, uncultured bacterium, river bankARB_9F80C0F0, denovo460 GII2.1_6145

AJ576343, uncultured bacterium, hindgut homogenate of Pachnoda ephippiata larva

Desulfarculus

ARB_2EB6DECF, denovo3036 GII2.1_284ARB_2EB6DECF, denovo38 GI2.1_214

ARB_986F247C, denovo2054 GII3.1_553EU487879, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bed

ARB_783FDF21, denovo1576 GII3.1_1060FM242399, uncultured delta proteobacterium, sediment

ARB_FA468FAA, denovo730 GII3.1_1182KF697533, uncultured bacterium, activated sludge from inclined plate membrane bioreactor

ARB_740BDBE3, denovo1219 GI1.1_1259

uncultured

De

sulfa

rcu

lale

s_D

esu

lfarc

ula

cea

e

0.05

Figure S3: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Deltaproteobacteria, order: Desulfarculales (black bold: sequences retrieved in this study and closest relatives).

Page 143: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

135

Supplementary Material

133 sequences (seq), uncultured bacteria, including 8 seq obtained by Pfeffer et al., 2012,and Lake Grevelingen surface sediments, August, station 3 (3 seq),Lake Grevelingen surface sediments, August, station 1 (1 seq)

85 seq, uncultured bacteria related to Desulforhopalus sp.,Lake Grevelingen surface sediments, March and August, station 1 (2 seq),Lake Grevelingen surface sediments, March and August, station 2 (2 seq)Lake Grevelingen surface sediments, August, station 3 (4 seq)

9 seq, uncultured bacteria, including Lake Grevelingen surface sediments, August, station 3 (2 seq)

12 seq, uncultured bacteria, including Lake Grevelingen surface sediments, August, station 3 (2 seq)

Am176855, uncultured bacterium, mangrove sediment

131 seq, uncultured Desulfopila sp.,including Lake Grevelingen surface sediments, March, station 1 (1 seq),Lake Grevelingen surface sediments, August, station 2 (1 seq), Lake Grevelingen surface sediments, August, station 3 (2 seq)

30 seq, uncultured bacteria related to Desulfotalea sp.,including Lake Grevelingen surface sediments, March, station 1 (1 seq),Lake Grevelingen surface sediments, August, station 3 (1 seq),

43 seq, uncultured bacteria (SEEP−SRB4),including Lake Grevelingen surface sediments, March and August, station 2 (2 seq),

2 seq, uncultured bacteria, sediments from Gulf of Mexico and Peru Margin

59 seq, uncultured bacteria,including Lake Grevelingen surface sediments, March, station 1 (1 seq) and 3 (2 seq),Lake Grevelingen surface sediments, August, station 2 (5 seq)

35 seq, uncultured bacteria related to Desulfofustis sp.

2 seq, uncultured bacteria

273 seq, uncultured bacteria related to Desulfocapsa sp.including Lake Grevelingen surface sediments, March, station 1 (1 seq) and 2 (2 seq),Lake Grevelingen surface sediments, August, station 2 (2 seq) and 3 (1 seq),

2 seq, uncultured bacteria, microbial mat and muddy salt marsh sediment

70 seq, uncultured bacteria related to Desulfocapsa sp.,including Lake Grevelingen surface sediments, March, station 1 (1 seq) and 2 (1 seq),Lake Grevelingen surface sediments, August, station 1 (1 seq)

14 seq, uncultured bacteria, including Lake Grevelingen surface sediments, August, station 2 (1 seq)

16 seq, uncultured bacteria, subsurface aquifer sediment and volcanic soil

Jx223964, uncultured bacterium, subsurface aquifer sediment

30 seq, uncultured bacteria, including 1 seq obtained by Pfeffer et al., 2012,Lake Grevelingen surface sediments, August, station 2 (1 seq) and station 3 (4 seq)

5 seq, uncultured bacteria,

2 seq, uncultured bacteria, subsurface aquifer sediments

6 seq, uncultured bacteria, including Lake Grevelingen surface sediments, March, station 1 (1 seq) and station 2 (1 seq), and August, station 2 (1 seq)

478 seq, uncultured bacteria related to Desulfobulbus sp.(see Desulfobulbus tree; following figure)

44 seq, uncultured bacteria related to Desulfurivibrio sp.

42 seq, uncultured bacteria, including 2 seq obtained by Malkin et al., 2014, 1 seq by Pfeffer et al., 2012, Lake Grevelingen surface sediments, March, station 1 (2 seq) and station 3 (2 seq),and August, station 2 (2 seq) and station 3 (2 seq)

7 seq, uncultured bacteria, including 5 seq obtained by Pfeffer et al., 2012

2 seq, uncultured bacterium, marine sediment (1 seq) and Lake Grevelingen surface sediments, August, station 3 (1 seq)

0.10

including Lake Grevelingen surface sediments, March and August, station 1 (2 seq)

De

sulfo

ba

cte

rale

s_D

esu

lfob

ulb

ace

ae

Figure S4a: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Deltaproteobacteria, order: Desulfobacterales, family: Desulfobulbaceae (black bold: sequences retrieved in this study and clos-est relatives).

Page 144: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

136

group 1: 18 sequences uncultured bacteria

group 2: 7 sequences uncultured bacteria

group 3: 3 sequences uncultured bacteria

group 4: 24 sequences uncultured bacteria

Dq676284, uncultured delta proteobacterium, suboxic freshwater−pond sedimentCp002364, Desulfobulbus propionicus DSM 2032, freshwater mudAy548789, Desulfobulbus propionicusCp002364, Desulfobulbus propionicus DSM 2032, freshwater mud

Ay160859, uncultured bacterium, Cubitermes orthognathusAb189697, uncultured bacterium, Termes comis

Af482433, uncultured bacterium, granular sludgeX95180, Desulfobulbus elongatusDq831531, uncultured Desulfobulbus sp., brackish sedimentJq815615, uncultured Desulfobulbus sp., Tinto River sediment (extreme acid environment)

KF020725, Desulfomicrobium sp. enrichment culture clone LDC−, soilGq133652, uncultured bacterium, ASBR reactor treating swine waste

Eu104773, uncultured delta proteobacterium, anaerobic biofilm reactorGq138460, uncultured bacterium, ASBR reactor treating swine waste

Gq138956, uncultured bacterium, ASBR reactor treating swine wasteKC260892, uncultured bacterium, bottlenose dolphin mouthEu488160, uncultured bacterium, siliciclastic sediment from Thalassia sea grass bed

Hm069062, uncultured bacterium, sediment of Rupa LakeAb546247, Desulfobulbus rhabdoformis, sludge sample of crude−oil storage tankU12253, Desulfobulbus rhabdoformis

Ef442993, Desulfobulbus sp. DSM 2033Jn688053, Desulfobulbus sp. S4, microbial consortium enriched from mangrove soilHM750216, Desulfobulbus alkaliphilusAJ866934, Desulfobulbus mediterraneus, tidal flat sedimentEU142055, uncultured bacterium, methane seep sediment

AF354663, Desulfobulbus mediterraneusAUCW01000075, Desulfobulbus mediterraneus DSM 13871Eu528183, uncultured bacterium, sedimentAF099055, psychrophilic sulfate−reducing bacterium LSv55AM404380, uncultured delta proteobacterium, marine sedimentFj716980, uncultured bacterium, marine sediment from Cullercoats, Northumberland, United Kingdom Am882629, uncultured delta proteobacterium, sediment from oil polluted water from petrochemical treatment plantAb110550, Desulfobulbus japonicus, harbor sediment in JapanAB110549, Desulfobulbus japonicus, harbor sediment in Japan

U85473, Desulfobulbus sp. BG25JQ209390, uncultured bacterium, bottlenose dolphin mouthJQ211736, uncultured bacterium, bottlenose dolphin mouthJQ215032, uncultured bacterium, bottlenose dolphin mouth

Gq134058, uncultured bacterium, ASBR reactor treating swine wasteAB243815, uncultured bacterium, Niigata oil well

FJ792164, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFr853023, uncultured bacterium, active hydrothermal chimney, Juan de Fuca Ridge

Aj969450, uncultured delta proteobacterium, rock fragments, Rainbow, deep sea hydrothermal vent fieldFr852968, uncultured bacterium, active hydrothermal chimney, Dudley, Juan de Fuca Ridge

Fr853007, uncultured bacterium, active hydrothermal chimney, Dudley, Juan de Fuca RidgeAy672535, uncultured bacterium, hydrothermal vent, East Pacific Rise, Pacific OceanAy672502, uncultured bacterium, hydrothermal vent, East Pacific Rise, Pacific Ocean

AM404330, uncultured delta proteobacterium, marine sedimentAF420336, uncultured proteobacteriumAy592579, uncultured bacterium, Napoli mud volcano, Eastern Mediterranean, sediment layer 0−8 cmGU584659, uncultured bacterium, marine hydrocarbon seep sedimentJQ287080, uncultured bacterium, East Pacific RiseFj5746, uncultured bacterium, iron oxide sediments, Tonga Arc

EU555141, uncultured bacterium, Dudley hydrothermal ventLake Grevelingen surface sediments, March, station 3 AM268767, uncultured prokaryote, hot fluid vent, site Irina II

group 5: 6 sequences uncultured bacteria

Jn874112, uncultured bacterium, TrapR1, East Pacific RiseAY327886, uncultured delta proteobacterium, surface of vent snail footJn873947, uncultured bacterium, TrapR1, East Pacific Rise

Fr852967, uncultured bacterium, active hydrothermal chimney, Dudley, Juan de Fuca Ridge

group 6: 41 seq, mainly uncultured bacteria and 2 seq obtained by Pfeffer et al., 2012 (JX091042.2; JX091026.1)

Ay531563, uncultured bacterium, Indian Ocean hydrothermal ventAY531586, uncultured bacterium, Indian Ocean hydrothermal ventKc682614, uncultured bacterium, inactive hydrothermal sulfide chimney, Tui Malila vent field, Lau Basin

Lake Grevelingen surface sediments, August, station 3Lake Grevelingen surface sediments, August, station 3Lake Grevelingen surface sediments, August, station 3

JX224538, uncultured bacterium, subsurface aquifer sedimentJX225752, uncultured bacterium, subsurface aquifer sedimentJX225113, uncultured bacterium, subsurface aquifer sedimentLake Grevelingen surface sediments, March, station 3 Lake Grevelingen surface sediments, August, station 2

Lake Grevelingen surface sediments, March, station 1 KF624269, uncultured bacterium, anoxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

Lake Grevelingen surface sediments, March, station 1 HM992531, uncultured Desulfobacteraceae bacterium, oil sands tailings settling basin

AB525452, uncultured bacterium, deep−sea methane seep sediment off Joetsu, Japan SeaFn554116, uncultured sediment bacterium, Logatchev hydrothermal vent field,Anya’s GardenJn662305, uncultured delta proteobacterium, particulate detritus, c tubeworm Ridgeia piscesae

Lake Grevelingen surface sediments, March, station 2 Lake Grevelingen surface sediments, August, station 3

HQ588389, uncultured bacterium, Amsterdam mud volcano sedimentHq153922, uncultured bacterium, white microbial film on pebble substrate in the aphotic zone of a shallow hydrothermal vent

group 7: 5 sequences uncultured bacteria, carbonate chimney in Lost City hydrothermal vent

Hq153967, uncultured bacterium, white filamentous microbial mat in the photic zone of a shallow hydrothermal ventKC682422, uncultured bacterium, substrate mineral chips

JF738104, uncultured bacterium, Lau Basin hydrothermal sulfide chimneyFR852988, uncultured bacterium, active hydrothermal chimney collected from the Dudley site of Juan de Fuca Ridge

group 8: 40 seq; including 17 seq obtained by Schauer et al., 2014; 11 seq by Pfeffer et al. 2012; 7 seq by Malkin et al., 2014; 4 seq uncultured bacteria obtained from marine sediment

JF268391, uncultured bacterium, Hikurangi margin, Wairarapa − Takahae, station 309; SO191_3_309Gu302481, uncultured bacterium, marine sediments (0−24 m sediment depth) from Mississippi Canyon 118, Gulf of Mexico

GU208270, uncultured prokaryote, Dongping Lake sedimentFQ658831, uncultured soil bacterium, PAH−contaminated soil; retention systems which treat road runoffs

group 9 : 59 sequences, including Lake Grevelingen surface sediments, i.e. 3 seq station 1, March; 1 seq station 1, August; 1 seq station 2, March; 2 seq station 2, August; 2 seq station 3, March; 3 seq station 3, August

FJ264588, uncultured bacterium, methane seep sedimentFJ712577, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean SeaFJ264787, uncultured bacterium, methane seep sedimentGU180162, uncultured bacterium, sediment and soil slurryHQ153843, uncultured bacterium, white microbial film on pebble substrate in the aphotic zone of a shallow hydrothermal vent, Volcano 1, South Tonga Arc at a depth of 189.1 m

group 10: 9 seq, uncultured bacteria, marine sediment

FJ358873, uncultured delta proteobacterium, marine reef sandy sedimentFR744613, uncultured bacterium, production water from the Halfdan oil fieldAj535237, uncultured delta proteobacterium, marine sediment above hydrate ridge

Lake Grevelingen surface sediments, August, station 2 JN256008, uncultured delta proteobacterium, surface of pereopod of new species of Yeti crab from mound 12JN662199, uncultured delta proteobacterium, particulate detritus collected from grabs of the vestimentiferan tubeworm Ridgeia piscesae in low−flow environmentAY355303, uncultured bacterium, epibiontic

group 11: 113 seq, including Lake Grevelingen surface sediments, i.e. 7 seq station 1, March; 6 seq station 2, March; 8 seq station 2, August; 1 seq station 3, March; 11 seq station 3, August

FJ813526, uncultured bacterium, marine sedimentLake Grevelingen surface sediments, August, station 2

KF596661, uncultured bacterium, soilFM179887, uncultured delta proteobacterium, marine sediments

AY771956, uncultured delta proteobacterium, marine surface sedimentHQ588373, uncultured bacterium, Amsterdam mud volcano sediment

Lake Grevelingen surface sediments, March, station 1 Lake Grevelingen surface sediments, August, station 2

JN977397, uncultured bacterium, Jiaozhao Bay sedimentDQ831546, uncultured delta proteobacterium, marine sedimentDQ831536, uncultured delta proteobacterium, freshwater sedimentLake Grevelingen surface sediments, August, station 3Lake Grevelingen surface sediments, March, station 2 Lake Grevelingen surface sediments, August, station 2 Lake Grevelingen surface sediments, March, station 1 HQ857627, uncultured delta proteobacterium, hydrocarbon contaminated saline alkaline soils

JN860370, uncultured Desulfobulbus sp., low temperature hydrothermal oxides at the South West Indian RidgeAB722257, uncultured bacterium, freshwater iron−rich microbial matAB754135, uncultured bacterium, lake waterHM141864, uncultured bacterium, in situ samplers incubated in the Mahomet aquifer

JX434192, uncultured bacterium, gold mine boreholeJX434246, uncultured bacterium, gold mine borehole

AB188785, uncultured Desulfobulbus sp., cold seep sedimentJF344301, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

Lake Grevelingen surface sediments, August, station 3Lake Grevelingen surface sediments, August, station 1

GQ246446, uncultured bacterium, North Yellow Sea sedimentsLake Grevelingen surface sediments, March, station 2 JN977170, uncultured bacterium, Jiaozhao Bay sedimentJN977290, uncultured bacterium, Jiaozhao Bay sedimentLake Grevelingen surface sediments, August, station 2

KC682706, uncultured bacterium, inner conduit of inactive hydrothermal sulfide chimney, ABE vent field, Lau BasinJQ287075, uncultured bacterium, East Pacific Rise

KC682691, uncultured bacterium, inner conduit of inactive hydrothermal sulfide chimney, ABE vent field, Lau BasinKC470997, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

Lake Grevelingen surface sediments, August, station 3Lake Grevelingen surface sediments, March, station 3Lake Grevelingen surface sediments, August, station 3

AM882608, uncultured delta proteobacterium, sediment from oil polluted water from petrochemical treatment plantLake Grevelingen surface sediments, March, station 3

Lake Grevelingen surface sediments, March, station 1Lake Grevelingen surface sediments, August, station 3DQ833475, uncultured bacterium, sediment of Guanting ReservoirAB722237, uncultured bacterium, freshwater iron−rich microbial matAB754082, uncultured bacterium, lake waterAB240495, uncultured bacterium, PCR−derived sequence from root−tip (0 to 40 mm) of Phragmites at Sosei River in Sappro, Japan

group 12: 9 seq, uncultured bacteria, marine sediment, hydrothermal vents

Desulfobulb

us

0.1

Figure S4b: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Deltaproteobacteria, order: Desulfobacterales, family: Desulfubulbaceae, genus: Desulfubulbus (black bold: sequences retrieved in this study and closest relatives).

Page 145: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

137

Supplementary Material

Kc471241, uncultured bacterium, marine sediment Okinawa TroughARB_7212B3B6, denovo1556 GII2.1_787

AY771938, uncultured delta proteobacterium, marine surface sedimentJf344655, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

Am882597, uncultured delta proteobacterium, sediment from oil polluted water Gu302441, uncultured bacterium, marine sediments Mississippi Canyon

Jf344367, uncultured delta proteobacterium, hydrocarbon polluted marine sediments Jf344621, uncultured delta proteobacterium, hydrocarbon polluted marine sediments Fn396669, uncultured marine bacterium, Arctic marine surface sediment

KF741562, uncultured bacterium, Talbert salt marsh sedimentJf344336, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

AB250573, uncultured bacterium, microbial matGQ259292, uncultured bacterium, deep sedimentARB_9B5D30FB, denovo2586 GII3.1_7180

Jq580293, uncultured delta proteobacterium, sediments Rodas Beach with crude oilDQ394998, uncultured delta proteobacterium, harbor sediment

JQ586318, uncultured bacterium, arctic marine sedimentFN600637, Olavius sp. associated proteobacterium Delta 1

Jf344583, uncultured delta proteobacterium, hydrocarbon polluted marine sediments ARB_E4F47108, denovo2691 GI3.1_2447

DQ831553, uncultured Desulfosarcina sp., marine sedimentARB_3BDC7AF5, denovo1926 GI3.1_5119

JQ579984, uncultured delta proteobacterium, sediments from Figueiras BeachFJ264667, uncultured bacterium, methane seep sediment

ARB_E6ACD753, denovo1584 GII3.1_4528AY222311, uncultured delta proteobacterium, host hindgut caecum

Am882638, uncultured delta proteobacterium, sediment from oil polluted water AXXK01000236, delta proteobacterium PSCGC 5451

ARB_AFCDBB4E, denovo2464 GI1.1_733EU925915, uncultured bacterium, sediment from station DBS1, northern Bering Sea

ARB_527A343A, denovo125 GII3.1_1095ARB_CFF4FC25, denovo2734 GII3.1_1621

ARB_5666EC06, denovo2457 GII3.1_1313ARB_35A5F4CB, denovo3527 GII3.1_3458

ARB_60341410, denovo2446 GII2.1_510ARB_7E58A22C, denovo1560 GI2.1_4724

Jf344664, uncultured delta proteobacterium, hydrocarbon polluted marine sediments DQ395018, uncultured delta proteobacterium, harbor sediment

AASZ01000534, metagenome, gutless worms collected in the Mediterranean SeaAJ620511, Olavius crassitunicatus delta−proteobacterial endo, marine sedimentJq580141, uncultured delta proteobacterium, sediments from Rodas Beach crude oil

FJ786116, uncultured bacterium, prawn farm sedimentARB_64DE2518, denovo1514 GI1.1_1793

JQ586291, uncultured bacterium, arctic marine sedimentARB_2308CAA0, denovo1932 GI1.1_3278

ARB_92BF177B, denovo3292 GII3.1_5475ARB_8199C09B, denovo1844 GII3.1_1277

Jf344447, uncultured delta proteobacterium, hydrocarbon polluted marine sediments ARB_21640F45, denovo2946 GII2.1_7360

ARB_B959CA3E, denovo1101 GI3.1_326AXXL01000032, delta proteobacterium PSCGC 5342

KC470928, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughJf344436, uncultured delta proteobacterium, hydrocarbon polluted marine sediments ARB_2375B3D7, denovo2041 GII2.1_1621ARB_2375B3D7, denovo2116 GI1.1_1921

FM242225, uncultured delta proteobacterium, sedimentARB_DCA38B18, denovo2024 GII2.1_3527

AM040123, uncultured delta proteobacterium, sandy sedimentsARB_BAC5426A, denovo3444 GII2.1_5360

Jn495286, uncultured organism, Guerrero Negro Hypersaline Mat Jf344372, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

ARB_70856E3C, denovo2194 GI2.1_783Jf344690, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

FN396636, uncultured marine bacterium, Arctic marine surface sedimentJf344696, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

ARB_44DC591A, denovo2386 GI2.1_2938ARB_35BFAC0E, denovo1701 GI2.1_3256

GQ472418, uncultured bacterium, lake sedimentFJ516927, uncultured Desulfobacteraceae bacterium, upper sediment

FJ716475, uncultured bacterium, Frasassi cave system, anoxic lake waterGU208253, uncultured prokaryote, Dongping Lake sediment

GU208215, uncultured prokaryote, Dongping Lake sedimentFj485545, uncultured delta proteobacterium, wall biomat sample in El Zacaton

AJ620510, Olavius crassitunicatus delta−proteobacterial endo, marine sedimentJQ712425, uncultured delta proteobacterium, hydrothermal vent fluids

ARB_2FDA25E1, denovo2589 GII3.1_649ARB_923AE546, denovo3737 GII2.1_1274

ARB_5F670FDF, denovo3450 GII2.1_2063Fm165241, uncultured bacterium, tubeworm Lamellibrachia sp. from cold seeps

ARB_B59ABBF4, denovo2404 GII2.1_6581ARB_BBF73459, denovo2912 GI1.1_3098

ARB_4DE6ABCD, denovo560 GII3.1_3361EF125388, uncultured bacterium, naked tidal flat sediment near mangrove

JF428822, uncultured delta proteobacterium, zero water exchange shrimp pond sedimentGQ246372, uncultured bacterium, North Yellow Sea sediments

Jf344682, uncultured delta proteobacterium, hydrocarbon polluted marine sedimentsARB_8BA374D8, denovo1060 GII2.1_1391ARB_8BA374D8, denovo1625 GI1.1_2788

EF999398, uncultured bacterium, Pearl River Estuary sediments at 50cm depthARB_17225923, denovo35 GI1.1_4513

JX221943, uncultured bacterium, subsurface aquifer sedimentJX225840, uncultured bacterium, subsurface aquifer sediment

ARB_7896FCD9, denovo3335 GII3.1_131HQ697836, uncultured bacterium, hydrocarbon contaminated saline−alkali soil

JX240710, uncultured delta proteobacterium, coastal soilGU289732, Desulfonatronobacter acidivorans, soda lake

Desulfonatronobacter

Jf268373, uncultured bacterium, Hikurangi marginAJ704684, uncultured delta proteobacterium, marine sediment

ARB_242A7D9D, denovo1494 GII1.1_2254ARB_F85668AB, denovo305 GII3.1_161

ARB_222F3CA4, denovo584 GII3.1_2561ARB_3E5AFB8D, denovo1509 GII3.1_3622

ARB_3367F4DF, denovo977 GII3.1_2082ARB_242A7D9D, denovo2742 GI1.1_1973

ARB_5F9B7D1E, denovo1873 GII3.1_5804ARB_46062F0A, denovo791 GI3.1_1456

ARB_8C59A6BA, denovo1742 GII3.1_2529GU291339, uncultured bacterium, acid mine drainage contaminated salt marsh sedimentsFJ497421, uncultured delta proteobacterium, Vailulu’u Seamount

ARB_F6791D49, denovo1990 GII3.1_998DQ394951, uncultured delta proteobacterium, harbor sediment

Hq405625, bacterium enrichment culture clone AOM−SR−B8, enrichment anaerobic oxidation of methaneAM176871, uncultured bacterium, mangrove sediment

GQ259270, uncultured bacterium, deep sedimentAJ704683, uncultured delta proteobacterium, marine sediment

ARB_579A7566, denovo2160 GI2.1_3055ARB_42D72CEA, denovo486 GI2.1_275

Fn554099, uncultured sediment bacterium, Logatchev hydrothermal vent HQ691980, uncultured delta proteobacterium, stratified lagoon

ARB_B5B2E1EF, denovo1135 GI2.1_2083ARB_42D72CEA, denovo1053 GII1.1_751

HM141895, uncultured bacterium, in situ samplers incubated in the Mahomet aquiferARB_3B3150E, denovo2691 GII3.1_3664

HQ174942, uncultured bacterium, Cherokee Road Extension Blue HoleARB_70290A1B, denovo94 GI2.1_1300

ARB_4B1EF6D, denovo1731 GI2.1_342ARB_DBAFDA21, denovo2860 GII2.1_5120

ARB_A6EACD9, denovo1419 GII3.1_8756ARB_128F1521, denovo1842 GII2.1_12592

FM242223, uncultured delta proteobacterium, sedimentAm882554, uncultured gamma proteobacterium, sediment from oil polluted water

ARB_E2F92A82, denovo2276 GI1.1_1174Am882618, uncultured delta proteobacterium, sediment from oil polluted water

ARB_61BA32E3, denovo2873 GI3.1_134ARB_39A85234, denovo716 GII3.1_1509

ARB_E2B4FFB0, denovo287 GII2.1_10006Am882636, uncultured delta proteobacterium, sediment from oil polluted water

ARB_768D549B, denovo2723 GII3.1_9116ARB_ED8C662E, denovo3475 GII2.1_540

ARB_F56DA353, denovo3774 GII2.1_8690ARB_942E6C15, denovo717 GII3.1_5336

DQ394899, uncultured delta proteobacterium, harbor sedimentARB_1C79E7C1, denovo123 GII3.1_4596

ARB_97748389, denovo525 GI2.1_2980ARB_87BD5B07, denovo2143 GII3.1_1826

EU542429, uncultured bacterium, sediment and soil slurryHQ190518, uncultured bacterium, Zhongyuan oil fieldAm882592, uncultured delta proteobacterium, sediment from oil polluted water

Jn539400, uncultured organism, Guerrero Negro Hypersaline Mat ARB_40933119, denovo1876 GI1.1_500ARB_40933119, denovo411 GII1.1_601

AB530178, uncultured bacterium, marine sedimentAm882593, uncultured delta proteobacterium, sediment from oil polluted water

uncu

lture

d

AY851292, Algorimarina butyricaARB_C3823A0C, denovo2233 GII3.1_11024

ARB_62175954, denovo1151 GII1.1_5521ARB_62175954, denovo639 GI1.1_635

Jn493110, uncultured organism, Guerrero Negro Hypersaline Mat Jn509712, uncultured organism, Guerrero Negro Hypersaline Mat

FJ813564, uncultured bacterium, marine sedimentJn533073, uncultured organism, Guerrero Negro Hypersaline Mat

Jq580317, uncultured delta proteobacterium, sediments from Rodas Beach crude oilJf344658, uncultured delta proteobacterium, hydrocarbon polluted marine sedimentsARB_AB025D00, denovo2174 GII3.1_2112

FJ517135, uncultured Desulfobacteraceae bacterium, waterJf344433, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

DQ811834, uncultured delta proteobacterium, mangrove soilDQ811835, uncultured delta proteobacterium, mangrove soil

FJ873366, uncultured bacterium, sediments from the Okhotsk SeaARB_2B982B23, denovo78 GII3.1_1891

EU488132, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bedEU385712, uncultured bacterium, subseafloor sediment of the South China Sea

Ab177130, uncultured bacterium, ethane hydrate bearing subseafloor sediment at the Peru margin ARB_8376DD2A, denovo2689 GII2.1_12013

FN820322, uncultured bacterium, marine sediment, mud volcanoJN977360, uncultured bacterium, Jiaozhao Bay sedimentGQ357013, uncultured bacterium, methane seep sediment

ARB_609A165E, denovo1706 GII3.1_4850ARB_276A6BD5, denovo3922 GII3.1_9884

EU181465, uncultured bacterium, marine sedimentsGU982811, uncultured bacterium, marine sediment from the Western Pacific Ocean

JX138851, uncultured delta proteobacterium, marine sedimentsARB_A42CEBFC, denovo210 GI2.1_484JQ817421, uncultured bacterium, subseafloor sediment at the Formosa Ridge

JF495316, uncultured bacterium, sediment from anoxic fjordARB_CA3EC558, denovo365 GII2.1_11472

GU553724, uncultured bacterium, subseafloor sediment at the Yung−An RidgeARB_D98560D6, denovo3652 GII2.1_1870

JN469616, uncultured organism, Guerrero Negro Hypersaline Mat Jn495515, uncultured organism, Guerrero Negro Hypersaline Mat

Fj416106, uncultured bacterium, sediment fNorthern Bering SeaARB_F2527D65, denovo2119 GII3.1_5230

EU287209, uncultured bacterium, arctic surface sedimentARB_ACA31ABA, denovo1320 GI1.1_4519

AY771936, uncultured delta proteobacterium, marine surface sedimentARB_FC6CBA6, denovo2741 GI1.1_105ARB_7F8C1CB7, denovo3949 GII2.1_7114

ARB_47DE09C, denovo2094 GII3.1_9138DQ811820, uncultured delta proteobacterium, mangrove soilFN396758, uncultured marine bacterium, Arctic marine surface sediment

KF624243, uncultured bacterium, suboxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden SeaARB_DECA4E83, denovo514 GI1.1_1612

ARB_21531FCD, denovo1023 GI1.1_2208HM245653, uncultured bacterium, pockmark sediments

DQ811797, uncultured delta proteobacterium, mangrove soilARB_5498C7B, denovo1552 GI2.1_1268

ARB_F96C7440, denovo3157 GII2.1_2571JF344653, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachFJ716454, uncultured bacterium, Frasassi cave system, anoxic lake water

JF344629, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachEU266854, uncultured delta proteobacterium, tar−oil contaminated aquifer sediments

AM882620, uncultured delta proteobacterium, sediment from oil polluted water from petrochemical treatment plantARB_89F1B619, denovo3034 GII3.1_5660

Jn533825, uncultured organism, Guerrero Negro Hypersaline Mat ARB_C183B8DF, denovo1811 GII3.1_3650

ARB_F53E48CB, denovo1126 GI3.1_3827ARB_F53E48CB, denovo2039 GII1.1_3417

KF741416, uncultured bacterium, Talbert salt marsh sedimentARB_BC7F083C, denovo345 GI2.1_712

KC358194, uncultured bacterium, large−discharge carbonate springsHQ691983, uncultured delta proteobacterium, stratified lagoon

AM404371, uncultured delta proteobacterium, marine sedimentJQ580124, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilARB_FBF16789, denovo2108 GII3.1_2264ARB_FBF16789, denovo2361 GI2.1_1158

EU048693, uncultured bacterium, marine sediment, South Slope, South China SeaARB_49159E95, denovo2425 GII3.1_9745

GQ261770, uncultured delta proteobacterium, deep sea sediment associated with whale fallsARB_8ADD6B04, denovo2164 GII3.1_4296ARB_857E05F1, denovo1899 GII2.1_12697

EU652612, uncultured bacterium, Yellow Sea sedimentGQ348508, uncultured delta proteobacterium, Saanich Inlet, 200 m depth

EU386062, uncultured bacterium, subseafloor sediment of the South China SeaJf344630, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

FJ517136, uncultured Desulfobacteraceae bacterium, waterARB_AF14643E, denovo783 GII1.1_4238

ARB_88ACEA80, denovo2073 GI3.1_2120ARB_24074B89, denovo2258 GII3.1_4875

EU488487, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bedDQ811826, uncultured delta proteobacterium, mangrove soil

Jn506728, uncultured organism, Guerrero Negro Hypersaline Mat JN512076, uncultured organism, Guerrero Negro Hypersaline Mat

Jn535830, uncultured organism, Guerrero Negro Hypersaline Mat ARB_C0585716, denovo1150 GII2.1_2398

ARB_6DECB96F, denovo1068 GII2.1_4639FJ746161, uncultured bacterium, ocean sediment, 5306 m water depth, during Cruise Knox02rrAB858610, uncultured bacterium, subseafloor massive sulfide deposit

ARB_11D292FD, denovo839 GII1.1_186Jn449287, uncultured organism, Guerrero Negro Hypersaline Mat

Jn461602, uncultured organism, Guerrero Negro Hypersaline Mat ARB_BC1B7B7C, denovo2903 GII1.1_4184

ARB_DE50BAE, denovo2728 GII2.1_7574ARB_E97A75AA, denovo3575 GII3.1_6200

JN672650, uncultured bacterium, surface layer sediment of the East China seaARB_8D6E9E8D, denovo264 GII2.1_2368

EU246034, uncultured organism, hypersaline microbial matAJ620501, Olavius ilvae Delta 3 endosymbiont, marine sediment

JQ580056, uncultured delta proteobacterium, sediments from Figueiras BeachARB_3CB8B01, denovo822 GII2.1_5065AM493254, Olavius algarvensis associated proteobacterium Del, marine sediment

ARB_3A6149F0, denovo438 GI1.1_1942AB188777, uncultured delta proteobacterium, cold seep sediment

ARB_53E34389, denovo597 GI2.1_641ARB_BEC97052, denovo2382 GI2.1_1911

ARB_5769CAF9, denovo1696 GII2.1_8379Fr853018, uncultured bacterium, active hydrothermal chimney Juan de Fuca Ridge

Fn553934, uncultured sediment bacterium, Logatchev hydrothermal vent fieldFJ535333, uncultured bacterium, sediment

ARB_7EEB8DA1, denovo1999 GI2.1_157ARB_BD08D4DA, denovo2785 GII2.1_4055ARB_7171ABDD, denovo2495 GI3.1_1328AM889193, uncultured delta proteobacterium, Sediments

Jn860325, uncultured Desulfobacteraceae bacterium, low temperature hydrothermal oxides South West Indian RidgeAJ535249, uncultured delta proteobacterium, marine sediment above hydrate ridge

ARB_F2B93C6B, denovo3181 GII3.1_3701ARB_742457D2, denovo3235 GII3.1_3463

Gu302438, uncultured bacterium, marine sediments Mississippi CanyonARB_EC6D159F, denovo3009 GII2.1_4940

ARB_86E2CFD9, denovo1535 GI2.1_196HQ588558, uncultured bacterium, Amsterdam mud volcano sediment

FN820295, uncultured bacterium, marine sediment, mud volcanoFn421323, uncultured delta proteobacterium, marine sediments

ARB_EC6B3962, denovo1757 GII3.1_527ARB_1132C70D, denovo666 GII3.1_2969

JF344558, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachFJ712476, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean Sea

JQ036271, uncultured bacterium, Hydrate Ridge carbonate 2518JF344483, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

HQ588578, uncultured bacterium, Amsterdam mud volcano sedimentARB_47DF84C1, denovo133 GI2.1_1635ARB_47DF84C1, denovo1667 GII3.1_7440

JF344359, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachHQ003583, uncultured Desulfobacteraceae bacterium, Carrizo shallow lake

JQ816993, uncultured bacterium, subseafloor sediment at the Good Weather RidgeFJ716406, uncultured bacterium, Frasassi cave system, anoxic lake water

EU385687, uncultured bacterium, subseafloor sediment of the South China Sea

SE

EP

−S

RB

1

Jn474717, uncultured organism, Guerrero Negro Hypersaline Mat ARB_FCD1073C, denovo1850 GII3.1_2677DQ154853, uncultured bacterium, hypersaline microbial mat

HM141837, uncultured bacterium, in situ samplers incubated in the Mahomet aquiferJQ580445, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_C2FCF446, denovo1290 GII2.1_8359ARB_DFC45A49, denovo1961 GI3.1_4202AJ937690, uncultured bacterium, mud from a terrestrial volcano

Hq622295, bacterium enrichment culture clone Oil_45, polluted estuarine sediment PalestineCp000859, Desulfococcus oleovorans Hxd3

FJ712452, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean SeaARB_DE12A1A0, denovo106 GII3.1_8455

ARB_8A60CE22, denovo2792 GI2.1_115ARB_9DAF10FF, denovo1830 GI2.1_2003

Jn427089, uncultured organism, Guerrero Negro Hypersaline MatJn513553, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_EE004B84, denovo1007 GII1.1_6352ARB_B4C116D9, denovo2961 GII1.1_74

Ay592356, uncultured bacterium, Amsterdam mud volcano, Eastern MediterraneanDQ395063, uncultured delta proteobacterium, harbor sediment

ARB_565A6172, denovo168 GI1.1_583ARB_715AC8A9, denovo1506 GII3.1_5149

ARB_A58971EA, denovo1416 GI3.1_1406ARB_E30E2486, denovo2288 GII3.1_7373

Jf344183, uncultured delta proteobacterium, petroleum spot over marine sediments from AJ241022, uncultured delta proteobacterium Sva0405DQ394892, uncultured delta proteobacterium, harbor sediment

KF741487, uncultured bacterium, Talbert salt marsh sedimentAF286034, saltmarsh clone LCP−80

AM176857, uncultured bacterium, mangrove sedimentARB_231AEC8D, denovo1538 GII1.1_7681ARB_2DE56776, denovo1282 GII3.1_6573

ARB_26FE5501, denovo1968 GII3.1_510EU488125, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bed

JF344657, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachARB_2CB6E1DA, denovo2662 GII2.1_13512ARB_DC9284AB, denovo3677 GII3.1_10827

ARB_8D032F64, denovo159 GII1.1_1709ARB_E5FB189, denovo22 GI2.1_3017

ARB_890E687C, denovo491 GII3.1_7669ARB_9BF6A65E, denovo1751 GI2.1_3671

ARB_809B64D8, denovo303 GII2.1_13944ARB_7CA9BAA1, denovo1264 GII2.1_466

ARB_9B8D811B, denovo3373 GII1.1_2350ARB_F5D39F51, denovo89 GII3.1_2169

FJ264783, uncultured bacterium, methane seep sedimentDQ831539, uncultured delta proteobacterium, marine sedimentAM745163, uncultured delta proteobacterium, marine sediments

JF344689, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachEU290703, uncultured bacterium, sediment 4−6cm depth

JF344569, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachAB530218, uncultured bacterium, marine sediment

EU592471, uncultured bacterium, hypersaline sedimentJQ580328, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

JF344202, uncultured delta proteobacterium, petroleum spot over marine sediments from figueiras beachFJ712547, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains, East Mediterranean Sea

Desu

lfoco

ccus

ARB_A55FD4F6, denovo374 GI2.1_643JQ580337, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

JF344676, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachAy592214, uncultured bacterium, Kazan mud volcano, Eastern Mediterranean

GU553709, uncultured bacterium, subseafloor sediment at the Yung−An RidgeJQ816985, uncultured bacterium, subseafloor sediment at the Good Weather RidgeJQ580267, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_ED592767, denovo91 GII2.1_7106JQ580366, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilAy592199, uncultured bacterium, Kazan mud volcano, Eastern Mediterranean

ARB_180B36CB, denovo816 GI3.1_1982Hq405618, bacterium enrichment culture clone AOM−SR−B34, enrichment anaerobic oxidation of methane

FJ813588, uncultured bacterium, marine sedimentFN820352, uncultured bacterium, marine sediment, mud volcano

FN820349, uncultured bacterium, marine sediment, mud volcanoFN820318, uncultured bacterium, marine sediment, mud volcano

AF354163, uncultured delta proteobacterium, marine methane seepAJ535248, uncultured delta proteobacterium, marine sediment above hydrate ridge

JQ580408, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilFn553961, uncultured sediment bacterium, Logatchev hydrothermal vent field

Fm165265, uncultured bacterium, chitineous tube of the vestimentiferan tubeworm Lamellibrachia sp. JQ580515, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilHQ711383, uncultured delta proteobacterium, marine sediment

uncu

lture

d

JQ925093, uncultured bacterium, cold seep sediment from oxygen mimimum zone in Pakistan MarginARB_AC5CC84B, denovo3507 GII3.1_47

ARB_D2CDF56A, denovo3569 GII3.1_1966JF428850, uncultured bacterium, zero water exchange shrimp pond sediment

Desulfosarcina

Figure S5a: Bacterial 16S rRNA gene amplicon sequences assigned to in the class of Deltaproteobac-teria, order: Desulfobacterales, family: Desulfobacteraceae (black bold: sequences retrieved in this study and closest relatives).

Page 146: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

138

GU584771, uncultured bacterium, marine hydrocarbon seep sedimentFM242377, uncultured delta proteobacterium, sediment

ARB_7BB7CC2B, denovo1528 GI2.1_1861ARB_D1E9F9BC, denovo1688 GII2.1_1597

ARB_A0038697, denovo3562 GII3.1_6063ARB_18096835, denovo477 GII3.1_5967ARB_20167A57, denovo1173 GI1.1_1953

ARB_6E8E107, denovo1703 GII1.1_5238ARB_D0A2A25A, denovo3037 GII3.1_8560

He648188, uncultured delta proteobacterium, mercury contaminated sedimentsARB_DC50913D, denovo2564 GI2.1_308

AY177788, uncultured Desulfosarcina sp., Antarctic sedimentARB_1A22C78C, denovo2306 GII3.1_8988KF741381, uncultured bacterium, Talbert salt marsh sediment

AM745205, uncultured delta proteobacterium, marine sedimentsAJ237603, Desulfosarcina cetonicaHE797782, uncultured bacterium, anoxic brackish photosynthetic biofilms

M34407, Desulfosarcina variabilisM26632, Desulfosarcina variabilisFJ516893, uncultured Desulfobacteraceae bacterium, upper sediment

Y17286, Desulfosarcina ovataAB587702, delta proteobacterium 28bB2TJn532272, uncultured organism, Guerrero Negro Hypersaline Mat

FM242231, uncultured delta proteobacterium, sedimentARB_F37E9CA4, denovo1201 GII3.1_9072ARB_713AB900, denovo1352 GII2.1_5864

FN396663, uncultured marine bacterium, Arctic marine surface sedimentARB_CB5BEF9B, denovo2020 GII2.1_2856

JQ214259, uncultured bacterium, bottlenose dolphin mouthARB_2B785AAB, denovo3244 GII3.1_1828

EU290724, uncultured bacterium, sediment 4−6cm depthJQ215139, uncultured bacterium, bottlenose dolphin mouth

ARB_8D28BB07, denovo302 GI2.1_5299Am882600, uncultured delta proteobacterium, sediment from oil polluted water

De

sulfo

sarc

ina

ARB_48683D96, denovo1131 GI2.1_2394ARB_48683D96, denovo432 GII2.1_7718

Jf344673, uncultured delta proteobacterium, hydrocarbon polluted marine sediments ARB_DCF3EA7D, denovo3912 GII3.1_6692ARB_902EB904, denovo2089 GI1.1_200ARB_9896FE39, denovo774 GII2.1_5886

Jf344442, uncultured delta proteobacterium, hydrocarbon polluted marine sediments DQ630712, uncultured delta proteobacterium, salt marshARB_E400D27A, denovo2526 GI2.1_3217

KF925363, Candidatus Magnetananas sp. Rongcheng CityHq857738, Candidatus Magnetananas tsingtaoensis

ARB_8B4B91B, denovo1795 GII1.1_3318ARB_AB672D1F, denovo4081 GII1.1_4099

ARB_A41502C5, denovo1324 GI1.1_3580Jn438841, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_B521EEA2, denovo1656 GII2.1_1288ARB_E0A88F1, denovo2036 GI3.1_1767

ARB_E0DCFAF3, denovo1979 GI2.1_798ARB_DBAEF3D1, denovo2878 GII2.1_13288

Jn484364, uncultured organism, Guerrero Negro Hypersaline Mat JN463006, uncultured organism, Guerrero Negro Hypersaline Mat

Jn482322, uncultured organism, Guerrero Negro Hypersaline Mat Ef077226, uncultured delta proteobacterium, Gulf of Mexico, gas seep sedimen

HQ588478, uncultured bacterium, Amsterdam mud volcano sedimentFR823373, uncultured delta proteobacterium, associated sedimentFR823363, uncultured delta proteobacterium, gas seep sediment

AXAM01000010, Desulfosarcina sp. BuS5HG513093, uncultured delta proteobacterium, marine sediments

ARB_5888730, denovo1188 GII3.1_10608ARB_F0E82616, denovo420 GII2.1_2648

FR823364, uncultured delta proteobacterium, gas seep sedimentEU570889, uncultured bacterium, Nitinat Lake at a depth of 50 m

ARB_3CE6694D, denovo1374 GI2.1_1017ARB_92706565, denovo2432 GII2.1_422

ARB_B99B5F84, denovo2534 GII2.1_1677Jn520563, uncultured organism, Guerrero Negro Hypersaline Mat Jn528692, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_4101ED26, denovo1240 GI3.1_2030ARB_E852A741, denovo3862 GII2.1_3081AB630474, uncultured bacterium, aquatic moss pillars

Fj712404, uncultured bacterium, Kazan Mud Volcano, Anaximander MountainsKF268892, uncultured bacterium, marine sediment

AJ620497, Olavius algarvensis Delta 4 endosymbiont, marine sedimentFj712498, uncultured bacterium, Kazan Mud Volcano, Anaximander Mountains,

Jn387434, uncultured microorganism, Zodletone spring sedimentARB_28682B51, denovo2197 GI2.1_784

Am882623, uncultured delta proteobacterium, sediment from oil polluted water ARB_76C9EFC4, denovo149 GII1.1_819ARB_76C9EFC4, denovo2253 GI2.1_1760

ARB_8FFDF050, denovo208 GII3.1_7256ARB_19CA1269, denovo3467 GII3.1_7023

JF344320, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachARB_87AB5B4E, denovo1028 GII3.1_1044

ARB_C029BEB9, denovo2068 GII3.1_3423JF344560, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beach

ARB_85F032BD, denovo3750 GII2.1_1454ARB_DF16678C, denovo386 GII3.1_11802

FR823365, uncultured delta proteobacterium, gas seep sedimentFR823374, uncultured delta proteobacterium, associated sediment

JQ580521, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilFM242228, uncultured delta proteobacterium, sediment

JN387432, uncultured microorganism, Zodletone spring source sedimentEU290686, uncultured bacterium, sediment 10−12cm depth

FJ802207, denitrifying bacterium enrichment culture clone NO, Ebro River Delta sedimentARB_C7FB7BD3, denovo4038 GII2.1_5507

FJ802214, denitrifying bacterium enrichment culture clone NO, Ebro River Delta sedimentAB630477, uncultured bacterium, aquatic moss pillarsKC662297, uncultured bacterium, shrimp sediment

EU266914, uncultured Desulfobacteraceae bacterium, tar−oil contaminated aquifer sedimentsDQ146482, Desulfatirhabdium butyrativoransAUCU01000039, Desulfatirhabdium butyrativorans DSM 18734

ARB_34FBA2FC, denovo2748 GII3.1_5675ARB_1216BFC8, denovo523 GII3.1_10002

ARB_FC4CC3BF, denovo1568 GII3.1_8534ARB_C45E6319, denovo1783 GII3.1_3436

Jf344627, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from figueiras beachARB_D6AA1912, denovo1579 GI2.1_4294

ARB_B02A7801, denovo3004 GII3.1_6642FJ485367, uncultured delta proteobacterium, wall biomat sample in El Zacaton at 273m depth

ARB_D7DF1E65, denovo183 GII3.1_662ARB_3B67EA53, denovo1846 GI1.1_44

FJ502253, uncultured bacterium, Lake Cadagno chemocline; 12.5mKF741543, uncultured bacterium, Talbert salt marsh sediment

JF268407, uncultured bacterium, Hikurangi margin, Wairarapa − Takahae, station 309; SO191_3_309ARB_118B74EC, denovo680 GII2.1_440

HQ174906, uncultured bacterium, Cherokee Road Extension Blue HoleJf344612, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

JX504502, uncultured bacterium, oolitic sandAB530220, uncultured bacterium, marine sediment

KF596671, uncultured bacterium, soilARB_19855F1B, denovo18 GI2.1_3831

AM086163, uncultured bacterium, lake profundal sediment

De

sulfa

tirh

ab

diu

m

Jn446846, uncultured organism, Guerrero Negro Hypersaline Mat Jn455820, uncultured organism, Guerrero Negro Hypersaline Mat

Gu302437, uncultured bacterium, marine sediments Mississippi Canyon ARB_566A0BD4, denovo661 GII3.1_2960

ARB_D9F5669B, denovo10 GI2.1_3947ARB_7581FAA0, denovo3528 GII1.1_2970

ARB_4D3AC268, denovo1169 GI2.1_1067FM242402, uncultured delta proteobacterium, sediment

Am882622, uncultured delta proteobacterium, sediment from oil polluted water FM242376, uncultured delta proteobacterium, sediment

Am882604, uncultured delta proteobacterium, sediment from oil polluted water AJ237607, Desulfobacterium indolicum

FJ813595, uncultured bacterium, marine sedimentARB_3A7EB538, denovo512 GII3.1_4964

HQ178881, uncultured bacterium, sediment treated with Mr. Frosty slow freeze methodFQ659849, uncultured soil bacterium, PAH−contaminated soil; retention systems which treat road runoffs

EU385675, uncultured bacterium, subseafloor sediment of the South China SeaAY851291, Algidimarina propionicaGQ356960, uncultured bacterium, methane seep sediment

De

sulfo

ba

cte

riu

m

ARB_69CF1588, denovo3217 GII2.1_765ARB_4D5BC411, denovo2235 GII1.1_8125EF125422, uncultured bacterium, mangrove soil

JN038987, uncultured delta proteobacterium, Chongxi wetland soilDQ811793, uncultured delta proteobacterium, mangrove soil

DQ811836, uncultured delta proteobacterium, mangrove soilAY216453, uncultured bacterium, temperate estuarine mud

ARB_E9C88F6B, denovo45 GII2.1_6638ARB_B6B3A7A9, denovo3493 GII1.1_310

JX222415, uncultured bacterium, subsurface aquifer sedimentARB_A67B6D71, denovo2403 GII3.1_10460

ARB_6E5ADDED, denovo3435 GII2.1_1449ARB_6E5ADDED, denovo408 GI3.1_2254

EU290687, uncultured bacterium, sediment 10−12cm depthARB_BF175B5A, denovo2686 GII1.1_64Jn521333, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_4BCB9692, denovo2286 GII1.1_1807ARB_1C9EAE25, denovo1438 GII3.1_1

AM745219, uncultured delta proteobacterium, marine sedimentsARB_B8B5FD8, denovo1541 GII2.1_6936

ARB_EC4D5ED7, denovo1034 GII3.1_1420ARB_5FE96A7F, denovo1725 GII3.1_11421AB121109, uncultured delta proteobacterium, cold seep sediment

ARB_8270E60B, denovo4084 GII2.1_2351ARB_EA29AB53, denovo1053 GI2.1_3353

JX222455, uncultured bacterium, subsurface aquifer sedimentEF999346, uncultured bacterium, Pearl River Estuary sediments at 6cm depth

Jn530486, uncultured organism, Guerrero Negro Hypersaline Mat ARB_E9534EC3, denovo248 GI2.1_3488

Am882642, uncultured delta proteobacterium, sediment from oil polluted water ARB_2869D04F, denovo1318 GII3.1_720

JQ245689, uncultured bacterium, mud volcanoDQ811807, uncultured delta proteobacterium, mangrove soil

JQ516448, uncultured Desulfobacteraceae bacterium, healthy tissueARB_350E391, denovo2371 GI2.1_3686

JQ739023, uncultured bacterium, Lonar sediment surface rocksAF177428, delta proteobacterium S2551

AB614135, Desulfatitalea tepidiphila, tidal flat sedimentJQ245513, uncultured bacterium, mud volcano

Desulfatitalea

AY222313, uncultured delta proteobacterium, host hindgut caecumDQ004679, uncultured bacterium, marine hydrocarbon seep sedimentARB_7FED19F4, denovo57 GII1.1_2044Jn662198, uncultured delta proteobacterium, tubeworm Ridgeia piscesae in low−flow environment

Jn442364, uncultured organism, Guerrero Negro Hypersaline Mat ARB_D6DE1887, denovo2591 GII2.1_5313

Jn451747, uncultured organism, Guerrero Negro Hypersaline Mat Jn456653, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_67DDF75D, denovo1296 GII1.1_1741Jn438994, uncultured organism, Guerrero Negro Hypersaline Mat

JN511301, uncultured organism, Guerrero Negro Hypersaline Mat Jn522159, uncultured organism, Guerrero Negro Hypersaline Mat Jn451253, uncultured organism, Guerrero Negro Hypersaline Mat

Jn453804, uncultured organism, Guerrero Negro Hypersaline Mat JN450823, uncultured organism, Guerrero Negro Hypersaline Mat

JN534382, uncultured organism, Guerrero Negro Hypersaline Mat Jn501880, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_1CA67942, denovo3200 GII1.1_9293Jn451179, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_3AE4005F, denovo2475 GII3.1_4668ARB_FF2383AD, denovo2345 GII3.1_2479

ARB_54481FDD, denovo2635 GII2.1_3097Jn459735, uncultured organism, Guerrero Negro Hypersaline Mat

Jn454132, uncultured organism, Guerrero Negro Hypersaline Mat Jn510800, uncultured organism, Guerrero Negro Hypersaline Mat

Fj712416, uncultured bacterium, Kazan Mud Volcano, Anaximander MountainsARB_4C2031DA, denovo1511 GI3.1_613

Jn522209, uncultured organism, Guerrero Negro Hypersaline MatJN452209, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_2EB5FAA4, denovo2738 GII3.1_6817Jn458684, uncultured organism, Guerrero Negro Hypersaline Mat

Jn459841, uncultured organism, Guerrero Negro Hypersaline Mat JN450773, uncultured organism, Guerrero Negro Hypersaline Mat

Jn442849, uncultured organism, Guerrero Negro Hypersaline Mat Jn450721, uncultured organism, Guerrero Negro Hypersaline Mat

JN460649, uncultured organism, Guerrero Negro Hypersaline Mat U45991, Desulfonema ishimotoniiU45992, Desulfonema ishimotoniiARB_A7DF0EB4, denovo3805 GII3.1_11577

JQ036269, uncultured bacterium, Eel River Basin sediment 2688ARB_1DFD6BEC, denovo952 GII1.1_2275

JN038818, uncultured delta proteobacterium, Chongxi wetland soilJX521572, uncultured bacterium, terrestrial sulfidic spring

Jn440228, uncultured organism, Guerrero Negro Hypersaline Mat ARB_D6FCA9F0, denovo1736 GII3.1_2524

ARB_85BA90CD, denovo2250 GI1.1_2008ARB_85BA90CD, denovo400 GII2.1_7007

U45990, Desulfonema limicolaARB_782235F9, denovo3920 GII2.1_7847

JQ215065, uncultured bacterium, bottlenose dolphin mouth

De

sulfo

ne

ma

ARB_14DE0EB5, denovo764 GII3.1_1091FJ814756, uncultured Desulfobacterales bacterium, Carmel Canyon

ARB_EB7F7EB9, denovo37 GII1.1_11048U45989, Desulfonema magnum

ARB_138F9ACC, denovo158 GII2.1_13064ARB_860FB93E, denovo844 GII3.1_3379

Jn485046, uncultured organism, Guerrero Negro Hypersaline Mat ARB_CD61D4ED, denovo1166 GII1.1_241ARB_CD61D4ED, denovo2718 GI1.1_256

ARB_82AEB832, denovo3766 GII3.1_7784Jn432947, uncultured organism, Guerrero Negro Hypersaline Mat Jn436445, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_86F3D39D, denovo1322 GII3.1_5031Am882643, uncultured delta proteobacterium, sediment from oil polluted water Jn470593, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_3847EEB8, denovo1986 GI1.1_4326ARB_385A98C2, denovo944 GII2.1_698

JX224716, uncultured bacterium, subsurface aquifer sedimentJX224719, uncultured bacterium, subsurface aquifer sediment

ARB_F6EFB277, denovo1388 GII3.1_8824ARB_1B95CE8A, denovo1016 GII3.1_5381

AM882627, uncultured delta proteobacterium, sediment from oil polluted water ARB_7BDF5E2E, denovo3523 GII3.1_3456

ARB_C4BDFDE5, denovo667 GI3.1_271ARB_9709B744, denovo2912 GII3.1_10066

ARB_B470EE51, denovo659 GII3.1_2742ARB_61518FF, denovo3305 GII3.1_9312

ARB_4E1BBC63, denovo1224 GI2.1_837JX223505, uncultured bacterium, subsurface aquifer sediment

AB237683, uncultured bacterium, deep subsurface groundwater from sedimentary rock milieuJX222865, uncultured bacterium, subsurface aquifer sediment

JX225750, uncultured bacterium, subsurface aquifer sedimentARB_8DA8F9E5, denovo433 GI2.1_3360

AB355075, uncultured bacterium, sediment, Manzallah LakeEf999366, uncultured bacterium, Pearl River Estuary sediments

DQ415752, uncultured bacterium, Frasassi sulfidic cave stream biofilmARB_7E558345, denovo924 GII3.1_3115

ARB_BC245D35, denovo1031 GII3.1_2710Fq659033, uncultured soil bacterium, PAH−contaminated soil

ARB_EA44E0C5, denovo1195 GI1.1_2747ARB_77B1D6BD, denovo1947 GII2.1_1107

ARB_F9FC6349, denovo1840 GII3.1_9728ARB_369D1854, denovo601 GI1.1_2578

FJ516891, uncultured Desulfobacteraceae bacterium, upper sedimentARB_D0E35DBE, denovo1637 GII2.1_5707

AB525467, uncultured bacterium, deep−sea methane seep sediment off Joetsu, Japan SeaJN454071, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_7CB39937, denovo3870 GII2.1_5347Jn486306, uncultured organism, Guerrero Negro Hypersaline Mat

Jn443588, uncultured organism, Guerrero Negro Hypersaline Mat Cu925954, uncultured bacterium, mesophilic anaerobic digester wastewater sludge

ARB_A049BE09, denovo1909 GII3.1_1493ARB_E0A335A9, denovo2058 GI2.1_966

AF154058, Eubostrichus dianae epibacterium 2_25DQ415749, uncultured bacterium, Frasassi sulfidic cave stream biofilm

EU246060, uncultured organism, hypersaline microbial matARB_8F55C3F4, denovo2751 GII2.1_6026

DQ415820, uncultured bacterium, Frasassi sulfidic cave stream biofilmDQ133938, uncultured bacterium, limestone−corroding stream biofilm

DQ415847, uncultured bacterium, Frasassi sulfidic cave stream biofilmJn537551, uncultured organism, Guerrero Negro Hypersaline Mat Jn512532, uncultured organism, Guerrero Negro Hypersaline Mat

ARB_42A2A09C, denovo2685 GII2.1_2390AM086154, uncultured bacterium, lake profundal sediment

DQ133936, uncultured bacterium, limestone−corroding stream biofilmDQ133911, uncultured bacterium, limestone−corroding stream biofilm

DQ133934, uncultured bacterium, limestone−corroding stream biofilmJn522187, uncultured organism, Guerrero Negro Hypersaline Mat ARB_988C1618, denovo3179 GII3.1_3268

AF099065, Desulfofrigus fragileARB_29F48A4E, denovo655 GI3.1_4246

HE600063, uncultured Desulfofrigus sp., marine sedimentARB_A5B339C9, denovo2168 GII2.1_7178

AF099064, Desulfofrigus oceanenseAB015240, uncultured delta proteobacterium, deepest cold−seep area of the Japan Trench

EU052250, uncultured delta proteobacterium, anode from microbial fuel cell fed with marine planktonEU052237, uncultured delta proteobacterium, anode from microbial fuel cell fed with marine plankton

AB530175, uncultured bacterium, marine sedimentARB_1EFDF7DA, denovo1912 GI1.1_3304

ARB_F38F9C78, denovo438 GII3.1_9848ARB_B3F8690C, denovo3171 GII2.1_2892

ARB_C3BF8B7, denovo3566 GII2.1_13654Jf344495, uncultured delta proteobacterium, hydrocarbon polluted marine sediments

ARB_972BEFDB, denovo2749 GI1.1_3156ARB_972BEFDB, denovo831 GII1.1_1216

ARB_EDD81177, denovo4017 GII2.1_4348ARB_824ACC52, denovo1977 GII3.1_7096ARB_CEB514CF, denovo327 GII3.1_2683

ARB_5499095F, denovo3967 GII2.1_121ARB_EA25DAB2, denovo301 GI1.1_5383

KC471130, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughKC471154, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

ARB_230591CA, denovo2332 GII1.1_5560KC471196, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

ARB_E3CF55F2, denovo3886 GII3.1_9557ARB_875F4162, denovo3943 GII3.1_12311

KC471054, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughHQ703812, uncultured bacterium, Lianyungang marine sediment

ARB_5366F7CC, denovo2113 GII3.1_12485KC471037, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughKC471011, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughKC471046, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

AF321820, Desulfofaba hanseniiJX222073, uncultured bacterium, subsurface aquifer sediment

Desulfofaba_1

ARB_42DA9ED8, denovo1976 GI1.1_1431ARB_FAF7585E, denovo2155 GII3.1_6836ARB_700BB89E, denovo2433 GI2.1_1599

ARB_B43CC987, denovo87 GII2.1_481Ay592706, uncultured bacterium, Napoli mud volcano, Eastern Mediterranean,

ARB_6CB546A4, denovo3222 GII3.1_3868AY268891, Desulfofaba fastidiosa

EU487968, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bed

Desulfofaba_2

JX222769, uncultured bacterium, subsurface aquifer sedimentDQ415842, uncultured bacterium, Frasassi sulfidic cave stream biofilm

ARB_7A94311A, denovo1579 GII3.1_3316ARB_102BDE9A, denovo1602 GI2.1_1649

HE613444, Desulfatiferula sp. BE2801, estuary sedimentFR667781, uncultured bacterium, iron snow from acidic coal mining−associated Lake 77

Jf344222, uncultured delta proteobacterium, petroleum spot over marine sediments DQ826724, Desulfatiferula olefinivorans

ARB_E11A1166, denovo399 GI2.1_871JX222822, uncultured bacterium, subsurface aquifer sediment

JX223240, uncultured bacterium, subsurface aquifer sedimentARB_2CD45AD4, denovo2141 GI2.1_1838

JX222458, uncultured bacterium, subsurface aquifer sedimentJX223512, uncultured bacterium, subsurface aquifer sediment

JX457247, uncultured bacterium, gutGQ866071, uncultured bacterium, gut

ARB_3A4D20D7, denovo2746 GII3.1_2705

Desulfatiferula

Figure S5b: Bacterial 16S rRNA gene amplicon sequences assigned to in the class of Deltaproteobac-teria, order: Desulfobacterales, family: Desulfobacteraceae (black bold: sequences retrieved in this study and closest relatives).

Page 147: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

139

Supplementary Material

Jn452561, uncultured organism, Guerrero Negro Hypersaline Mat ARB_5E38B580, denovo3953 GII1.1_1602

JX391346, uncultured bacterium, marine sedimentKf623992, uncultured bacterium, Janssand tidal flat, Wadden Sea

KF799406, uncultured bacterium, gutKF268874, uncultured bacterium, marine sediment

JX391709, uncultured bacterium, marine sedimentJX391368, uncultured bacterium, marine sediment

FJ949149, uncultured Desulfobacteraceae bacterium, calcareous sandy sedimentJX391090, uncultured bacterium, marine sedimentARB_9E2E722F, denovo2789 GII1.1_7427

JX391589, uncultured bacterium, marine sedimentFj716993, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandARB_9D945629, denovo867 GII1.1_234

U51844, delta proteobacterium WNARB_B6C76362, denovo522 GII3.1_7062

ARB_8AA8018, denovo3096 GII2.1_5852JQ212736, uncultured bacterium, bottlenose dolphin mouth

Fj717001, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandAF099054, Desulfobacter sp. ASv25GQ261762, uncultured delta proteobacterium, deep sea sediment associated with whale falls

GQ356938, uncultured bacterium, methane seep sedimentGQ356949, uncultured bacterium, methane seep sedimentFR733673, Desulfobacter hydrogenophilus

DQ057079, Desulfobacter psychrotoleransFJ717003, uncultured bacterium, marine sediment from Cullercoats, Northumberland

Fj716965, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandJF343993, uncultured delta proteobacterium, hydrocarbon polluted marine sediments from rodas beach

Fj716962, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandU85469, Desulfobacter sp. BG8KF268856, uncultured bacterium, marine sediment

ARB_DD288351, denovo2652 GI2.1_4358JX225079, uncultured bacterium, subsurface aquifer sediment

ARB_7BB6FB78, denovo2419 GI3.1_4466ARB_A58E3C3B, denovo3178 GII3.1_10337

JX120461, uncultured bacterium, subsurface aquifer sedimentJX120460, uncultured bacterium, subsurface aquifer sediment

JX125449, uncultured bacterium, subsurface aquifer sedimentJX120466, uncultured bacterium, subsurface aquifer sediment

AF418180, Desulfobacter postgateiM26633, Desulfobacter postgateiAGJR02000005, Desulfobacter postgatei 2ac9, anaerobic sediment of brackish water ditch near Jadebusen

Eu864440, uncultured bacterium, Xiao River EU864484, uncultured bacterium, Xiao River receiving treated oxytetracycline production wastewater

JX120459, uncultured bacterium, subsurface aquifer sedimentJX225700, uncultured bacterium, subsurface aquifer sediment

Jx271977, uncultured bacterium, activated sludge in lab−scale reactor EU234252, uncultured bacterium, downstream of Wang Yang River, receiving penicillin G production wastewater

AGJR02000008, Desulfobacter postgatei 2ac9, anaerobic sediment of brackish water ditch near JadebusenFJ716310, uncultured bacterium, Sawmill Sink water column, at 10.3 m water depth

KF933845, Desulfobacula sp. TS−7ARB_68E1E8F6, denovo1052 GII3.1_1606

AY222309, uncultured delta proteobacterium, host hindgut caecumARB_55128A66, denovo3319 GII2.1_11393

EF442984, sulfate−reducing bacterium JHA1AY177795, uncultured Desulfobacula sp., Antarctic sediment

FR823371, uncultured delta proteobacterium, gas seep sedimentJF514280, uncultured bacterium, sea

ARB_691E454A, denovo157 GI2.1_919ARB_BCB1D27F, denovo1223 GII3.1_11648

EU290701, uncultured bacterium, sediment 0−2cm depthAY177800, uncultured Desulfobacula sp., Antarctic sedimentKF268860, uncultured bacterium, marine sediment

JQ753193, uncultured bacterium, Antarctic sea iceGU197423, uncultured bacterium, intertidal sediment, depth 0−15 cmAy922192, uncultured delta proteobacterium, grey whale bone, Pacific Ocean, Santa Cruz Basin

FR823375, uncultured delta proteobacterium, associated sedimentFR823376, uncultured delta proteobacterium, associated sediment

JF344268, uncultured delta proteobacterium, petroleum spot over marine sediments from figueiras beachKC471161, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughKC471246, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

KC471078, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughARB_9E645E6E, denovo210 GII3.1_1506

ARB_AD101018, denovo343 GII2.1_4906ARB_9900558E, denovo1638 GII2.1_10513

ARB_ADF00A92, denovo796 GII3.1_7275KC471244, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa Trough

KC471084, uncultured bacterium, marine sediment at Yonaguni Knoll in the Okinawa TroughARB_4A9FE3FF, denovo1772 GII2.1_879

ARB_2423619F, denovo2566 GI2.1_305ARB_18F285C2, denovo2618 GII2.1_3476

ARB_E1D7A46E, denovo1731 GII1.1_2949KC492863, uncultured Desulfobacula sp., Baltic Sea redoxcline, 119 m depthAY177803, uncultured Desulfobacula sp., Antarctic sediment

AJ937698, uncultured bacterium, mud from a terrestrial volcanoARB_A9ED3C90, denovo127 GII3.1_1097

ARB_C5D21D56, denovo1906 GII3.1_607ARB_752C5DFE, denovo3090 GII3.1_10845

ARB_56BD7EAD, denovo3258 GII2.1_9782Fj628185, uncultured bacterium, brackish water from anoxic fjord

ARB_C5116C1C, denovo1634 GII2.1_4094ARB_A9D2324B, denovo1307 GI1.1_2919ARB_7BB990DB, denovo1697 GII3.1_8393

FJ628265, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 22GU145490, uncultured bacterium, Black Sea, potential density 15.81

KC545723, uncultured Desulfobacula sp., oxycline region of a coastal marine water columnKC545728, uncultured Desulfobacula sp., oxycline region of a coastal marine water column

AM774314, delta proteobacterium JS_SRB50Hy, tidal flat sedimentAJ237606, Desulfobacula phenolicaGQ261767, uncultured delta proteobacterium, deep sea sediment associated with whale fallsGQ348356, uncultured delta proteobacterium, Saanich Inlet, 200 m depthGU197424, uncultured bacterium, intertidal sediment, depth 0−15 cm

DQ831540, uncultured Desulfobacula sp., marine sedimentARB_CB393774, denovo2285 GII1.1_285

ARB_831C7D25, denovo1192 GII1.1_11241ARB_F227D1D4, denovo344 GII3.1_2379

ARB_FE2B548D, denovo2584 GI2.1_237ARB_F4D910C0, denovo2876 GII2.1_5307

Kc000940, unidentified marine bacterioplankton, surface seawater samples from South China SeaARB_8741E9EF, denovo551 GI1.1_5282

ARB_6AAFAFB, denovo1020 GII1.1_1771ARB_E6C2968D, denovo1222 GI1.1_1092

HQ405662, bacterium enrichment culture clone AOM−SR−B24, enrichment mediating the anaerobic oxidation of methaneARB_8921930, denovo1642 GI3.1_5066

EU570896, uncultured bacterium, Nitinat Lake at a depth of 13 mARB_FE450D44, denovo2673 GII2.1_270

Dq925890, uncultured bacterium, hydrothermal vent Guaymas BasinAB189365, uncultured Desulfobacterium sp., cold seep sediment

AJ704693, uncultured delta proteobacterium, marine sedimentAJ704685, uncultured delta proteobacterium, marine sediment

ARB_7FCB72A1, denovo1108 GI1.1_3343ARB_844A3AFA, denovo994 GII3.1_5707

JF268367, uncultured bacterium, Hikurangi margin, Wairarapa − Takahae, station 309; SO191_3_309KF624105, uncultured bacterium, anoxic pyrite−particle colonization experiment, Janssand tidal flat, Wadden Sea

JF344167, uncultured delta proteobacterium, petroleum spot over marine sediments from figueiras beachFN820309, uncultured bacterium, marine sediment, mud volcanoJQ580395, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oil

ARB_F3E53A99, denovo2676 GII3.1_2047AY197397, uncultured bacterium, Guaymas Basin hydrothermal vent sediments

ARB_B802BE9, denovo1369 GII3.1_343FJ497571, uncultured delta proteobacterium, Vailulu’u Seamount

ARB_4BCB0690, denovo2794 GII1.1_6932JX225725, uncultured bacterium, subsurface aquifer sediment

JX224838, uncultured bacterium, subsurface aquifer sedimentARB_38687E91, denovo1775 GII3.1_357

JX224531, uncultured bacterium, subsurface aquifer sedimentKF059938, uncultured Desulfobacula sp., oil well water

JX224964, uncultured bacterium, subsurface aquifer sedimentARB_90F03020, denovo418 GI3.1_3548

Desu

lfobacu

la

ARB_903AB4BE, denovo400 GI1.1_2096ARB_E811D690, denovo2159 GII3.1_4428

ARB_6E68406, denovo3323 GII3.1_11303ARB_2658D044, denovo736 GI2.1_1998

Ef687285, uncultured bacterium, sediment underneath a sulfide−oxidizing mat, Chefren mud volcanoARB_5E2AC7C5, denovo2114 GII1.1_9836

ARB_B40486E4, denovo3407 GII1.1_957ARB_BDBAEB51, denovo2557 GII2.1_2691

ARB_7F4AC00D, denovo2965 GII2.1_4835ARB_5AFBF727, denovo3032 GII3.1_1347

ARB_99C327D9, denovo3509 GII2.1_3753ARB_F6FC891D, denovo2989 GII3.1_6268

HQ405610, bacterium enrichment culture clone AOM−SR−B20, enrichment mediating the anaerobic oxidation of methaneEf687314, uncultured bacterium, sediment underneath a sulfide−oxidizing mat, Chefren mud volcano, Nile Deep Sea Fan

FR823370, uncultured delta proteobacterium, gas seep sedimentAM883182, uncultured bacterium, host trophosome tissue

AY177799, uncultured Desulfobacteraceae bacterium, Antarctic sedimentJn455195, uncultured organism, Guerrero Negro Hypersaline Mat

EU570886, uncultured bacterium, Nitinat Lake at a depth of 50 mARB_1852A786, denovo4080 GII1.1_371

ARB_E77D75, denovo2622 GII3.1_3718ARB_138286CE, denovo2476 GI2.1_431ARB_138286CE, denovo3966 GII3.1_12408

ARB_12774C80, denovo2714 GII2.1_11932AY197376, uncultured proteobacterium, Guaymas Basin hydrothermal vent sediments

DQ925877, uncultured bacterium, hydrothermal vent chimneys of Guaymas BasinARB_6BE57BCD, denovo466 GI2.1_3816

ARB_51E6D082, denovo2560 GII1.1_904ARB_4D98103A, denovo590 GI1.1_2711

ARB_F2CADC85, denovo1645 GII2.1_320ARB_8DA9F1F2, denovo2813 GI3.1_4146

AY197389, uncultured proteobacterium, Guaymas Basin hydrothermal vent sedimentsCP001087, Desulfobacterium autotrophicum HRM2AF099056, Desulfobacterium sp. LSv25GQ261805, uncultured delta proteobacterium, deep sea sediment associated with whale falls

ARB_85D093E0, denovo3220 GII1.1_3563U85474, Desulfobacterium sp. BG33U85471, Desulfobacterium sp. BG18

FR733668, Desulfococcus niaciniEF442983, Desulfobacterium zeppelinii

AF418178, Desulfobacterium vacuolatumARB_DCBB76E5, denovo3877 GII1.1_2385ARB_3CD31A93, denovo1687 GI3.1_2927

ARB_42406D1, denovo1081 GI3.1_4342

Desu

lfobact

erium

_2

ARB_F4D45108, denovo1094 GII1.1_537ARB_F23C63C1, denovo2424 GII2.1_6963

JN252194, Candidatus Desulfamplus magnetomortis BW−1EU047537, uncultured sulfate−reducing bacterium, benzene−degrading, sulfate−reducing enrichment culture

Candidatus Desulfamplus

ARB_D6D2D42B, denovo157 GII3.1_9438ARB_45B9A382, denovo999 GII3.1_1990AB530188, uncultured bacterium, marine sediment

ARB_B33F233E, denovo1250 GII3.1_11390ARB_3811BEBC, denovo272 GI2.1_266ARB_B99FC159, denovo735 GII2.1_4119

ARB_217ED137, denovo1265 GII3.1_6977ARB_6F53D9A2, denovo941 GII2.1_6606

ARB_6CFF573D, denovo1362 GII1.1_9507ARB_6CFF573D, denovo289 GI2.1_1883

ARB_C7AE2C97, denovo2652 GII3.1_3409ARB_7D646BD3, denovo593 GI2.1_647

ARB_4AFD410D, denovo1138 GI2.1_2087ARB_12CB3B73, denovo3860 GII3.1_10456

GQ348489, uncultured delta proteobacterium, Saanich Inlet, 200 m depthARB_1BD1B139, denovo185 GII1.1_842

ARB_AE84C44F, denovo827 GII2.1_8140EU885094, uncultured bacterium, GI tract

JQ216349, uncultured bacterium, bottlenose dolphin mouthARB_843ECF4E, denovo2893 GII3.1_8486

JX391413, uncultured bacterium, marine sedimentARB_BE36ACE0, denovo3825 GII3.1_3441

ARB_DFC8E39B, denovo3243 GII2.1_7317KF268778, uncultured bacterium, marine sediment

ARB_9BF7C65, denovo232 GI3.1_3136GQ259286, uncultured bacterium, deep sediment

ARB_F7DCEB29, denovo227 GII2.1_2296ARB_14204A90, denovo1504 GII2.1_6800ARB_13808028, denovo2840 GI3.1_827

GU127076, uncultured bacterium, sediment; anoxic zone; tucurui hydroeletric power plant reservoirARB_51EF3CEE, denovo718 GII3.1_4007

HM243869, uncultured bacterium, middle sediment from Honghu LakeFJ485347, uncultured delta proteobacterium, wall biomat sample in El Zacaton at 273m depth

JQ580417, uncultured delta proteobacterium, sediments from Rodas Beach polluted with crude oilARB_6F8CC735, denovo624 GII2.1_730

KF548223, uncultured bacterium, aerobic tank of a full−scale produced water treatment plantARB_57B3B254, denovo617 GI1.1_920

AB530187, uncultured bacterium, marine sedimentARB_BFBB7707, denovo738 GII3.1_701

Figure S5c: Bacterial 16S rRNA gene amplicon sequences assigned to in the class of Deltaproteobac-teria, order: Desulfobacterales, family: Desulfobacteraceae (black bold: sequences retrieved in this study and closest relatives).

Page 148: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

140

Ay280427, uncultured epsilon proteobacterium, hydrothermal vent sulfide chimney, Juan de Fuca RidgeJn662023, uncultured epsilon proteobacterium, ubeworm Ridgeia piscesae

AB278151, uncultured bacterium, obtained from Mariana TroughAY531599, uncultured bacterium, Indian Ocean hydrothermal vent

JQ712438, uncultured epsilon proteobacterium, hydrothermal vent fluidsJn662026, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae AJ441211, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretionsAJ576002, uncultured epsilon proteobacterium, deep sea hydrothermal vent

Jn874168, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJn873951, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

AY672507, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanAY672537, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

EU555121, uncultured bacterium, Dudley hydrothermal ventAF449232, uncultured epsilon proteobacterium, Riftia pachyptila’s tube

JQ611095, uncultured bacterium, biofilm near white vent off Kueishan IslandJQ611085, uncultured bacterium, biofilm near white vent off Kueishan Island

Ay280407, uncultured epsilon proteobacterium, hydrothermal vent sulfide chimney, Juan de Fuca RidgeARB_64F4AD65, denovo915 GI2.1_384

ARB_305B8363, denovo513 GII1.1_682Jn662076, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae

ARB_EA631019, denovo1504 GI2.1_3731AY672526, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

AY672530, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanAY672521, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

AY672520, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanAJ441209, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretions

Gu369886, uncultured epsilon proteobacterium, shallow hydrothermal vent region

uncultured 32

AF449253, uncultured epsilon proteobacterium, Riftia pachyptila’s tube

uncultured 37

AJ431222, proteobacterium BHI80−16AF420345, uncultured proteobacterium

Jn874159, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJQ712462, uncultured epsilon proteobacterium, hydrothermal vent fluids

Jn662191, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae AP009179, Sulfurovum sp. NBC37−1AP009179, Sulfurovum sp. NBC37−1AP009179, Sulfurovum sp. NBC37−1

JQ611086, uncultured bacterium, biofilm near white vent off Kueishan IslandJQ611159, uncultured bacterium, core sediment at 0−3 cm depth in yellow vent off Kueishan Island

AJ431221, proteobacterium BHI60−14AB091292, Sulfurovum lithotrophicum, deep−sea hydrothermal vent environmentsJn873995, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

FR853000, uncultured bacterium, active hydrothermal chimney collected from the Dudley site of Juan de Fuca RidgeJQ712487, uncultured epsilon proteobacterium, hydrothermal vent fluidsJn662247, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae

uncultured (9 sequences)

AY531576, uncultured bacterium, Indian Ocean hydrothermal ventAF449235, uncultured epsilon proteobacterium, Riftia pachyptila’s tube

Jn874207, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific RiseJn874109, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

Jn874231, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific RiseAY672533, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanJQ611145, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

JQ611112, uncultured bacterium, biofilm near yellow vent off Kueishan IslandHq153866, uncultured bacterium, white microbial film shallow hydrothermal vent

Jn662036, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae Jn662064, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae

FJ535344, uncultured bacterium, sedimentFJ535271, uncultured bacterium, iron matJn662133, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae

AF449231, uncultured epsilon proteobacterium, Riftia pachyptila’s tubeJQ611132, uncultured bacterium, core sediment at 0−5 cm depth in white vent off Kueishan Island

JQ611123, uncultured bacterium, core sediment at 21−24 cm depth in white vent off Kueishan IslandJQ712402, uncultured epsilon proteobacterium, hydrothermal vent fluids

AB842288, uncultured bacterium, gut of S.crosnieri

182

JF747909, uncultured bacterium, Frasassi cave system sulfidic springJF747989, uncultured bacterium, Frasassi cave system sulfidic spring

GQ355039, uncultured bacterium, springEF444740, uncultured bacterium, Thermopiles hot springs

Dq295592, uncultured epsilon proteobacterium, microbial mat from sulfidic spring Dq295601, uncultured epsilon proteobacterium, microbial mat from sulfidic spring Dq295602, uncultured epsilon proteobacterium, microbial mat from sulfidic spring

AF121887, bacterium 2BP−7GQ227834, uncultured bacterium, ground swab in mangrove forest

GQ355011, uncultured epsilon proteobacterium, springJn662280, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae

EF061978, uncultured epsilon proteobacterium, mangrove sedimentGU583971, uncultured bacterium, mangrove soilGU583970, uncultured bacterium, mangrove soilJX403047, uncultured bacterium, bitumen

GU180159, uncultured bacterium, sediment and soil slurryGU180158, uncultured bacterium, sediment and soil slurry

EU488101, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bedKC009865, uncultured Campylobacteraceae bacterium, shallow fluidized muds off the French Guiana coast

KF465256, uncultured Sulfurovum sp., Alang−Sosiya ship breaking yardARB_5E9258CC, denovo3386 GII3.1_7847

ARB_482679F6, denovo1298 GII1.1_1746KC763867, uncultured bacterium, oil−contaminated soilGU120550, uncultured epsilon proteobacterium, inside erupting mud volcano, temperature = 27.7 C

GU056082, uncultured bacterium, soil sample above gas and oil fieldDQ394915, uncultured epsilon proteobacterium, harbor sediment

JQ989773, uncultured bacterium, subseafloor sediment at the KP−9EF061977, uncultured epsilon proteobacterium, mangrove sediment

JQ989775, uncultured bacterium, subseafloor sediment at the KP−9KF758699, uncultured Sulfurovum sp., ocean waterKF758616, uncultured Sulfurovum sp., ocean water

KF758740, uncultured Sulfurovum sp., ocean waterJQ989774, uncultured bacterium, subseafloor sediment at the KP−9KF758756, uncultured Sulfurovum sp., ocean water

KF758605, uncultured Sulfurovum sp., ocean waterJQ989767, uncultured bacterium, subseafloor sediment at the KP−9

KF758583, uncultured Sulfurovum sp., ocean waterJQ989770, uncultured bacterium, subseafloor sediment at the KP−9

AF420360, uncultured proteobacteriumJX403027, uncultured bacterium, underground waterEF467546, uncultured bacterium, Frasassi sulfidic cave stream biofilm

Jn662200, uncultured epsilon proteobacterium, Ridgeia piscesae tubewormAF420343, uncultured proteobacterium

120

uncultured

uncultured

8

JQ036280, uncultured bacterium, Hydrate Ridge sediment 2518

uncultured 11

JQ611092, uncultured bacterium, biofilm near white vent off Kueishan IslandAJ441206, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretions

EU570833, uncultured bacterium, Nitinat Lake at a depth of 37.5 mKc545718, uncultured Desulfopila sp., oxycline region

GQ354918, uncultured Desulfobacterales bacterium, springGQ354989, uncultured delta proteobacterium, spring

Jn662041, uncultured proteobacterium, tubeworm Ridgeia piscesae Jn662234, uncultured delta proteobacterium, tubeworm Ridgeia piscesae

ARB_870C0E9A, denovo896 GII2.1_6517ARB_D61337BA, denovo3327 GII3.1_2152

GU120555, uncultured epsilon proteobacterium, inside erupting mud volcanoAB530185, uncultured bacterium, marine sediment

KF758719, uncultured Salimesophilobacter sp., ocean waterKF758696, uncultured Sulfurovum sp., ocean water

DQ394916, uncultured bacterium, harbor sedimentGQ354985, uncultured epsilon proteobacterium, spring

Fn554089, uncultured sediment bacterium, Logatchev hydrothermal vent fieldFJ615414, uncultured bacterium, high rate anaerobic methane−oxidizing submerged membrane bioreactor

HM445989, uncultured bacterium, wastewaterEU234154, uncultured bacterium, effluent of the WWTP for treating penicillin G production wastewater

AJ431220, proteobacterium Dex80−27FJ535262, uncultured bacterium, iron mat

GU220732, uncultured bacterium, low temperature hydrothermal precipitates of East Lau Spreading CenterFR852993, uncultured bacterium, active hydrothermal chimney collected from the Dudley site of Juan de Fuca RidgeFR853069, uncultured bacterium, active hydrothermal chimney collected from the Dudley site of Juan de Fuca Ridge

KF624533, uncultured bacterium, suboxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaAB842287, uncultured bacterium, gut of S.crosnieri

KF545018, uncultured Sulfurovum sp., deep−sea sedimentJn662077, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae in high−flow environment

Ay280435, uncultured epsilon proteobacterium, exterior of a deep−sea hydrothermal vent sulfide chimney, Juan de Fuca RidgeJn662174, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae in high−flow environment

Jn874024, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJn874212, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific Rise

Jn662066, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae in high−flow environmentJn662253, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae in high−flow environmentJn662149, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae in high−flow environmentJQ712481, uncultured epsilon proteobacterium, hydrothermal vent fluids

Jn662098, uncultured epsilon proteobacterium, vestimentiferan tubeworm Ridgeia piscesae in high−flow environmentJQ712460, uncultured epsilon proteobacterium, hydrothermal vent fluids

JQ712484, uncultured epsilon proteobacterium, hydrothermal vent fluidsJQ712444, uncultured epsilon proteobacterium, hydrothermal vent fluidsAJ431224, proteobacterium Dex80−90

JN874140, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NJN873965, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’N

JN874006, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NAY672518, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

JN873957, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NJN874072, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NJN874274, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific Rise, 9 degrees 50’N

JN873975, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NAQWF01000016, Sulfurovum sp. SCGC AAA036−F05

JN873991, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NAY672522, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanJN873979, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’N

AY672531, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

uncultured 25

AB842280, uncultured bacterium, gut of S.crosnieriAy280389, uncultured epsilon proteobacterium, exterior of a deep−sea hydrothermal vent sulfide chimney, Juan de Fuca Ridge

JX521118, uncultured bacterium, terrestrial sulfidic springJQ611148, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

JQ611152, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandJQ611146, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandJQ611125, uncultured bacterium, core sediment at 21−24 cm depth in white vent off Kueishan Island

JQ611156, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandJQ611139, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

JQ611117, uncultured bacterium, core sediment at 21−24 cm depth in white vent off Kueishan IslandJQ611179, uncultured bacterium, porewater fluid near yellow vent off Kueishan Island

JQ611147, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandJQ611151, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

JQ611137, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandAY672540, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

Cam

pyl

obact

era

les_

Helic

obact

era

ceae_Sulfu

rovu

m

uncultured

Figure S6a: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Epsilonproteobac-teria, order: Campylobacterales, family: Helicobacteraceae, genus: Sulfurovum (black bold: sequences retrieved in this study and closest relatives).

Page 149: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

141

Supplementary Material

uncultured 16

JX570598, endosymbiont of Ridgeia piscesaeJQ712409, uncultured epsilon proteobacterium, hydrothermal vent fluids

AF449239, uncultured epsilon proteobacterium, Riftia pachyptila’s tubeARB_A468E98B, denovo1651 GII3.1_10383

JQ199082, uncultured bacterium, seawater; next to dolphin UGQ357014, uncultured bacterium, methane seep sediment

JX016315, uncultured bacterium, marine bulk waterHM057704, uncultured bacterium, ocean water from the Yellow SeaHQ203962, uncultured bacterium, seawater

ARB_BDD1DF7F, denovo1949 GI1.1_1153FR666865, uncultured bacterium, microbial matFR666868, uncultured bacterium, microbial mat

FR666860, uncultured bacterium, microbial matFR666861, uncultured bacterium, microbial mat

FR666863, uncultured bacterium, microbial matFR666862, uncultured bacterium, microbial mat

JQ712383, uncultured epsilon proteobacterium, hydrothermal vent fluidsARB_7969B766, denovo561 GI2.1_1182FR666864, uncultured bacterium, microbial matGq261829, uncultured epsilon proteobacterium, deep sea sediment whale falls

FR666869, uncultured bacterium, microbial matGu061298, uncultured epsilon proteobacterium, Yellow Sea intertidal beach

FR683719, uncultured marine bacterium, marine biome, fjord, coastal waterHe576783, uncultured epsilon proteobacterium, biorector inoculated microbial mat

KF799694, uncultured bacterium, gutKF799695, uncultured bacterium, gutKF799571, uncultured bacterium, gutARB_B31CB100, denovo3137 GII1.1_1105

AY548997, uncultured bacterium, marine environmentFn429802, uncultured epsilon proteobacterium, marine sediments (1.1 mbsf)

Jf268371, uncultured bacterium, Hikurangi margin, Wairarapa − TakahaeJQ712406, uncultured epsilon proteobacterium, hydrothermal vent fluidsAB189374, uncultured epsilon proteobacterium, cold seep sediment

JQ712417, uncultured epsilon proteobacterium, hydrothermal vent fluidsJQ712419, uncultured epsilon proteobacterium, hydrothermal vent fluids

JQ712382, uncultured epsilon proteobacterium, hydrothermal vent fluids

uncultured 6

JX521734, uncultured bacterium, terrestrial sulfidic spring

uncultured 4

uncultured

JX521782, uncultured bacterium, terrestrial sulfidic springAy922188, uncultured epsilon proteobacterium, grey whale bone

uncultured 22

uncultured 9

Gu061286, uncultured epsilon proteobacterium, Yellow Sea intertidal beach ARB_F642CBD4, denovo370 GI2.1_670Gu061291, uncultured epsilon proteobacterium, Yellow Sea intertidal beach seawater

EU050947, uncultured bacterium, sediment from the Kings Bay, Svalbard, ArcticKF799046, uncultured bacterium, gut

Eu573099, bacterium enrichment culture clone EB24.3, North Sea, Ekofisk oil field JQ862033, uncultured epsilon proteobacterium

JX391668, uncultured bacterium, marine sedimentJX391694, uncultured bacterium, marine sediment

EF999357, uncultured bacterium, Pearl River Estuary sediments at 6cm depthJN018454, uncultured bacterium, Gulf of Mexico stable isotope probing incubationsJn233497, uncultured marine bacterium, ocean water Cape Lookout, NC from 505m depth

KC682345, uncultured bacterium, substrate mineral chipsKC682364, uncultured bacterium, substrate mineral chips

AB015529, uncultured epsilon proteobacterium, deep−sea sediments from different depthsJn233480, uncultured marine bacterium, ocean water continental slope off Cape Lookout

ARB_87561165, denovo588 GI3.1_1443AJ810558, uncultured epsilon proteobacterium, superficial sediments

uncultured 4

Dq357825, epsilon proteobacterium Oy−M7, adult oyster mantleEF645952, uncultured bacterium, Namibian upwelling systemEF028937, uncultured bacterium, Zostera marina roots

GU584643, uncultured bacterium, marine hydrocarbon seep sedimentFj717168, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717163, uncultured bacterium, marine sediment from Cullercoats, Northumberland

JQ586324, uncultured bacterium, arctic marine sedimentARB_21BC0A68, denovo2795 GI2.1_5638EU142059, uncultured bacterium, methane seep sediment

EU236320, uncultured bacterium, 13 m depth in the Pacific Ocean

uncultured 50

uncultured 6

L42994, Arcobacter sp. Solar Lake, cyanobacterial mat from hypersaline lakeEF031096, uncultured bacterium, endorheic hypersaline atalasohaline lake

EF031092, uncultured bacterium, endorheic hypersaline atalasohaline lakeEU512920, Arcobacter marinus, seawater associated with starfish

JX403054, uncultured bacterium, bitumenFN356234, uncultured bacterium, oil production waterEu573100, bacterium enrichment culture clone EB27.1, North Sea, Ekofisk oil field

FR675876, Arcobacter molluscorumFR675875, Arcobacter molluscorumFR675874, Arcobacter molluscorum CECT 7696

AB893930, uncultured bacterium, digestive tractJN596130, bacterium A1(2011)JN596128, bacterium A2(2011b)

ARB_18DB62B1, denovo487 GI2.1_2840ARB_3F2E08E3, denovo592 GII2.1_8864

EU669904, Arcobacter mytili, musselsHG792246, uncultured bacterium, intestinal contents of Chinese Mitten Crab

AF513455, Arcobacter halophilusKF268902, uncultured bacterium, marine sedimentFJ202783, uncultured bacterium, Montastraea faveolata kept in aquarium for 23 days

uncultured 7

KC527432, uncultured bacterium, White Plague Disease infected coral

uncultured 25

uncultured 4

AJ271655, Arcobacter sp. A3b2EU142043, uncultured bacterium, methane seep sediment

AJ271654, Arcobacter sp. D1a1

uncultured 9

FR683715, uncultured marine bacterium, marine biome, fjord, coastal waterAJ810559, uncultured epsilon proteobacterium, superficial sediments

ARB_DD34D980, denovo2326 GII2.1_3882Fj717102, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717122, uncultured bacterium, marine sediment from Cullercoats, Northumberland

HQ203797, uncultured bacterium, seawaterEU052239, uncultured epsilon proteobacterium, anode from microbial fuel cell fed with marine plankton

GU584648, uncultured bacterium, marine hydrocarbon seep sediment

uncultured 15

uncultured 26

AB530236, uncultured bacterium, marine sedimentFR666867, uncultured bacterium, microbial mat

Jf344171, uncultured epsilon proteobacterium, petroleum spot over marine sediments EU245155, uncultured organism, hypersaline microbial mat

EF123613, uncultured epsilon proteobacterium, black band diseased (BBD) coral tissuesJX570599, endosymbiont of Ridgeia piscesae

DQ218324, uncultured Arcobacter sp., microbial mat

uncultured 15

FJ628351, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 28FR666871, uncultured bacterium, microbial matARB_2017809, denovo3313 GII3.1_10544

HM057719, uncultured bacterium, ocean water from the Yellow SeaCP001999, Arcobacter nitrofigilis DSM 7299, rootsL14627, Arcobacter nitrofigilis

CP001999, Arcobacter nitrofigilis DSM 7299, rootsCP001999, Arcobacter nitrofigilis DSM 7299, rootsCP001999, Arcobacter nitrofigilis DSM 7299, rootsEu106661, Arcobacter nitrofigilis

JX403044, uncultured bacterium, bitumenJF904743, uncultured Arcobacter sp., mangrove soil

EF648116, uncultured Arcobacter sp., aerobic activated sludgeHQ203800, uncultured bacterium, seawater

JN118552, Arcobacter sp. sw028, a−surface water of South PacificHM057804, uncultured epsilon proteobacterium, ocean water from the Yellow Sea

HM057749, uncultured epsilon proteobacterium, ocean water from the Yellow SeaHM057779, uncultured epsilon proteobacterium, ocean water from the Yellow SeaHQ203831, uncultured bacterium, seawaterJQ712049, uncultured bacterium, surface seawater offshore QingdaoJQ712030, uncultured bacterium, surface seawater offshore Qingdao

JQ712046, uncultured bacterium, surface seawater offshore QingdaoHM057765, uncultured epsilon proteobacterium, ocean water from the Yellow SeaJQ712055, uncultured bacterium, surface seawater offshore Qingdao

AM084114, Arcobacter sp. R−28314JQ845784, uncultured Arcobacter sp., wash water from wash basin 2 of a spinach−processing plant

HE565359, Arcobacter venerupis, SeaARB_AD7A5D23, denovo709 GI1.1_4652

JQ845782, uncultured Arcobacter sp., wash water from wash basin 2 of a spinach−processing plantDQ234254, uncultured Arcobacter sp., mangroveHM057679, uncultured epsilon proteobacterium, ocean water from the Yellow SeaFJ946551, uncultured epsilon proteobacterium, arctic meltwaterJX403059, uncultured bacterium, bitumen

HQ860689, uncultured bacterium, stream waterFJ612319, uncultured bacterium, lake water

uncultured 44

JQ845771, uncultured Arcobacter sp., wash water from wash basin 2 of a spinach−processing plant

group

DQ234101, uncultured Arcobacter sp., mangroveFN428747, uncultured bacterium, Mahananda river waterFR691497, uncultured bacterium, Mahananda river sediment

HQ860691, uncultured bacterium, stream water

uncultured 21

mainly uncultured including Arcobacter butzleri 26

JX912503, Arcobacter butzleriAM084124, Arcobacter sp. R−28214JQ086899, uncultured Arcobacter sp., push core sediment sample from the vadose zone of a hydrocarbon contaminated aquifer

JF497852, uncultured bacterium, activated sludgeJN397809, uncultured bacterium, river bank

Z93999, uncultured bacterium

uncultured 44

JF915357, Arcobacter cryaerophilusHQ860539, uncultured bacterium, stream waterKF703990, Arcobacter sp. 69948, American crow feces

EU560811, uncultured bacterium, intestinal floraAF324539, uncultured epsilon proteobacterium BioIuz K34

AB355029, uncultured bacterium, surface water, Manzallah LakeCU925313, uncultured bacterium, mesophilic anaerobic digester which treats municipal wastewater sludge

CU926906, uncultured bacterium, mesophilic anaerobic digester which treats municipal wastewater sludgeCU924753, uncultured bacterium, mesophilic anaerobic digester which treats municipal wastewater sludge

uncultured 62

JQ213140, uncultured bacterium, bottlenose dolphin mouthFJ959725, uncultured bacterium, blowhole

uncultured 20

uncultured 19

uncultured 4

JQ611225, uncultured bacterium, venting fluid in yellow vent off Kueishan IslandKC527475, uncultured bacterium, White Plague Disease infected coralAY035822, Candidatus Arcobacter sulfidicus

uncultured 9

AF144693, oilfield bacterium ’FWKO B’

uncultured 7

JQ712410, uncultured epsilon proteobacterium, hydrothermal vent fluidsJQ712384, uncultured epsilon proteobacterium, hydrothermal vent fluids

AY542568, uncultured epsilon proteobacterium, Gulf of Mexico seafloor sedimentARB_34C3AB2A, denovo2381 GI2.1_228

AF458287, uncultured epsilon proteobacterium, Mono Lake, California

Cam

pylo

bacte

rale

s_C

am

pylo

bacte

raceae

_A

rcobacte

r

Figure S6b: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Epsilonproteobacteria, order: Campylobacterales family: Campylobacteraceae, genus: Acrobacter (black bold: sequences retrieved in this study and closest relatives).

Page 150: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

142

AFRZ01000001, Sulfurimonas gotlandica GD1, sulfidic area of central Baltic Sea

Fj717113, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717120, uncultured bacterium, marine sediment from Cullercoats, Northumberland

JQ712116, uncultured bacterium, oil−contaminated seawaterKF268813, uncultured bacterium, marine sedimentKF268857, uncultured bacterium, marine sediment

Ay922189, uncultured epsilon proteobacterium, grey whale bone, Pacific Ocean, Santa Cruz BasinGQ261771, uncultured epsilon proteobacterium, deep sea sediment associated with whale falls

uncultured 3

JQ287474, uncultured bacterium, East Pacific RiseJQ287473, uncultured bacterium, East Pacific Rise

JQ287433, uncultured bacterium, East Pacific RiseKC492834, uncultured Sulfurimonas sp., Baltic Sea redoxcline, 119 m depth

ARB_CC0227FD, denovo2021 GI2.1_1757FJ497608, uncultured epsilon proteobacterium, Vailulu’u SeamountFJ497591, uncultured bacterium, Vailulu’u SeamountFj717063, uncultured bacterium, marine sediment from Cullercoats, Northumberland

uncultured 8

Fj717036, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717079, uncultured bacterium, marine sediment from Cullercoats, Northumberland

Fj717049, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717093, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717084, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFJ717065, uncultured bacterium, marine sediment from Cullercoats, Northumberland

Fj717096, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717067, uncultured bacterium, marine sediment from Cullercoats, Northumberland

uncultured 11

DQ228650, uncultured bacterium, sulfide−microbial incubatorGQ261797, uncultured epsilon proteobacterium, deep sea sediment associated with whale falls

uncultured 4

AB176067, uncultured bacterium, PCR−derived sequence from vent area waterAB176111, uncultured bacterium, PCR−derived sequence from vent area waterAB176109, uncultured bacterium, PCR−derived sequence from vent area water

AY704397, uncultured bacterium, oceanic crustARB_4672D8F7, denovo1916 GI2.1_90EU487967, uncultured bacterium, siliciclastic sedment from Thalassia sea grass bed

FJ624448, uncultured bacterium, Lau Basin hydrothermal vent sedimentGQ848475, uncultured bacterium, deep sea hydrothermal vent sediment

GQ848408, uncultured bacterium, deep sea hydrothermal vent sedimentGU198381, bacterium enrichment culture clone 16S−9, marine sediment

JX120496, uncultured bacterium, subsurface aquifer sedimentARB_40E96DC5, denovo1140 GI2.1_3851

AB176113, uncultured bacterium, PCR−derived sequence from vent area waterGQ349272, uncultured epsilon proteobacterium, Saanich Inlet, 120 m depth

Fj717167, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandJQ287448, uncultured bacterium, East Pacific Rise

Jn873929, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseAb193899, uncultured bacterium, obtained from Mariana Trough hydrothermal vent waterJn874362, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific RiseJn873940, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

AB015257, unidentified epsilon proteobacterium, deepest cold−seep area of the Japan TrenchARB_EC842CDC, denovo3733 GII1.1_1431

Jn874097, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseARB_C5F1B8A5, denovo885 GI3.1_315AY211683, uncultured epsilon proteobacterium, Gulf of Mexico seafloor sediment

JQ712411, uncultured epsilon proteobacterium, hydrothermal vent fluids

uncultured 5

FJ497252, uncultured epsilon proteobacterium, Vailulu’u SeamountJN637808, uncultured marine microorganism, middle port

JN637788, uncultured marine microorganism, middle portFJ497618, uncultured gamma proteobacterium, Vailulu’u Seamount

HQ203919, uncultured bacterium, seawaterFJ497584, uncultured delta proteobacterium, Vailulu’u Seamount

FJ535288, uncultured bacterium, iron mat

mainly uncultured including Sulfurimonas autotrophica 25

FJ535336, uncultured bacterium, sedimentFJ535346, uncultured bacterium, sediment

DQ394918, uncultured epsilon proteobacterium, harbor sedimentARB_251565CE, denovo454 GII2.1_1533

JX391546, uncultured bacterium, marine sedimentAb697024, uncultured bacterium, Bacterial mat

Jq723654, uncultured Campylobacterales bacterium, biofilm AF068806, uncultured bacterium VC2.1 Bac32

EF644801, uncultured bacterium, MAR Logatchev hydrothermal vent system, 3000m water depth, sulfide sampleAF367490, uncultured epsilon proteobacterium

AF367496, uncultured epsilon proteobacteriumAY225613, uncultured epsilon proteobacterium, pumice fragments exposed to a Mid−Atlantic Ridge vent

EU574671, uncultured bacterium, hydrothermal vent mat; Champagne Vent on N.W. EifukuEU574674, uncultured bacterium, hydrothermal vent mat; Champagne Vent on N.W. Eifuku

uncultured 21

JX391632, uncultured bacterium, marine sedimentKF624431, uncultured bacterium, suboxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

Fr853059, uncultured bacterium, active hydrothermal chimney Juan de Fuca RidgeJq072913, uncultured prokaryote, active sulfide structures Juan de Fuca RidgeFr852963, uncultured bacterium, active hydrothermal chimney Juan de Fuca Ridge

Fr853045, uncultured bacterium, active hydrothermal chimney Juan de Fuca RidgeARB_E46846F5, denovo2532 GII1.1_85

Fr852998, uncultured bacterium, active hydrothermal chimney Juan de Fuca RidgeARB_D69AC317, denovo4058 GII1.1_2141Fr852969, uncultured bacterium, active hydrothermal chimney Juan de Fuca RidgeFJ497579, uncultured bacterium, Vailulu’u Seamount

DQ145977, bacterium ML2−86JX391568, uncultured bacterium, marine sediment

GQ261791, uncultured epsilon proteobacterium, deep sea sediment associated with whale fallsEF999352, uncultured bacterium, Pearl River Estuary sediments at 6cm depth

AY218555, uncultured bacteriumEF219007, uncultured bacterium, Ridge Flank Abyssal Hill hydrothermal site on the east Pacific rise

EF219001, uncultured bacterium, Ridge Flank Abyssal Hill hydrothermal site on the east Pacific riseJQ287049, uncultured bacterium, East Pacific Rise

JQ036261, uncultured bacterium, Eel River Basin sediment 2688EF687416, uncultured bacterium, sulfide−oxidizing mat, Chefren mud volcano, Nile Deep Sea Fan, Eastern Mediterranean

AJ441199, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretionsEU652648, uncultured bacterium, Yellow Sea sediment

EU662310, uncultured bacterium, microbial matAJ441200, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretions

DQ676325, uncultured epsilon proteobacterium, suboxic freshwater−pond sedimentKF464419, uncultured Sulfurimonas sp., Alang−Sosiya ship breaking yard

JQ611140, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandAF355050, uncultured epsilon proteobacterium

AY225615, uncultured epsilon proteobacterium, iron−rich substrate exposed to a Mid−Atlantic Ridge ventEF441903, uncultured epsilon proteobacterium, endolithic community in the acidic Rio Tinto basin

DQ071281, uncultured epsilon proteobacterium, deep−sea hydrothermal ventDQ071286, uncultured epsilon proteobacterium, deep−sea hydrothermal ventDQ228651, uncultured bacterium, sulfide−microbial incubator

KC682636, uncultured bacterium, inactive hydrothermal sulfide chimney, Tui Malila vent field, Lau BasinJQ611149, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

mainly uncultured including Sulfurimonas denitrificans 49

AY532543, uncultured bacterium, uranium−contaminated aquiferFJ628245, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 50

FJ628235, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 37.5KC682581, uncultured bacterium, inactive hydrothermal sulfide chimney, Tui Malila vent field, Lau Basin

ARB_314DFE6B, denovo3359 GII3.1_4254AY672532, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific Ocean

JQ611150, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan IslandAJ441204, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretions

AJ575997, uncultured epsilon proteobacterium, deep sea hydrothermal ventFJ628204, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 37.5

Ay251063, uncultured proteobacterium, deep−sea hydrothermal vent biofilm from the Edmond vent field JQ586311, uncultured bacterium, arctic marine sediment

ARB_FBD03AE6, denovo1223 GI2.1_522FJ628251, uncultured bacterium, brackish water from anoxic fjord Nitinat Lake at depth 50

KF624356, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaARB_7A60E850, denovo2375 GI2.1_220

ARB_97CD83C8, denovo1803 GII3.1_2388FJ535253, uncultured bacterium, iron mat

AJ575998, uncultured epsilon proteobacterium, deep sea hydrothermal ventARB_DFCFDB0A, denovo1215 GI1.1_2525

AJ441201, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretionsJn662179, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae in high−flow environmentDQ890422, uncultured bacterium, empty intestine

AJ441198, uncultured epsilon proteobacterium, Paralvinella palmiformis mucus secretionsHM154512, uncultured bacterium, Yellowstone Lake hydrothermal vent biofilmHM154504, uncultured bacterium, Yellowstone Lake hydrothermal vent biofilm

GU291331, uncultured bacterium, acid mine drainage contaminated salt marsh sedimentsKc682699, uncultured bacterium, inner conduit of inactive hydrothermal sulfide chimneyJQ287031, uncultured bacterium, East Pacific Rise

GU291328, uncultured bacterium, acid mine drainage contaminated salt marsh sedimentsGU291330, uncultured bacterium, acid mine drainage contaminated salt marsh sediments

JX403024, uncultured bacterium, underground waterKf624439, uncultured bacterium, suboxic sulfur−particle colonization experiment, Janssand tidal flat

uncultured 9

uncultured 3

JX521790, uncultured bacterium, terrestrial sulfidic springHQ729106, uncultured epsilon proteobacterium, bacterial matJX521773, uncultured bacterium, terrestrial sulfidic springJX521779, uncultured bacterium, terrestrial sulfidic springHQ729098, uncultured epsilon proteobacterium, bacterial matJX521774, uncultured bacterium, terrestrial sulfidic springHQ729096, uncultured epsilon proteobacterium, bacterial mat

Gu120613, uncultured epsilon proteobacterium, active mining area ARB_FD05D18F, denovo2886 GI1.1_2548

AM982624, uncultured bacterium, pig saw dust spent beddingFj037619, uncultured bacterium, biofilms of iron−ixodizing bacteria

JN637800, uncultured marine microorganism, middle portJn662232, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae

ARB_E5EE2E71, denovo1628 GI2.1_503JN637811, uncultured marine microorganism, middle portGU583978, uncultured bacterium, mangrove soilHM057666, uncultured bacterium, ocean water from the Yellow Sea

EF444735, uncultured bacterium, Thermopiles hot springsJX403049, uncultured bacterium, bitumenJX403058, uncultured bacterium, bitumen

FJ535290, uncultured bacterium, iron matFJ535325, uncultured bacterium, sedimentAB189712, Alviniconcha sp. gill symbiont

DQ228648, uncultured bacterium, sulfide−microbial incubatorAy251059, uncultured proteobacterium, deep−sea hydrothermal vent biofilm from the Edmond vent field DQ925870, uncultured bacterium, hydrothermal vent chimneys of Guaymas Basin

AB252048, Sulfurimonas paralvinellae, Paralvinella sp. polychaetes’ nestJn873966, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJn874012, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

Jn874124, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJn874136, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise

Jn874166, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseAy280406, uncultured epsilon proteobacterium, exterior of a deep−sea hydrothermal vent sulfide chimney, Juan de Fuca RidgeAY672505, uncultured bacterium, hydrothermal vent, 9 degrees North, East Pacific Rise, Pacific OceanJn662183, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae in high−flow environment

JQ712432, uncultured epsilon proteobacterium, hydrothermal vent fluidsJQ712447, uncultured epsilon proteobacterium, hydrothermal vent fluids

AY225611, uncultured epsilon proteobacterium, organic−rich substrate exposed to a Mid−Atlantic Ridge vent

uncultured 23

EF036297, uncultured bacterium, Zostera marina rootsAJ515712, uncultured epsilon proteobacterium

AJ515713, uncultured epsilon proteobacteriumJQ712418, uncultured epsilon proteobacterium, hydrothermal vent fluids

Jn662192, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae AY216438, uncultured bacterium, temperate estuarine mud

KF465274, uncultured Sulfurimonas sp., Alang−Sosiya ship breaking yardEU574677, uncultured bacterium, hydrothermal vent mat; Champagne Vent on N.W. EifukuAY211663, uncultured epsilon proteobacterium, Gulf of Mexico seafloor sediment

GQ357037, uncultured bacterium, methane seep sedimentGQ357021, uncultured bacterium, methane seep sedimentAB013263, uncultured epsilon proteobacterium, Nankai Trough sediments at a depth of 3,843m

ARB_3F56C503, denovo1533 GI2.1_190AY225612, uncultured epsilon proteobacterium, organic−rich substrate exposed to a Mid−Atlantic Ridge ventHq153998, uncultured bacterium, white filamentous microbial mat in the photic zone of a shallow hydrothermal ventAB188787, uncultured Helicobacter sp., cold seep sediment

FN553987, uncultured sediment bacterium, Logatchev hydrothermal vent field,site F, watercolumn depth = 3000 mFn554146, uncultured sediment bacterium, Logatchev hydrothermal vent field,Anya’s GardenJn874121, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseJQ712373, uncultured epsilon proteobacterium, hydrothermal vent fluids

FJ535303, uncultured bacterium, iron matJN874138, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’N

Jn874070, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific RiseKC682592, uncultured bacterium, inactive hydrothermal sulfide chimney, Tui Malila vent field, Lau Basin

JQ712437, uncultured epsilon proteobacterium, hydrothermal vent fluidsKF464524, uncultured bacterium, Alang−Sosiya ship breaking yardKF624354, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

GQ848454, uncultured bacterium, deep sea hydrothermal vent sedimentGQ848449, uncultured bacterium, deep sea hydrothermal vent sediment

JN874382, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific Rise, 9 degrees 50’NJN874185, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific Rise, 9 degrees 50’N

JN874190, uncultured bacterium, TrapR2 located 115 m southwest of Ty/Io vents, East Pacific Rise, 9 degrees 50’NJN873941, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’N

EU491825, uncultured bacterium, seafloor lavas from the East Pacific RiseEU491791, uncultured bacterium, seafloor lavas from the East Pacific Rise

EU491826, uncultured bacterium, seafloor lavas from the East Pacific RiseDQ070790, uncultured epsilon proteobacterium, basalt glass from 9N latitude East Pacific RiseEU491816, uncultured bacterium, seafloor lavas from the East Pacific RiseEU491792, uncultured bacterium, seafloor lavas from the East Pacific Rise

GQ848457, uncultured bacterium, deep sea hydrothermal vent sedimentJN873959, uncultured bacterium, TrapR1 located 30 m southwest of Bio9 vent complex, East Pacific Rise, 9 degrees 50’NJQ287099, uncultured bacterium, East Pacific Rise

JQ586308, uncultured bacterium, arctic marine sedimentARB_8340C096, denovo2767 GII2.1_5839KF624353, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

KF624411, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaKF624438, uncultured bacterium, suboxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

KF624409, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaFj717137, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717173, uncultured bacterium, marine sediment from Cullercoats, Northumberland

uncultured 18

Fj717179, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFj717157, uncultured bacterium, marine sediment from Cullercoats, NorthumberlandFJ615170, uncultured epsilon proteobacterium, Saanich Inlet, depth of 10m in anoxic marine fjord

ARB_1401C19A, denovo2596 GI1.1_1762EU652653, uncultured bacterium, Yellow Sea sediment

FJ535347, uncultured bacterium, sedimentKF465277, uncultured Sulfurimonas sp., Alang−Sosiya ship breaking yardKF465242, uncultured Sulfurimonas sp., Alang−Sosiya ship breaking yard

KF464511, uncultured bacterium, Alang−Sosiya ship breaking yardGU369904, uncultured epsilon proteobacterium, white filamentous microbial mat in the photic region of a shallow hydrothermal vent

KF465252, uncultured Sulfurimonas sp., Alang−Sosiya ship breaking yardEF379628, uncultured bacterium, sea water from Victoria Harbor (Hong Kong) area

ARB_3223D2E0, denovo508 GI3.1_1312KF624336, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

GQ261804, uncultured epsilon proteobacterium, deep sea sediment associated with whale fallsAY922185, uncultured epsilon proteobacterium, grey whale bone, Pacific Ocean, Santa Cruz Basin (N33.30 W119.22) depth 1674 meters

KF624343, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaARB_C6318A, denovo1788 GI1.1_3689

DQ395005, uncultured epsilon proteobacterium, harbor sedimentEU652651, uncultured bacterium, Yellow Sea sediment

ARB_5EA6F342, denovo917 GI2.1_381ARB_53282A96, denovo2773 GII2.1_8103

ARB_1615A492, denovo2370 GI2.1_3681EU652650, uncultured bacterium, Yellow Sea sediment

GQ349003, uncultured epsilon proteobacterium, Saanich Inlet, 100 m depthGQ261792, uncultured epsilon proteobacterium, deep sea sediment associated with whale falls

FJ497261, uncultured bacterium, Vailulu’u SeamountKF545003, uncultured Sulfurimonas sp., deep−sea sediment

JN977221, uncultured bacterium, Jiaozhao Bay sedimentAB530237, uncultured bacterium, marine sediment

KF624267, uncultured bacterium, anoxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaKF624342, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden SeaKF624358, uncultured bacterium, oxic sulfur−particle colonization experiment, Janssand tidal flat, Wadden Sea

JQ196547, uncultured bacterium, seawater; next to dolphin EARB_77205F9F, denovo889 GI2.1_254

JQ586310, uncultured bacterium, arctic marine sedimentFJ205271, uncultured epsilon proteobacterium, active hydrothermal field sediments, depth:1922m

EU652667, uncultured bacterium, Yellow Sea sedimentEU266014, uncultured bacterium, Nitinat Lake at a depth of 26 m

KF801506, uncultured epsilon proteobacterium

Cam

pylo

bacte

raceae

_S

ulfurim

onas

mainly uncultured related to Sulfuricurvum

256

uncultured 7

Figure S6c: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Epsilonproteobac-teria, order: Campylobacterales family: Campylobacteraceae, genus: Sulfurimonas (black bold: sequences retrieved in this study and closest relatives).

Page 151: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

143

X82931, Sulfurospirillum multivoransJQ713915, uncultured Sulfurospirillum sp., sediments from shallow hydrothermal vents

CP007201, Sulfurospirillum multivorans DSM 12446CP007201, Sulfurospirillum multivorans DSM 12446JX403043, uncultured bacterium, surface water

JX521007, uncultured bacterium, terrestrial sulfidic springEu864435, uncultured bacterium, Xiao River wastewater

Jq087039, uncultured Sulfurospirillum sp.,sedimenthydrocarbon contaminated aquiferAY570615, uncultured bacterium, low−temperature biodegraded Canadian oil reservoir

JX521473, uncultured bacterium, terrestrial sulfidic springJX521469, uncultured bacterium, terrestrial sulfidic spring

JN685481, uncultured bacterium, production water sample from oil wellJF304793, uncultured Sulfurospirillum sp., mining wastewater

Fq659417, uncultured soil bacterium, PAH−contaminated soil; retention systems HM481314, uncultured bacterium, TCE−contaminated field site

FJ612331, uncultured bacterium, lake waterCU922216, uncultured bacterium, mesophilic anaerobic digester

Dq295746, uncultured epsilon proteobacterium, microbial mat in sulfidic springDq295743, uncultured epsilon proteobacterium, microbial mat in sulfidic spring

AB355036, uncultured bacterium, surface water, Manzallah LakeJF747911, uncultured bacterium, Frasassi cave system sulfidic springAB355034, uncultured bacterium, surface water, Manzallah Lake

AB186804, uncultured bacterium, uncultured clone from polychlorinated dioxin dechlorinating microcosmAY756596, uncultured bacterium, Siwhaho sediment

Jx271945, uncultured bacterium, activated sludge in lab−scale reactor JX521449, uncultured bacterium, terrestrial sulfidic spring

Fq660045, uncultured soil bacterium, PAH−contaminated soilFq659173, uncultured soil bacterium, PAH−contaminated soilAf144694, Campylobacter sp. DSM 806

Fq659063, uncultured soil bacterium, PAH−contaminated soilL14632, Campylobacter sp.Jx462539, uncultured Campylobacter sp., methanogenic bioelectrochemical reactor

AY780560, uncultured Sulfurospirillum sp., chlorinated ethene−degrading cultureFJ612315, uncultured bacterium, lake waterAY189928, Sulfurospirillum sp. JPD−1FQ659988, uncultured soil bacterium, PAH−contaminated soil; retention systems which treat road runoffs

FQ659274, uncultured soil bacterium, PAH−contaminated soil; retention systems which treat road runoffsHM481322, uncultured bacterium, TCE−contaminated field site

JF789596, uncultured Sulfurospirillum sp., Athabasca oil sands formation waterFJ612310, uncultured bacterium, lake water

DQ234114, uncultured Sulfurospirillum sp., mangroveFJ612314, uncultured bacterium, lake water

FJ612328, uncultured bacterium, lake waterFJ612333, uncultured bacterium, lake water

FJ612317, uncultured bacterium, lake waterKF933844, Sulfurospirillum sp. b10KF933843, Sulfurospirillum sp. b1

DQ234236, uncultured Sulfurospirillum sp., mangroveAY756183, arsenate−reducing bacterium NP4, groundwater

JX521339, uncultured bacterium, terrestrial sulfidic springFQ659969, uncultured soil bacterium, PAH−contaminated soil; retention systems which treat road runoffs

AB368775, Sulfurospirillum deleyianumAB286550, uncultured bacterium, activated sludge

AB246781, Sulfurospirillum cavolei, underground crude−oil storage cavityJx000050, uncultured bacterium, anaerobic wastewater reactorJx271924, uncultured bacterium, activated sludge in lab−scale reactor with dissolved oxygen above 2.5 mg/lAJ488086, uncultured bacterium

Kc179071, uncultured bacterium, activated sludge taken from a microaerobic bioreactor KF836224, uncultured bacterium, springs and wells fed by a deep, fractured rock aquifer in the Mojave DesertAF038843, Sulfurospirillum barnesii

U41564, Sulfurospirillum barnesiiCP003333, Sulfurospirillum barnesii SES−3, selenate−contaminated freshwater marsh sedimentCp003333, Sulfurospirillum barnesii SES−3, selenate−contaminated freshwater marsh sediment

AJ535704, Sulfurospirillum sp. EK7, river sedimentY13671, Sulfurospirillum deleyianumCP001816, Sulfurospirillum deleyianum DSM 6946CP001816, Sulfurospirillum deleyianum DSM 6946CP001816, Sulfurospirillum deleyianum DSM 6946

AJ249226, bacterium DCE10GU120580, uncultured epsilon proteobacterium, active mining area of Pitch LakeEU662599, uncultured bacterium, floating microbial mat from sulfidic waterJX399885, Sulfuricurvum sp. L12, enrichment culture of coal−tar from a natural coal−tar source

Am157656, epsilon proteobacterium AN−BI3A, hypersaline brine and seawater in the Mediterranean.HM468037, uncultured bacterium, solid waste with green color in tanning process before Cr recovery

FN356324, uncultured bacterium, oil production waterFN356312, uncultured bacterium, oil production water

DQ228139, Sulfurospirillum sp. C6, bioreactor inoculated with oil field produced waterAY570614, uncultured bacterium, low−temperature biodegraded Canadian oil reservoirAY570602, uncultured bacterium, low−temperature biodegraded Canadian oil reservoirGQ863490, Sulfurospirillum alkalitolerans

AY570598, uncultured bacterium, low−temperature biodegraded Canadian oil reservoirAY570558, uncultured bacterium, low−temperature biodegraded Canadian oil reservoirJx898111, uncultured bacterium, series of continously fed aerated tank reactors system

Gq133407, uncultured bacterium, ASBR reactor treating swine wasteGQ415373, uncultured bacterium, Daqing oilfield production water

AB546014, uncultured bacterium, #AR39 oil wellheadJQ088439, uncultured bacterium, crude oil reservoir

JQ214236, uncultured bacterium, bottlenose dolphin mouthJQ214253, uncultured bacterium, bottlenose dolphin mouth

JQ215461, uncultured bacterium, bottlenose dolphin mouthAB546035, uncultured bacterium, #AR39 oil wellhead

AB546080, uncultured bacterium, #SR123 oil wellheadAB192073, uncultured epsilon proteobacterium, Microcerotermes sp. 2AB192072, uncultured epsilon proteobacterium, Microcerotermes sp. 1

AJ576349, uncultured bacterium, hindgut homogenate of Pachnoda ephippiata larvaAB231076, uncultured epsilon proteobacterium, Neotermes koshunensis

JN619494, uncultured bacterium, insect gutJN620073, uncultured bacterium, insect gut

AB234540, uncultured epsilon proteobacterium, Macrotermes gilvusFJ792313, uncultured bacterium, carbonate chimney in Lost City Hydrothermal Field

FJ792353, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFJ792370, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFJ791985, uncultured bacterium, carbonate chimney in Lost City Hydrothermal Field

FJ792434, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFJ792359, uncultured bacterium, carbonate chimney in Lost City Hydrothermal Field

ACQI01000338, hydrothermal vent metagenome, Lost City hydrothermal chimney biofilmFJ792218, uncultured bacterium, carbonate chimney in Lost City Hydrothermal Field

FJ792357, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFJ792283, uncultured bacterium, carbonate chimney in Lost City Hydrothermal FieldFJ792210, uncultured bacterium, carbonate chimney in Lost City Hydrothermal Field

KC682802, uncultured bacterium, surface of a basalt collected from Kilo Moana vent field, Lau BasinJN637820, uncultured marine microorganism, middle port

ARB_E5A35806, denovo3111 GII1.1_6079JX391284, uncultured bacterium, marine sedimentJX391327, uncultured bacterium, marine sedimentU85965, Sulfurospirillum sp. SM−5

AF367491, uncultured epsilon proteobacteriumFn773278, bacterium endosymbiont of Osedax mucofloris, Osedax mucofloris extacted from whale boneAY548994, uncultured bacterium, marine environment

FN658698, epsilon proteobacterium ectosymbiont of Rimicaris FJ717081, uncultured bacterium, marine sediment from Cullercoats, Northumberland

FJ497338, uncultured bacterium, Vailulu’u SeamountFJ814746, uncultured Campylobacterales bacterium, Carmel Canyon

ARB_E2983623, denovo180 GII3.1_8793FJ497463, uncultured epsilon proteobacterium, Vailulu’u Seamount

FJ535312, uncultured bacterium, iron matAY211681, uncultured epsilon proteobacterium, Gulf of Mexico seafloor sediment

Y11561, Sulfurospirillum arcachonenseGU220754, uncultured bacterium, low temperature hydrothermal precipitates of East Lau Spreading Center

ARB_9981C26F, denovo2147 GII3.1_1394AF357198, Sulfurospirillum sp. Am−NAF357199, Sulfurospirillum sp. 18.1

HM128339, uncultured bacterium, Xiaochaidan LakeAF513952, uncultured Sulfurospirillum sp., Hawaiian lake water

AB550501, uncultured Sulfurospirillum sp.FJ497442, uncultured epsilon proteobacterium, Vailulu’u Seamount

JQ287455, uncultured bacterium, East Pacific RiseJQ286985, uncultured bacterium, East Pacific Rise

JQ287296, uncultured bacterium, East Pacific RiseJQ287108, uncultured bacterium, East Pacific Rise

KC682703, uncultured bacterium, inner conduit of inactive hydrothermal sulfide chimney, ABE vent field, Lau BasinJQ611155, uncultured bacterium, core sediment at 18−21 cm depth in yellow vent off Kueishan Island

KC682704, uncultured bacterium, inner conduit of inactive hydrothermal sulfide chimney, ABE vent field, Lau BasinFJ497620, uncultured bacterium, Vailulu’u Seamount

Ay280408, uncultured epsilon proteobacterium, exterior of a deep−sea hydrothermal vent sulfide chimney, Juan de Fuca RidgeKC357921, uncultured bacterium, large−discharge carbonate springs

Jn662188, uncultured epsilon proteobacterium, tubeworm Ridgeia piscesae in high−flow environmentEf462744, uncultured bacterium, dorsal surface of the deep sea hydrothermal vent polychaete worm Alvinella pompejana

Sulfurospirillum

Thiofractor 6

0.10

Figure S6d: Bacterial 16S rRNA gene amplicon sequences assigned to the class of Epsilonproteobacte-ria, order: Campylobacterales family: Campylobacteraceae, genus: Sulfurospirillum (black bold: sequences retrieved in this study and closest relatives).

Page 152: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 5

144

Supplementary Material

Figure S7: Characteristic PLFA profiles of (a) the average relative concentration of PLFAs for March (0–0.5 cm and 2–4 cm) and August (0–1 cm), (b) the average Δδ 13C values and (c) relative 13C incorpora-tion of PLFAs for S2 and S1/S3 March (0–1 cm).

Page 153: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 154: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 155: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

147

6. Potential recycling of thaumarchaeotal lipids by DPANN archaea in seasonally hypoxic surface marine sediments

Yvonne A. Lipsewers, Ellen C. Hopmans, Jaap S. Sinninghe Damsté, and Laura Villanueva

Thaumarchaeota are known to synthesize specific glycerol dibiphytanyl glycerol tetraethers (GDGTs), the distribution of which is affected by temperature which forms the base of the paleotemperature proxy TEX86. Lipids present in the marine surface sediments are believed to be mostly derived from pelagic Thaumarchaeota; however, some studies have evaluated the possibility that benthic archaea also contribute to the lipid fossil record. Here, we compared the archaeal abundance and composition by DNA-based methods and the archaeal intact polar lipid (IPL) diversity in surface sediments of a seasonally hypoxic marine lake to determine the potential biological sources of the archaeal IPLs detected in surface sediments under changing environmental conditions. The archaeal community changed from March (oxic conditions) to August (euxinic) from a Thaumarchaeota-dominated (up to 82%) to an archaeal community dominated by the DPANN superphylum (up to 95%). This marked change coincided with a one order of magnitude decrease in the total IPL-GDGT abundance. In addition, the polar head groups of the IPL-GDGTs changed from predominantly containing phospho- to glyco- polar head groups. This may indicate a transition to Thaumarchaeota growing in stationary phase or selective preservation of the GDGT pool. However, considering the apparent inability of the DPANN archaea to synthesize their own membrane lipids, we hypothesize that an alternative explanation might be that DPANN archaea use the lipids previously synthesized by the Thau-marchaeota to form their own cell membranes, which would indicate an active recycling of fossil IPLs in the marine surface sediment.

Submitted to Organic Geochemistry

Page 156: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

148

6.1. IntroductionArchaea occur ubiquitously, are abundant in aquatic and terrestrial habitats, and play an import-ant role in global biogeochemical cycles (Jarrell et al., 2011). Marine sediments have been shown to harbor a diverse archaeal community, of which ammonia-oxidizing Thaumarchaeota (marine group I.1a) are the most studied (Pester et al., 2011). Recent genomic studies have revealed the presence of other benthic archaeal groups, such as Miscellaneous Crenarchaeota Group (MCG), Marine Benthic Group-B and D, and archaea of the DPANN superphylum, in marine sediments, which are predicted to be involved in degradation of polymers and proteins under anoxic con-ditions (Castelle et al., 2015; Lloyd et al., 2013, Meng et al., 2014).

Archaeal lipids occur widespread in marine sediments and are commonly used as biomarkers of Archaea both in present-day systems and in past depositional environments. However, their biological sources are not well constrained, especially in the light of the expanding archaeal di-versity. Archaeal lipids in marine surface sediments are thought to be derived mainly from pelagic Thaumarchaeota, which due to grazing and packing in fecal pellets are efficiently transported to the sediment where they get preserved in the sedimentary record (Huguet et al., 2006b). Thaumarchaeota are known to synthesize isoprenoid glycerol dibiphytanyl glycerol tetraether (GDGT) containing 0–4 cyclopentane moieties (GDGT-0 to GDGT-4) as well as the GDGT crenarchaeol (Sinninghe Damsté et al., 2002), containing four cyclopentane moieties and a cy-clohexane moiety, which is considered to be characteristic of this phylum (Pearson et al., 2004; Zhang et al., 2006; Pester et al., 2011; Pitcher et al., 2011a; Sinninghe Damsté et al., 2012). The distribution of thaumarchaeotal GDGTs in the marine environment has been shown to be affected by temperature, i.e. with increasing temperature there is an increase in the relative abun-dance of cyclopentane-containing GDGTs (Schouten et al., 2002; Wuchter et al., 2004, 2005). Based on this relationship, the TEX86 paleotemperature proxy was developed and calibrated on sea surface temperature (e.g. Schouten et al., 2002; Kim et al., 2010) and it has been widely ap-plied for more than a decade (Schouten et al., 2013).

Although it is generally thought that GDGTs in marine sediments derive from surface-derived thaumarchaeotal biomass (e.g. Wakeham et al., 2003), some studies have addressed the possibility of a potential contribution of lipids from benthic archaea to the fossil record, which would be relevant for the reliability of TEX86 (Shah et al., 2008; Biddle et al., 2006; Lipp et al., 2008; Lipp and Hinrichs, 2009). Archaeal intact polar lipids (IPLs), where the core lipid (CL) GDGTs are attached to polar head groups from the building blocks of membranes of living cells with phos-pho head groups have been shown to degrade rapidly upon death of the source organism (White et al., 1979; Harvey et al., 1986), while IPLs with glyco polar head groups may be preserved over (much) longer timescales (Lipp et al., 2008; Lipp and Hinrichs, 2009; Bauersachs et al., 2010). IPLs of Archaea have been detected in surface marine sediments and used as biomarkers of the presence of living in situ benthic Archaea (Biddle et al., 2006; Lipp et al., 2008; Lipp and Hin-richs, 2009; Lengger et al., 2012). Evidence has been presented that the IPL-GDGTs produced in situ in the surface sediments are rapidly recycled upon burial (Lengger et al., 2012). Some studies have suggested that benthic marine archaea recycle fossil CL-GDGTs when producing IPL-GDGTs de novo to decrease energy requirements (Liu et al., 2011; Takano et al., 2010).

With the exception of Thaumarchaeota, the membrane lipid composition of other phyla of benthic archaea present in marine sediments has not been yet characterized and, therefore, their potential impact on the archaeal lipid pool preserved in the sedimentary record remains un-

Page 157: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

149

Introduction/Material and Methods

known. In surface sediments where oxygen is still available, Thaumarchaeota are expected to be dominant due to their oxygenic metabolism as nitrifiers (Könneke et al., 2005, Wuchter et al., 2006). However in subsurface sediments, where oxygen is no longer present, archaeal groups such as MBG-D, MCG, Thermoplasmatales and methanogens have been reported to be pre-dominant (Lloyd et al., 2013; Kubo et al., 2012), and may contribute to the total archaeal lipid pool in marine sediments. Recent studies have also observed a high presence of archaea of the superphylum DPANN both in surface and subsurface coastal marine sediments (Choi et al., 2016), as well as in freshwater systems (Ortiz-Alvarez et al., 2016; Ma et al., 2016). However, their contribution to the sedimentary archaeal lipid pool is expected to be negligible as their small genomes (Castelle et al., 2015) lack the genes of the membrane lipid biosynthetic pathway, suggesting that they rely on host cells or cell debris for the synthesis of their lipids (Waters et al., 2003; Jahn et al., 2004).

Here, we determined the archaeal abundance and composition using DNA-based methods in surface sediments of a seasonally hypoxic marine lake with different oxygen and sulfide bottom water concentrations and compared this with the composition of the archaeal IPLs. We aim to identify the potential biological sources of the archaeal IPL lipids detected in surface sediments under changing environmental conditions which could impact the biology of the producer but also the preservation potential of the archaeal lipids.

6.2. Material and Methods

6.2.1. Study site, sediment sampling and physicochemical analysesLake Grevelingen is a former estuary located within the Rhine-Meuse-Scheldt delta area of the Netherlands. The delta became a closed saline reservoir (salinity ~30) by dam construction at both the land side and sea side in the early 1970s. As a result of the absence of tides and strong currents, Lake Grevelingen experiences a seasonal stratification of the water column, which in turn, leads to a depletion of oxygen in the bottom waters (Hagens et al., 2015). Bottom water oxygen concentration at the deepest stations starts to decline in April, reaches hypoxic condi-tions by end of May (O2 < 63 µM), and further decreases to reach the state of anoxia in August (O2 < 0.1 µM), while re-oxygenation of the bottom water takes place in September (Seitaj et al., 2015).

Two sampling campaigns were performed on March 13th, 2012 (before the start of the annual O2 depletion) and on August 20th, 2012 (at the peak of the annual O2 depletion). Detailed water column, pore water and solid sediment chemistry of Lake Grevelingen over the year 2012 have been previously reported (Seitaj et al., 2015; Sulu-Gambari et al., 2016, Hagens et al., 2015). In-tact sediment cores were recovered at three stations along a depth gradient within the Den Osse basin, one of the deeper basins in Marine Lake Grevelingen: Station 1 (S1) was located in the deepest point (34 m) of the basin (51.747°N, 3.890°E), Station 2 (S2) at 23 m depth (51.749°N, 3.897°E) and Station 3 (S3) at 17 m depth (51.747°N, 3.898°E) (Fig. S1). Sediment cores were retrieved with a single core gravity corer (UWITEC) using PVC core liners (6 cm inner diameter, 60 cm length). Further details of sediment sampling have been previously described in detail (Chapter 4). Four sediment cores were sliced with a 1 cm resolution (for the purposes of this study we only focused on the top centimeter) for lipid and DNA/RNA analysis and kept at –80ºC until further processing.

Page 158: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

150

The total organic carbon (TOC) content of the sediment was determined on sub-samples that were freeze-dried, ground to a fine powder and analyzed by an a Thermo Finnigan Delta plus isotope ratio monitoring mass spectrometer (irmMS) connected to a Flash 2000 elemental analyzer (Thermo Fisher Scientific, Milan). Before the analysis, samples were first acidified with 2N hydrogen chloride (HCl) to remove the inorganic carbon (Nieuwenhuize et al., 1994). Con-centrations of TOC are expressed as mass % of dry sediment. Oxygen, sulfide, bottom water and pore water ammonium, nitrite and nitrate were determined as previously described (Malkin et al., 2014; Seitaj et al., 2015) and reported in Chapter 5.

6.2.2. DNA/RNA extractionSediment was centrifuged and the excess of water was removed by pipetting before proceeding with the extraction of nucleic acids from the sediment. DNA/RNA of surface (0–1 cm) sedi-ments was extracted with the RNA PowerSoil® Total Isolation Kit plus the DNA elution acces-sory (Mo Bio Laboratories, Carlsbad, CA). Concentration of DNA was quantified by Nanodrop (Thermo Scientific, Waltham, MA) and Fluorometric with Quant-iTTM PicoGreen® dsDNA Assay Kit (Life technologies, Netherlands).

6.2.3. 16S rRNA gene amplicon sequencing and sequencing analysesPCR reactions were performed with the universal, Bacteria and Archaea, primers S-D-Arch-0159-a-S-15 and S-D-Bact-785-a-A-21 (Klindworth et al., 2013) as previously described in Moore et al. (2015). The archaeal 16S rRNA gene amplicon sequences were analyzed by QIIME v1.9 (Caporaso et al., 2010). Raw sequences were demultiplexed and then quality-filtered with a minimum quality score of 25, length between 250–350, and allowing maximum two errors in the barcode sequence. Taxonomy was assigned based on blast and the SILVA database version 123 (Altschul et al., 1990; Quast et al., 2013). The 16S rRNA gene amplicon reads (raw data) have been deposited in the NCBI Sequence Read Archive (SRA) under BioProject no. PRJNA293286.

6.2.4. PCR amplification, cloning and archaeal 16S rRNA gene quantificationAmplification of the archaeal amoA gene was performed as described by Yakimov et al. (2011). PCR reaction mixture was the following (final concentration): Q-solution 1 × (PCR additive, Qiagen); PCR buffer 1 ×; BSA (200 µg ml–1); dNTPs (20 µM); primers (0.2 pmol µl–1); MgCl2 (1.5 mM); 1.25 U Taq polymerase (Qiagen, Valencia, CA, USA). PCR conditions for these ampli-fications were the following: 95°C, 5 min; 35 × 95°C, 1 min; 55°C, 1 min; 72°C, 1 min]; final ex-tension 72°C, 5 min. PCR products were gel purified (QIAquick gel purification kit, Qiagen) and cloned in the TOPO-TA cloning® kit from Invitrogen (Carlsbad, CA, USA) and transformed in E. coli TOP10 cells following the manufacturer’s recommendations. Recombinant clones plasmid DNAs were purified by Qiagen Miniprep kit and screening by sequencing (n ≥ 30) using M13R primer by BaseClear (Leiden, The Netherlands). Obtained archaeal amoA protein sequences were aligned with already annotated amoA sequences by using the Muscle application (Edgar, 2004). Phylogenetic trees were constructed with the Neighbor-Joining method (Saitou and Nei, 1987) and evolutionary distances computed using the Poisson correction method with a boot-strap test of 1,000 replicates. Quantification of archaeal 16S rRNA gene copies was performed by quantitative PCR (qPCR) by using the primers Parch519F and ARC915R as previously de-scribed (Pitcher et al., 2011b).

Page 159: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

151

Material and Methods/Results

6.2.5. Lipid extraction and analysisTotal lipids were extracted from surface (upper 0–1 cm) sediments after freeze-drying using a modified Bligh and Dyer method (Bligh and Dyer, 1959) following an experimental protocol previously described by Lengger et al. (2014). The extracts were then dissolved by adding solvent (hexane:isopropanol:H2O 718:271:10 [v/v/v/v]) and filtered through a 0.45 µm, 4 mm-diameter True Regenerated Cellulose syringe filter (Grace Davison, Columbia, MD, USA).

IPLs were analyzed with Ultra High Pressure Liquid Chromatography- High Resolution Mass spectrometry (UHPLC-HR MS). An Ultimate 3000 RS UHPLC, equipped with thermostated auto-injector and column oven, coupled to a Q Exactive Orbitrap MS with Ion Max source with heated electrospray ionization (HESI) probe (Thermo Fisher Scientific, Waltham, MA), was used. Separation was achieved on a YMC-Triart Diol-HILIC column (250 x 2.0 mm, 1.9 µm particles, pore size 12 nm; YMC Co., Ltd, Kyoto, Japan) maintained at 30°C. The following elution program was used with a flow rate of 0.2 mL min–1: 100% A for 5 min, followed by a linear gradient to 66% A: 34% B in 20 min, maintained for 15 min, followed by a linear gradient to 40% A: 60% B in 15 min, followed by a linear gradient to 30% A: 70% B in 10 min, where A = hexane/2-propanol/formic acid/14.8 M NH3aq (79:20:0.12:0.04 [v/v/v/v]) and B = 2-propa-nol/water/formic acid/ 14.8 M NH3aq (88:10:0.12:0.04 [v/v/v/v]). Total run time was 70 min with a re-equilibration period of 20 min in between runs. HESI settings were as follows: sheath gas (N2) pressure 35 (arbitrary units), auxiliary gas (N2) pressure 10 (arbitrary units), auxiliary gas (N2) T 50 ˚C, sweep gas (N2) pressure 10 (arbitrary units), spray voltage 4.0 kV (positive ion ESI), capillary temperature 275 °C, S-Lens 70 V. IPLs were analyzed with a mass range of m/z 375 to 2000 (resolution 70,000), followed by data-dependent MS2 (resolution 17,500), in which the ten most abundant masses in the mass spectrum (with the exclusion of isotope peaks) were fragmented successively (stepped normalized collision energy 15, 22.5, 30; isolation window 1.0 m/z). An inclusion list was used with a mass tolerance of 3 ppm to target specific compounds (supplementary File S1). The Q Exactive Orbitrap MS was calibrated within a mass accuracy range of 1 ppm using the Thermo Scientific Pierce LTQ Velos ESI Positive Ion Calibration Solution (containing a mixture of caffeine, MRFA, Ultramark 1621, and N-butylamine in an acetonitrile-methanol-acetic solution). IPLs were quantified by integrating the summed mass chromatograms (within 3 ppm mass accuracy) of the dominant adduct formed (in case of MH-, DH- and HPH-IPLs the ammoniated adduct) and the first isotopomer and reported as peak area response per gram of dry sediment extracted, due to the lack of quantitative standards (for details see Pitcher et al., 2011b).

The total lipid extracts from the surface sediments were further analyzed by acid hydrolysis to determine the composition and relative abundance of IPL-derived CL (resulting from the acid hydrolysis of IPLs) and CL-GDGTs using the method described by Lengger et al. (2012) and analyzed by via high-performance liquid chromatography-atmospheric pressure chemical ioniza-tion/mass spectrometry (HPLC-APCI/MS) (Schouten et al., 2007) using an internal C46 GTGT standard as described previously (Huguet al. 2006a).

6.3. Results

6.3.1. Physicochemical conditionsThe seasonal variation of the bottom water oxygen concentration in Lake Grevelingen strongly

Page 160: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

152

influenced the bottom and pore water concentrations of O2 and sulfide (Table 1). In March, bot-tom waters were fully oxygenated in all stations (299–307 μM), oxygen penetrated 1.8–2.6 mm deep in the sediment, and no free sulfide was recorded in the first few centimeters (Hagens et al., 2015). The width of the suboxic zone, operationally defined as the sediment layer located between the oxygen penetration depth (OPD) and the sulfide appearance (SAD), varied between 16–39 mm across the three stations in March 2012. In contrast, in August, oxygen was strongly depleted in the bottom waters at S1 (<0.1 μM) and S2 (11 μM), and no O2 was detected by mi-crosensor profiling in the surface sediment at these two stations. At station S3, the bottom water concentration of O2 remained higher (88 μM), and oxygen still penetrated into the surface sed-iment down to 1.1 mm. In August, free sulfide was present near the sediment-water interface at all three stations, and the concentration of sulfide in the pore water increased with water depth.

Bottom water ammonium (NH4+) concentrations in station S1 ranged from 3 µM in March to

11.5 µM in August; nitrite (NO2–) concentrations were relatively constant (0.7–1 µM) in March

and August. Nitrate concentrations ranged from 28 µM in March to <2 µM in August in station S1, whereas in stations S2 and S3 values varied between 28 µM in March to ~10 µM in August (Table 1; for detailed bottom water biogeochemistry see Seitaj et al., 2015; Sulu-Gambari et al., 2016; Hagens et al., 2015; Chapter 4).The TOC content of the sediments varied slightly between stations and seasons, ranging between 1.8 and 4.4% as described previously (Table 1; Chapter 4).

6.3.2. Archaeal lipid diversity and distributionThe archaeal IPL GDGT concentration (quantified as IPL-derived CLs) was almost an order of magnitude higher in March in all stations in comparison with the values in August, although this drop was less for station 3, where slightly lower concentrations of GDGT-0 and crenarchaeol were detected in March in comparison with the other stations (Table 2). In March, total IPL-GDGTs were mostly distributed between IPL-GDGT-0 (approximately 5 μg g–1 dry weight of sediment) and IPL-crenarchaeol (average 2.6 μg g–1). In August, the total IPL-GDGT abundance was lower in comparison with March (approximately an order of magnitude less), while the dis-tribution did not change (Table 2). CL-GDGT abundance was also higher in March (26 to 79 μg g–1 dry weight of sediment) than in August (5 to 15 μg g–1 dry weight of sediment) (Table 2). CL-GDGT abundances were also higher than those of IPL-GDGTs at all stations and seasons (Table 2). In station 1 it was on average three times higher both in March and in August. On the other hand in station 2, CL-GDGTs were in average 8 times higher than IPL-GDGTs in March, and 30 times higher in August, with CL-crenarchaeol as the most abundant CL-GDGT in August (8 μg g–1 dry weight of sediment; Table 2), 40 times more of that detected for IPL-cre-narchaeol (0.2 μg g–1; Table 2). In station 3, CL-GDGTs were on average 4 times higher in con-centration than IPL-GDGTs (Table 2).

The IPL composition was assessed by UHPLC-HRMS analysis and various GDGT-IPLs and archaeol-IPLS were specifically targeted by using an inclusion list as part of the analytical rou-tine (see supplementary File S1). Within the various IPL types detected, the distribution of observed cores was comparable to the distribution obtained after analysis of the IPL-derived CLs. IPLs with GDGT-0 and crenarchaeol as CL are the most abundant IPLs, with the highest concentrations in March (approximately 1010 response units g–1), while IPLs with GDGT-1 and -2 as CLs were two orders of magnitude less abundant (Table 3). In August, total IPL-GDGT concentrations decreased by two orders of magnitude as compared with the concentrations in

Page 161: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

153

ResultsTa

ble

1: P

hysic

oche

mic

al p

aram

eter

s of

the

botto

m w

ater

and

surf

ace

sedi

men

t (0–

1 cm

) in

the

thre

e st

atio

ns in

Lak

e G

reve

linge

n in

sprin

g (M

arch

) an

d in

sum

mer

(Aug

ust).

Dat

a in

clud

ed in

Sei

taj e

t al.,

201

5; S

ulu-

Gam

bari

et a

l., 2

016;

Hag

ens e

t al.,

201

5; C

hapt

er 4

and

5.

Bot

tom

wat

erSt

atio

n 1

Stat

ion

2St

atio

n 3

Aug

ust

Mar

chA

ugus

tM

arch

Aug

ust

Mar

ch

Tem

pera

ture

[°C

]5

175

175

19

O2 [

μM]

299

030

112

307

88

(oxi

c)(a

noxi

c)(o

xic)

(hyp

oxic

)(o

xic)

(hyp

oxic

)

NH

4+ [μ

M]

3.2

11.5

3.0

4.3

2.8

2.5

NO

2− [μ

M]

0.7

1.0

0.7

0.7

0.7

0.1

NO

3− [μ

M]

28.2

1.7

27.9

11.6

27.7

10.6

Surf

ace

sedi

men

t

HS−

[μM

]0

810

011

570

211

NH

4+ [μ

M]

279

656

165

550

7353

7

TOC

[%]

2.86

1.81

3.11

2.38

2.87

3.04

OPD

[mm

]1.

8±0.

040

2.6±

0.65

02.

4±0.

41.

1±0.

1

SAD

[mm

]17

.5±

0.7

0.9±

1.1

21.3

±2.

50.

6±0

41.8

±8.

64.

2±2.

7Bo

ttom

wat

er is

cla

ssifi

ed a

s ano

xic

with

O2 c

once

ntra

tion

belo

w 1

μM

and

hyp

oxic

bel

ow 6

3 μM

; OPD

: O2 p

enet

ratio

n de

pth;

SA

D: ∑

H2S

app

eara

nce

dept

h.

Page 162: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

154

Tabl

e 2:

Abu

ndan

ce a

nd d

istrib

utio

na o

f IP

L-de

rived

(rel

ease

d by

aci

d hy

drol

ysis)

and

CL

GD

GTs

(µg

g-1 o

f dr

y w

eigh

t) in

the

surf

ace

sedi

men

t (0–

1 cm

) of

the

thre

e st

atio

ns in

Lak

e G

reve

linge

n in

sprin

g (M

arch

) and

in su

mm

er (A

ugus

t).

Stat

ion

1St

atio

n 2

Stat

ion

3M

arch

Aug

ust

Mar

chA

ugus

tM

arch

Aug

ust

IPL-

deriv

ed

GD

GTs

GD

GT-

05.

1 (5

6.9)

0.8

(48.

2)5.

9 (6

0.5)

0.2

(45.

5)3.

0 (5

8.6)

1.7

(54.

6)

GD

GT-

10.

4 (4

.6)

0.1

(5.5

)0.

4 (4

.4)

0.02

(4.9

)0.

3 (5

)0.

1 (4

.6)

GD

GT-

20.

2 (2

.2)

0.1

(3.7

)0.

2 (2

.2)

0.01

(2.6

)0.

1 (2

.7)

0.1

(2.3

)

GD

GT-

30.

1 (0

.9)

0.02

(1.3

)0.

1 (0

.8)

0.01

(1.1

)0.

05 (1

)0.

03 (0

.9)

Cre

nb3.

1 (3

4.8)

0.6

(40.

8)3.

1 (3

1.6)

0.2

(44.

8)1.

7 (3

2.5)

1.1

(37.

2)

Cre

n’b

0.04

(0.5

)0.

01 (0

.5)

0.04

(0.4

)0.

01 (1

.1)

0.02

(0.3

)0.

01 (0

.4)

Tota

l9.

01.

69.

70.

55.

13.

0

CL-

GD

GTs

GD

GT-

012

.9 (4

4)2.

1 (4

3)36

.2 (4

6)5.

8 (3

9.3)

11.8

(45.

3)2.

9 (4

0.8)

GD

GT-

11.

1 (3

.7)

0.2

(3.6

)2.

9 (3

.7)

0.5

(3.5

)0.

9 (3

.5)

0.2

(3.5

)

GD

GT-

20.

5 (1

.6)

0.1

(1.6

)1.

3 (1

.6)

0.2

(1.4

)0.

4 (1

.5)

0.1

(1.5

)

GD

GT-

30.

2 (0

.8)

0.04

(0.7

)0.

6 (0

.8)

0.1

(0.7

)0.

2 (0

.7)

0.05

(0.7

)

Cre

n14

.5 (4

9.5)

2.5

(50.

5)37

.5 (4

7.6)

8.1

(54.

8)12

.7 (4

8.8)

3.7

(53)

Cre

n’0.

1 (0

.4)

0.03

(0.5

)0.

3 (0

.4)

0.04

(0.3

)0.

1 (0

.2)

0.04

(0.6

)To

tal

29.2

4.9

78.8

14.8

26.0

7.1

a Val

ues b

etw

een

pare

nthe

ses c

orre

spon

d to

the

frac

tiona

l (re

lativ

e) a

bund

ance

of

each

indi

vidu

al G

DG

T as

the

conc

entra

tion

of th

e in

divi

dual

GD

GT

divi

ded

by th

e su

m o

f th

e co

ncen

tratio

n of

all

GD

GTs

in th

at fr

actio

n. b

Cre

n =

Cre

narc

haeo

l, C

ren’

= c

rena

rcha

eol r

egio

isom

er.

Page 163: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

155

Results Ta

ble

3: A

bsol

ute

abun

danc

e as

resp

onse

uni

ts p

er g

ram

of

dry

wei

ght o

f de

tect

ed a

rcha

eal G

DG

T IP

Ls in

surf

ace

sedi

men

ts (0

–1cm

).

GD

GT

IPL

Stat

ion

1St

atio

n 2

Stat

ion

3

Mar

chA

ugus

tM

arch

Aug

ust

Mar

chA

ugus

t

GD

GT-

0-M

H9.

4×10

7 1.

3×10

7 3.

1×10

8 4.

4×10

7 1×

108

4.9×

107

GD

GT-

0-D

H6.

3×10

7 n.

d.7.

3×10

7 n.

d.6.

7×10

7 n.

d.

GD

GT-

0-D

H’

2×10

8 n.

d.2.

1×10

8 n.

d.1.

4×10

8 2.

6×10

7

GD

GT-

0-H

PH1.

7×10

10

3.3×

108

1.9×

1010

8.

5×10

8 1.

4×10

101.

8×10

9

GD

GT-

1-M

H9.

4×10

7 n.

d.8.

9×10

6 n.

d.1.

2×10

6 n.

d.

GD

GT-

1-D

H1.

9×10

8 n.

d.2.

1×10

8 n.

d.6.

2×10

7 n.

d.

GD

GT-

1-D

H’

n.d.

n.d.

n.d.

n.d.

8.2×

107

n.d.

GD

GT-

1-H

PH4.

6×10

8 n.

d.3.

4×10

8 n.

d.5.

2×10

8 n.

d.

GD

GT-

2-D

H1.

4×10

8 n.

d.2.

7×10

8 n.

d.1.

1×10

8 n.

d.

GD

GT-

3-D

H8×

106

n.d.

2.7×

107

n.d.

1×10

7 n.

d.

GD

GT-

3-D

H’

n.d.

n.d.

1.2×

107

n.d.

n.d.

n.d.

GD

GT-

4-D

H7.

8×10

7 n.

d.1.

5×10

8 n.

d.5×

107

n.d.

GD

GT-

4-D

H’

n.d.

n.d.

n.d.

n.d.

1.3×

107

n.d.

Cre

narc

haeo

l-MH

1.5×

108

5.3×

107

3.1×

108

5.7×

107

9.5×

107

7×10

7

Cre

narc

haeo

l-HPH

1.1×

1010

2.4×

108

8.5×

109

6.1×

108

6.6×

109

1.5×

109

Arc

haeo

l-DH

3.8×

108

n.d.

n.d.

n.d.

n.d.

n.d.

Tota

l ru

g–13.

0×10

108.

7×10

83.

0×10

101.

6×10

92.

2×10

103.

4×10

9

’isom

ers o

f sp

ecifi

c G

DG

Ts; n

.d.:

not d

etec

ted.

MH

: mon

ohex

ose,

DH

: Dih

exos

e, H

PH, h

exos

e ph

osph

ohex

ose.

Page 164: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

156

Tabl

e 4:

Arc

haea

l cla

sses

abu

ndan

ces (

cells

g–1

of

sedi

men

t) ca

lcul

ated

by

mul

tiply

ing

the

perc

enta

ges o

f to

tal a

rcha

eal 1

6S rR

NA

gen

e re

ads b

y th

e ar

chae

al 1

6S rR

NA

gen

e ab

unda

nce

(cop

y nu

mbe

r g–1

of

sedi

men

t) de

tect

ed in

the

surf

ace

sedi

men

t (0–

1cm

) of

the

thre

e st

atio

ns in

Mar

ch a

nd in

A

ugus

t, as

sum

ing

one

16S

rRN

A g

ene

copy

num

ber p

er g

enom

e.

orga

nism

Stat

ion

1St

atio

n 2

Stat

ion

3

Mar

chA

ugus

tM

arch

Aug

ust

Mar

chA

ugus

t

The

rmop

lasm

ata,

20a-

9n.

dn.

dn.

d7.

8 ×

108

n.d

2.0

×10

8

The

rmop

lasm

atal

es, A

MO

S1A

-411

3-D

04n.

dn.

d5.

5 ×

107

6.0

×10

74.

1 ×

108

6.0

×10

8

The

rmop

lasm

atal

es, C

CA

47n.

dn.

d1.

7 ×

108

2.4

×10

82.

3 ×

108

1.7

×10

7

The

rmop

lasm

atal

es, M

BG-D

& D

HV

EG

-1n.

d1.

1 ×

108

6.1

×10

87.

8 ×

108

1.2

×10

82.

0 ×

108

The

rmop

lasm

atal

es, V

C2.

1 A

rc6

n.d

n.d

1.1

×10

83.

0 ×

108

5.8

×10

71.

2 ×

109

Sum

of

read

s The

rmop

lasm

atal

esn.

d1.

1 ×

108

9.4

×10

81.

4 ×

109

8.2

×10

83.

7 ×

109

Misc

ella

neou

s Cre

narc

haeo

ta G

roup

, C3

9.9

×10

71.

1 ×

108

1.7

×10

82.

1 ×

109

n.d

9.0

×10

8

Tha

umar

chae

ota,

Mar

ine

Gro

up I,

Nitr

osop

umilu

s5.

8 ×

109

n.d

3.1

×10

92.

9 ×

109

2.2

×10

101.

2 ×

109

Misc

ella

neou

s Eur

yarc

haeo

tic G

roup

(ME

G)

n.d

n.d

n.d

1.4

×10

9n.

dn.

d

DPA

NN

, Woe

sear

chae

ota

DH

VE

G-6

1.1

×10

94.

2 ×

109

2.0

×10

91.

9 ×

1010

6.4

×10

91.

2 ×

1010

MBG

-Bn.

dn.

dn.

d9.

0 ×

108

n.d

1.0

×10

9

Oth

ers

4.9

×10

7n.

d2.

2 ×

108

6.0

×10

71.

5 ×

109

2.3

×10

9

Tota

l arc

haea

l abu

ndan

ce [c

ell g

–1]

7.1×

109

4.4×

109

6.5×

109

2.8×

1010

3.1×

1010

2.1×

1010

Tota

l CL-

GD

GTs

and

IPL-

deriv

ed G

DG

T m

g g–1

38.2

6.5

88.5

15.3

31.1

10.1

Cal

cula

ted

fem

togr

ams o

f G

DG

T pe

r cel

l1.

37.

53.

09.

81.

42.

9n.

d =

not

det

ecte

d. M

BG-D

, Mar

ine

Bent

hic

Gro

up D

; MBG

-B, M

arin

e Be

nthi

c G

roup

B.

Page 165: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

157

Results

Figu

re 1:

(A) P

erce

ntag

es o

f to

tal a

rcha

eal 1

6S rR

NA

gen

e rea

ds [%

] and

(B) a

rcha

eal 1

6S rR

NA

gen

e abu

ndan

ce [c

opy n

umbe

r g–1

of

sedi

men

t] de

tect

ed

in th

e su

rfac

e se

dim

ent (

0–1c

m) o

f th

e th

ree

stat

ions

in M

arch

and

in A

ugus

t. O

nly

arch

aeal

gro

ups w

ith p

erce

ntag

es >

3% ar

e re

port

ed. q

PCR

cond

ition

s, ef

ficie

ncy

= 9

0%; R

2=0.

991.

Page 166: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

158

March. In March both IPLs with GDGT-0 and crenarchaeol cores were mostly found with hexose phosphohexose (HPH) as IPL type, while in August the IPL type MH increased at the expense of HPH (Table 3).

6.3.3. Archaeal 16S rRNA gene diversity and abundanceArchaeal diversity was estimated by 16S rRNA gene amplicon sequencing using universal prim-ers. The archaeal community was dominated by Thaumarchaeota marine group I (MGI) in March both in station 1 and 3 with 72–82% of the total archaeal reads (Fig. 1A, Table S1). In station 2 these archaea were slightly less dominant but still comprised almost half of the total archaeal reads. Other archaeal groups were also present in the surface sediment in March in addition to MGI such as members of the DPANN Woesearchaeota (DHVE-6) (16–31%), Thermoplasmatales (3–14%), and MCG (1.4–2.5%). In August, the percentage of reads attribut-ed to the DPANN Woesearchaeota increased notably from 16 to 95% in station 1, from 31 to 66% in station 2, and from 21 to 57% in station 3 at the expense of the reads assigned to the Thaumarchaeota MGI (Fig. 1A, Table S1). In addition, a slight increase of percentage of reads was observed from March to August for archaea affiliated to the MCG and Thermoplasmatales groups. Total archaeal abundance estimated by qPCR with general archaea 16S rRNA gene prim-ers was comparable between March and August in all stations with slightly lower values in station 1 (average 6 × 109 gene copies g–1) and 1–2 × 1010 gene copies g–1 for station 2 and 3 (Fig. 1B).

The diversity of Thaumarchaeota was further analyzed by amplification, cloning and sequenc-ing of the amoA gene in the surface sediments of the three stations. All amoA sequences were closely related to sequences previously detected in marine surface sediments. Although certain variability was detected between the sequences recovered in different stations and seasons, no clear clustering was observed (Figs. S2A, B).

6.4. DiscussionThe seasonal changes in the oxygen and sulfide concentration in the bottom water of the ma-rine Lake Grevelingen trigger an important switch in the archaeal community composition in the surface sediment. The archaeal community changed considerably from a Thaumarchaeota MGI-dominated community in March, when oxygen was still present and sulfide was not (Table 1) to an archaeal community dominated by archaea of the DPANN superphylum. Besides, the seasonal change did not induce a diversity change within the Thaumarchaeota as indicated by the amoA gene phylogeny (Figs. S2A, B).

By multiplying the percentage of 16S rRNA gene reads of the Thaumarchaeota and the DPANN Woesearchaeota by the total archaea 16S rRNA gene copies per gram of sediments, we can estimate the abundances of these groups (assuming one 16S rRNA gene copy number per genome) (Table 4). For Thaumarchaeota, the abundance decreased dramatically in station 1 (from 6 × 109 cells g–1 to undetected) and in station 3 (2 × 1010 to 1.2 × 109 cells g–1), while in station 2 the abundance remained fairly constant (Table 4). On the other hand, members of the DPANN Woesearchaeota were the major component of the archaeal population in August in all stations with absolute numbers increasing in all stations (1 × 109 to 4.2 × 109 cells g–1 of sediment in station 1; 2 × 109 to 2 × 1010 cells g–1 in station 2, 6.4 × 109 to 1.2 × 1010 cells g–1 in station 3; Table 4). This change in the archaeal community is entirely compatible with the metabolism of both Thaumarchaeota and the DPANN archaea, which is aerobic (Könneke

Page 167: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

159

Results/Discussion

et al., 2005) and anaerobic (Castelle et al., 2015), respectively. Also, nitrification conducted by Thaumarchaeota has been shown to be inhibited by sulfide (Berg et al., 2015), which additionally explains the decrease in the MGI population upon increase of the sulfide concentration in the summer. In the same way, the abundance of the other archaeal benthic groups, which increased in the surface sediments in August, such as the Thermoplasmatales and the MCG, are predicted to have an anaerobic metabolism based on their genome. The environmental control of the change in the archaeal community diversity is also supported by the fact that this change is less evident in the surface sediment of station 3, where oxygen was still present and the sulfide con-centration in summer was substantially lower than at stations 1 and 2 (Table 1). Although the seasonality changes in the physicochemical conditions in the bottom and pore water induced a large change in the archaeal community diversity, the total archaeal abundance remained fairly constant between seasons and in the different stations (Fig. 1B). The archaeal community analy-sis was conducted by primer-based 16S rRNA gene amplicon sequencing. Therefore, we cannot completely rule out a possible primer bias (Sipos et al., 2010; Schloss et al., 2011), although the underrepresentation of DPANN sequences in the databases would imply that if any those prim-er biases would negatively detect DPANN in our samples, which in that case would induce to a underestimation of this group.

The large change in the archaeal community composition in the surface sediments upon the decrease of oxygen concentration in the summer coincided with an order of magnitude decrease of the total IPL-GDGTs detected. The IPL-GDGT profiles in March were compatible with a Thaumarchaeota-dominated population, due to the relatively high abundance of crenarchaeol, the specific CL of Thaumarchaeota (Sinninghe Damsté et al., 2002). In addition, the IPL type of crenarchaeol (and all other main GDGTs) switched from mostly HPH to MH from March to August (Table 3). The predominance of HPH IPL-crenarchaeol has been previously interpreted as an indication of the presence of an active Thaumarchaeotal population synthesizing mem-brane lipids in situ (Lengger et al., 2012, 2014), due to the labile nature of HPH IPLs (Harvey et al., 1986; Schouten et al., 2010). On the other hand, glycolipid-type IPLs (here mostly MH) have been considered as fossil signal (Lengger et al., 2012, 2014) or, alternatively, as the preferred IPL type in conditions where the Thaumarchaeota are in a stationary phase of growth (Elling et al., 2014). This latter explanation is not likely in the case where the surface sediment became fully sulfidic.

Recent studies suggest that members of the DPANN have a reduced genome with limited metabolic capabilities (Rinke et al., 2013; Castelle et al., 2015), suggesting that these archaea may have a symbiotic or parasitic lifestyle. In addition, they also lack most if not all the genes coding for the enzymes of the archaeal lipid biosynthetic pathway (Jahn et al., 2004; Villanueva et al., 2017). They are, therefore, not expected to contribute to the total IPL-GDGT pool by actively synthesizing lipids but may recycle the membrane lipid of other archaea. The decline of the total archaeal IPL-GDGTs in August coinciding with the switch to a DPANN-dominant population with similar or even higher cell abundances to that reported in March would therefore imply that the DPANN Woesearchaeota make use of the preserved (‘fossil’) pool of IPL-GDGTs to make their membranes. A question with respect to this membrane lipid acquisition mechanism is whether the pool of ‘fossil’ IPL-GDGTs is able to fulfill the membrane requirements of the large archaeal DPANN population detected in the surface sediments sampled in August in view of the much reduced concentration of IPL-GDGTs, which is an order of magnitude lower. This

Page 168: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

160

can only work when the DPANN archaea have a much smaller cell size than the Thaumarchae-ota that thrive in spring. The size of the DPANN Woesearchaeota present in Lake Grevelingen surface sediment is unknown. The only characterized archaeon from the DPANN superphylum is Nanoarchaeum equitans, which is considered to be a symbiont of the archaeon Ignicoccus sp. by growing on its surface. The lipids of N. equitans have been shown to be derived from Ignicoccus sp. (Jahn et al., 2004), but the uptake mechanisms remain unknown. N. equitans is 5 times smaller than its host, the archaeon Ignicoccus sp.. Assuming a similar size difference between the DPANN Woesearchaeota found in the Lake Grevelingen surface sediments and its potential host, suffi-cient IPLs would be available to sustain the DPANN population detected in the sediments in August.

Alternatively, the DPANN archaea could recycle the CL-GDGTs present in the sediment as previously suggested (Liu et al., 2011; Takano et al., 2010), which is presumed to involve hydro-lysis and reformation of the ether bond to the glycerol backbone as suggested by the 13C-glucose labeling studies of Takano et al. (2010). However, since the members of the DPANN generally lack the genes coding for the enzymes involved in producing the ether bonds to the glycerol backbone (Villanueva et al., 2017), this is unlikely. Alternatively, they could also use the CL-GDGTs and add the polar head groups de novo. Waters et al. (2003) observed the presence of genes for lipid modification, such as glycosylation, in the genome of N. equitans. We performed a search for the gene encoding cytidine diphosphate (CDP)-archaeol synthase (CarS, E.C. 2.7.7.67) that catalyzes the activation of 2,3-bis-O-geranylgeranylglyceryl diphosphate (DGGGP) by cyti-dine triphosphate (CTP) to form the intermediate for polar head group attachment (i.e. CDP-ar-chaeol), and we found homologs in most of the DPANN Micrarchaeota genomes and in some of the DPANN Parvarchaeota and Diapherotrites (perfomed by find function in JGI Integrated Microbial Genomes, IMG with genomes available in July 2017). However, the genes coding for the enzymes that catalyze the subsequent replacement of cytidine monophosphate of the cyt-idine diphosphate (CDP)-archaeol of CDP-diacylglycerol with a polar head group in DPANN genomes were not detected, in contrast to N. equitans (Waters et al., 2003). This may indicate that the members of the DPANN either do not have the capacity of adding the polar head groups to the CL-GDGT or that the polar head groups of their IPLs are different and added by enzymes different to those already characterized.

6.5. ConclusionWe observed a dramatic change in the archaeal community composition and lipid abundance in surface sediments of a seasonally hypoxic marine lake which corresponded to a switch from a Thaumarchaeota-dominated to a DPANN-dominated archaeal community, while the total IPLs detected were significantly reduced. Considering the reduced genome of the members of the superphylum DPANN and their apparent inability to synthesize their own membrane lipids, we hypothesize that they use the CLs previously synthesized by the Thaumarchaeota to form their membrane.

Acknowledgements

We acknowledge the crew of the R/V Luctor and Pieter van Rijswijk for their help in the field during sediment collection, Marcel van der Meer and Sandra Heinzelmann for support onboard

Page 169: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

161

Discussion/Conclusion/Supplementary Material

and assistance with incubations, Elda Panoto for technical support, and Eric Boschker and Filip Meysman for helpful discussions. This work was financially supported by the Darwin Centre for Biogeosciences to LV (grant no: 3062). LV and JSSD receive funding from the Soehngen Institute for Anaerobic Microbiology (SIAM) through a Gravitation Grant (024.002.002) from the Dutch Ministry of Education, Culture and Science (OCW).

Figure S1: Scheme of the sampling locations in Lake Grevelingen

Page 170: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 6

162

Tabl

e S1

: Per

cent

ages

of

tota

l arc

haea

l 16S

rRN

A g

ene

read

s [%

] det

ecte

d in

the

surf

ace

sedi

men

t (0–

1cm

) of

the

thre

e st

atio

ns in

Mar

ch a

nd in

Aug

ust.

Onl

y ar

chae

al g

roup

s with

per

cent

ages

>3%

are

repo

rted

.

orga

nism

Stat

ion

1St

atio

n 2

Stat

ion

3

Mar

chA

ugus

tM

arch

Aug

ust

Mar

chA

ugus

t

The

rmop

lasm

ata,

20a-

9n.

d.n.

d.n.

d.2.

7n.

d.0.

9

The

rmop

lasm

atal

es, A

MO

S1A

-411

3-D

04n.

d.n.

d.0.

80.

21.

32.

8

The

rmop

lasm

atal

es, C

CA

47n.

d.n.

d.2.

50.

80.

77.

9

The

rmop

lasm

atal

es, M

BG-D

& D

HV

EG

-1n.

d.2.

69.

32.

70.

40.

9

The

rmop

lasm

atal

es, V

C2.

1 A

rc6

n.d.

n.d.

1.7

1.1

0.2

5.6

Sum

of

read

s The

rmop

lasm

atal

esn.

d.2.

614

.44.

92.

617

.1

Misc

ella

neou

s Cre

narc

haeo

ta G

roup

, C3

1.4

2.6

2.5

7.4

n.d.

4.2

Tha

umar

chae

ota,

Mar

ine

Gro

up I,

Nitr

osop

umilu

s81

.9n.

d.48

.310

.172

.15.

6

Misc

ella

neou

s Eur

yarc

haeo

tic G

roup

(ME

G)

n.d.

n.d.

n.d.

5.1

n.d.

n.d.

DPA

NN

, Woe

sear

chae

ota

DH

VE

G-6

16.0

94.9

31.4

66.5

20.6

56.9

MBG

-Bn.

d.n.

d.n.

d.3.

2n.

d.4.

6

Oth

ers

0.7

n.d.

3.4

0.2

4.7

10.6

n.d

= n

ot d

etec

ted.

MBG

-D, M

arin

e Be

nthi

c G

roup

D; M

BG-B

, Mar

ine

Bent

hic

Gro

up B

.

Page 171: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

163

Supplementary Material

Lake Grevelingen sediment; 0-1 cm depth; 2 seq March: 1 x S2, 1 x S3; 6 seq August: 2 x S1, 2 x S2, 2 x S3

Lake Grevelingen sediment; 0-1 cm depth; March S2 uncultured crenarchaeote, Elkhorn Slough sediment (DQ148892)

Lake Grevelingen sediment; 0-1 cm depth; August S1 Lake Grevelingen sediment; 0-1 cm depth; August S2

Lake Grevelingen sediment; 0-1 cm depth; March S2 Lake Grevelingen sediment; 0-1 cm depth; August S3

Lake Grevelingen sediment; 0-1 cm depth; 5 seq March: 2 x S1, 2 x S2, 1 x S3 Lake Grevelingen sediment; 0-1 cm depth; March S3

Lake Grevelingen sediment; 0-1 cm depth; August S2 Lake Grevelingen sediment; 0-1 cm depth; 4 seq March: 1 x S1, 3 x S2; 5 seq August: 1 x S1, 2 x S2, 2 x S3 Lake Grevelingen sediment; 0-1 cm depth; 2 seq March: 1 x S1, 1 x S3; 2 seq August: 1 x S1, 1 x S3

Lake Grevelingen sediment; 0-1 cm depth; March S3 Lake Grevelingen sediment; 0-1 cm depth; August S2

uncultured crenarchaeote clone, Elkhorn Slough sediment (DQ148771) uncultured archaeon clone, aquarium biofilter (AB373356) Lake Grevelingen sediment; 0-1 cm depth; March S1

Lake Grevelingen sediment; 0-1 cm depth; March S3 uncultured archaeon, marine aquaculture (AM295172)

uncultured crenarchaeote clone, marine sediment (FJ656584) Nitrosopumilus maritimus (SCM1) (CP000866) Lake Grevelingen sediment; 0-1 cm depth; 4 seq March: 2 x S1, 2 x S3

Lake Grevelingen sediment; 0-1 cm depth; March S3 Lake Grevelingen sediment; 0-1 cm depth; March S1

Lake Grevelingen sediment; 0-1 cm depth; August S3 Lake Grevelingen sediment; 0-1 cm depth; August S1

Lake Grevelingen sediment; 0-1 cm depth; 4 seq March: 1 x S1, 2 x S2, 1 x S3; 4 seq August: 1 x S2, 3 x S3 Lake Grevelingen sediment; 0-1 cm depth; March; S1

Lake Grevelingen sediment; 0-1 cm depth; March; S1 Lake Grevelingen sediment; 0-1 cm depth; August S1

uncultured archaeon clone, deep sea sediments (AB289355) uncultured crenarchaeote clone, Black Sea water 100 m (EF414231) Lake Grevelingen sediment; 0-1 cm depth; August S3

Lake Grevelingen sediment; 0-1 cm depth; March: 1 x S1; 6 seq August: 3 x S1, 2 x S2, 1 x S3

Lake Grevelingen sediment; 0-1 cm depth; 20 seq March: 5 x S1, 8 x S2, 7 x S3; 24 seq August: 5 x S1, 12 x S2, 7 x S3

Lake Grevelingen sediment; 0-1 cm depth; March; S2 Lake Grevelingen sediment; 0-1 cm depth; March; S1

Lake Grevelingen sediment; 0-1 cm depth; August S1 Lake Grevelingen sediment; 0-1 cm depth; March S2

Lake Grevelingen sediment; 0-1 cm depth; March S3 Lake Grevelingen sediment; 0-1 cm depth; 3 seq March: 1 x S1, 1 x 2, 1 x S3; 3 seq August: 2 x S1, 1 x S3

Lake Grevelingen sediment; 0-1 cm depth; August S2 Lake Grevelingen sediment; 0-1 cm depth; August S3

uncultured archaeon clone, deep sea sediments (AB289348)uncultured thaumarchaeota clone, Arabian Sea surface sediment (KJ509763)

uncultured archaeon clone (deep sea water column) (AB289377) Lake Grevelingen sediment; 0-1 cm depth; August S3

Lake Grevelingen sediment; 0-1 cm depth; 3 seq March: 1 x S1, 1 x S2, 1 x S3; 3 seq August: 2 x S1, 1 x S3

Lake Grevelingen sediment; 0-1 cm depth; 8 seq March: 2 x S1, 2 x S2, 4 x S3; 8 seq August: 2 x S1, 4 x S2, 2 x S3

Lake Grevelingen sediment; 0-1 cm depth; March S2 uncultured crenarchaeote clone, San Francisco Bay sediment (DQ148664)

Lake Grevelingen sediment; 0-1 cm depth; 2 seq August S3 Lake Grevelingen sediment; 0-1 cm depth; 3 seq March: 1 x S1, 1 x S2, 1 x S3

uncultured crenarchaeote clone, wastewater (HM589775) uncultured crenarchaeote clone, wastewater (HM191510) uncultured archaeon clone, freshwater aquaria (JN183601)53

96

73

53

0.02

Figure S2A: Phylogenetic tree of amoA protein sequences recovered from the surface sediment (0–1 cm) of Stations 1, 2 and 3 in Lake Grevelingen during March and August (black bold: sequences retrieved in this study and closest relatives) constructed with the Neighbor-Joining method (Saitou and Nei, 1987). Scale bar indicates 2% sequence dissimilarity. The evolutionary distances were computed using the Poisson correc-tion method with a bootstrap test of 1,000 replicates (values higher than 50% are shown on the branches).

Page 172: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

164

Chapter 6 Supplementary Material

uncultured crenarchaeote clone, Gulf of California (EU340510) uncultured crenarchaeote clone, Black Sea water 70 m (DQ148740) uncultured archaeon clone, Peruvian OMZ (FJ799200)

uncultured crenarchaeote clone, South China Sea 100m (EU885994) uncultured crenarchaeote clone, North Pacific subtropical gyre 130 m (EF106932)

uncultured crenarchaeote clone, tropical marine estuarine sediment (FJ601621) uncultured crenarchaeote clone, Elkhorn Slough sediments (DQ148805) uncultured crenarchaeote clone, ETNP OMZ 200 m (DQ148759)

uncultured crenarchaeote clone, intertidal sandy flat (EU099940) uncultured ammonia-oxidizing archaeon clone, salt marsh sediment (EU925267)

Candidatus Nitrosoarchaeum limnia (SFB1) (CM001158) uncultured crenarchaeote clone (Monterey Bay 30 m) (DQ148820)

uncultured crenarchaeote clone, Changjiang Estuary sediment (EU025183) uncultured Thaumarchaeota clone, Arabian Sea 170 m (KF512326)

uncultured archaeon clone, Peruvian OMZ (FJ799198) uncultured crenarchaeote clone, Pacific Ocean 150 m (GU364066)

Lake Grevelingen sediment; 0-1 cm depth; March S1 Lake Grevelingen sediment; 0-1 cm depth; August S2

uncultured crenarchaeote clone, San Francisco Bay sediment (DQ148670.1) Lake Grevelingen sediment; 0-1 cm depth; March S1 Lake Grevelingen sediment; 0-1 cm depth; March S2

uncultured crenarchaeote clone, Elkhorn Slough South Marsh (FJ227870) uncultured archaeon clone, aquarium biofilter (AB373264)

Cenarchaeum symbiosum (A) (DQ397569) uncultured crenarchaeote clone, San Francisco Bay sediment (DQ148654)

Candidatus Nitrosoarchaeum koreensis (My1) (HQ331117) uncultured crenarchaeote clon, Gulf of California (DQ148574)

uncultured crenarchaeote clone, Elkhorn Slough sediment (DQ148796) uncultured crenarchaeote clone, San Francisco Bay sediment (DQ148671) uncultured archaeon clone, hydrothermal vents (AB451499)

uncultured crenarchaeote clone, Gulf of California (EU340541) uncultured crenarchaeote clone, North pacific subtropical gyre 4000 m (EF106928) uncultured crenarchaeote clone, Black Sea 70 m (DQ148697) uncultured archaeon, sponge tissue (DQ333425)

uncultured crenarchaeote clone, ETNP OMZ 200 m (DQ148764) uncultured crenarchaeote, ETNP OMZ 200 m (DQ148750)

uncultured crenarchaeote clone, North East Atlantic 2502 m (EU810223) uncultured crenarchaeote clone, North East Atlantic 2502 m (EU810225)

uncultured crenarchaeote clone, ETNP OMZ 200 m (DQ148769) uncultured archaeon, sponge tissue (DQ333419) uncultured thaumarchaeota clone, Arabian Sea 1050 m (KF512379)

uncultured crenarchaeote clone, Gulf of California (EU340549) uncultured crenarchaeote clone, ETNP OMZ 200 m (DQ148746)

uncultured crenarchaeote clone, Gulf of Mexico (GQ250723) uncultured crenarchaeote clone, Juan de Fuca Ridge 2267 m (EU864296)

uncultured crenarchaeote clone, North Pacific subtropical gyre 700 m (EF106899) uncultured crenarchaeote clone, ETNP OMZ 200 m (DQ148762)

uncultured archaeon clone, Lake water (FN773449) uncultured archaeon clone, Lake water (FN773418)

uncultured archaeon clone, Lake water (FN773404) uncultured crenarchaeote clone, rhizosphere (EU667867)

Candidatus Nitrosotalea devanaterra (JN227489) uncultured crenarchaeote clone, forest soil (EF530119)

Candidatus Nitrosocaldus yellowstonii (EU239961) Candidatus Nitrososphaera vienensis (FR773159)

Candidatus Nitrososphaera gargensis (EU281321) uncultured crenarchaeote clone, vapour cave (DQ672686)99

97

78100

99

6353

54

69

55

99

6794

51

83

51

57

87

76

86

54

81

0.02

Figure S2B: Phylogenetic tree of amoA protein sequences recovered from the surface sediment (0–1 cm) of Stations 1, 2 and 3 in Lake Grevelingen during March and August (black bold: sequences retrieved in this study and closest relatives) constructed with the Neighbor-Joining method (Saitou and Nei, 1987). Scale bar indicates 2% sequence dissimilarity. The evolutionary distances were computed using the Poisson correc-tion method with a bootstrap test of 1,000 replicates (values higher than 50% are shown on the branches).

Page 173: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 174: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 175: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

167

7. Synthesis

The role of chemolithoautotrophic microorganisms has been considered to be of minor impor-tance in coastal marine sediments although it has not been investigated in depth. Additionally, the impact of seasonal hypoxic/anoxic conditions on microbial chemolithoautotrophy in coastal marine sediments has not been examined. Therefore, in this thesis the diversity, abundance and activity of specific groups of chemolithoautotrophic microorganisms involved in the nitrogen and sulfur cycling, such as ammonia oxidizing archaea and bacteria, denitrifying and anammox bacteria, sulfate reducing and sulfur oxidizing bacteria, as well as the total microbial communi-ty, has been determined in coastal sediments by using DNA/RNA- and lipid-based biomarker approaches. Results show the presence and potential importance of chemolithoautotrophs in the biogeochemical cycling of carbon, nitrogen and sulfur. Additionally, temporal changes of the diversity, abundance and activity of chemolithoautotrophic microorganisms coinciding with seasonal hypoxia in coastal sediments have been determined.

7.1. Physicochemical factors of the sediments harboring chemolithoautotrophs studied in the framework of this thesisIn this thesis, coastal sediments with different sediment properties and oxygen conditions were examined. In this section, I focus on the most important environmental factors affecting chem-olithoautotrophic microorganisms targeted in this thesis, i.e. aerobic (ammonia oxidizing archaea and bacteria, AOA and AOB) and anaerobic ammonia oxidizing (anammox) bacteria, denitrify-ing bacteria, as well as sulfate reducing and sulfur oxidizing bacteria including sulfide-dependent denitrifiers. Effects on the seasonal and spatial distribution of the diversity, abundance and ac-tivity of the target organisms, as well as applied methods are discussed in detail in the following sections.

Sediments of the Oyster Grounds (Chapter 2), a deep depression in the central North Sea and a temporary deposition center for sediment, play an important role in the carbon and nitrogen cycle in the region (van Raaphorst et al., 1998; Weston et al., 2008). High levels of benthic fauna has been reported for this area (Duineveld et al., 1991; van Raaphorst et al., 1992) even during summer stratification. Oxygen concentrations in the water column commonly decrease from winter to summer (as seen in our case between the sampling performed in February and August), as well as the oxygen penetration depth (OPD) into the sediment. In this thesis, sediments sam-pled in February, May and August were investigated.

Also Lake Grevelingen sediments (Chapters 4, 5, and 6) were exposed to varying oxygen concntrations, i.e. bottom water hypoxia due to summer stratification of the water column, and sulfide accumulation in the sediment pore water due to intense sulfate reduction. Sediments, lo-cated at different water depths of a deep basin within Lake Grevelingen were investigated (S1 at 34 m, S2 at 23 m, and S3 at 17m). Bottom waters were fully oxygenated in March (S1-S3), where-as in August oxygen was strongly depleted in the bottom waters and no oxygen has been detect-ed in the surface sediments (S1, S2). Bottom water oxygen concentration at S3 remained higher in August (~88 μM), and oxygen penetrated down to 1.1 mm into the sediment. In August, free sulfide was present near the sediment-water interface at all three stations, and the accumulation of sulfide in the sediment pore water increased with water depth (S3 to S1).

Page 176: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

168

In general, my results support that the oxygen concentration of the bottom water and the OPD into the sediment is one of the major factors shaping the temporal and spatial distribution of the diversity, abundance and activity of chemolithoautotrophic microorganisms, due to dif-ferent oxygen requirements of their respective metabolism. Next to the availability of electron donors and acceptors, I have concluded that other factors such as bioturbation (Kristensen, 2000; Meyer et al., 2005; Beman et al., 2012) and the interactions between different microbial groups are important factors allowing for the activity of aerobic and anaerobic microorgan-isms together in the same niche. The abundance and activity of Thaumarchaeota and anammox bacteria was detected in Oyster Grounds sediments (Fig. 1) independent of the season and decreasing oxygen concentrations in the bottom water, whereas in Lake Grevelingen sediments, the abundance and activity of Thaumarchaeota was restricted to March, when the bottom water was still oxygenated (Fig. 2). These results suggest that the presence of bioturbation, oxygen-ating deeper layers in Oyster Grounds sediments, favors the presence and activity of aerobic ammonia oxidizing microorganisms, which could in turn fuel the activity of anammox bacteria by consuming oxygen and providing nitrite (Meyer et al., 2005; Laverock et al., 2013), compared to Lake Grevelingen sediments, where bioturbation was absent and sulfide accumulates in the sediment pore water during summer hypoxia.

In comparison to Oyster Grounds sediments, temporal and spatial differences of the com-munity structure, as well as of the abundance and activity of studied microorganisms were more pronounced in Lake Grevelingen sediments (Fig. 2), most likely due to the full development of hypoxia and rising sulfide concentrations as it has been shown recently for Black Sea sediments (Jessen et al., 2017). High sulfide concentrations are not only toxic for benthic bioturbating an-imals as sulfide has also been shown to inhibit ammonia oxidation (Joye and Hollibaugh, 1995; Berg et al., 2015) and the last steps of denitrification (Sørensen, 1980; Porubsky et al., 2009). My results showed that decreasing oxygen and rising sulfide concentrations induce drastic changes in the archaeal and bacterial community structure, and limit the abundance and/or the activity as well of specific groups involved in the sedimentary nitrogen and sulfur cycle, i.e. Thaumarchae-ota (AOA), AOB, anammox (Chapters 2, 3, 4) and denitrifying bacteria, sulfate and sulfur oxi-dizing bacteria including sulfide-dependent denitrifiers in Lake Grevelingen sediments (Chapters 4 and 6), as well as of chemolithoautotrophic bacteria (Chapter 5). Additionally, the potential activity of anammox bacteria was relatively low in Lake Grevelingen sediments compared to Oyster Grounds sediments, suggesting that this group is strongly affected by increasing sulfide concentrations. The activity of other groups, i.e. denitrifying bacteria, sulfate reducing and sul-fur oxidizing bacteria including sulfide-dependent denitrifiers, was relatively low in seasonally hypoxic Lake Grevelingen sediments independent of the season, suggesting that the activity of these groups is as well affected by decreasing oxygen and rising sulfide concentrations between spring and summer.

Both anammox and denitrification pathways have been suggested to be more important in sediments with low sulfide concentrations, but also with low carbon input (Burgin and Ham-ilton, 2007). The TOC content in Lake Grevelingen sediments is high (2.5−4.5%) and fresh organic rich sediment rapidly accumulates (> 2 cm y−1), which might additionally inhibit the activity of anammox and denitrifying bacteria. The TOC content in Oyster Grounds sediments was low (0.15−0.38%), which might explain the higher activity of anammox activity compared to Lake Grevelingen sediments. However, sediment incubation experiments (using labeled ni-

Page 177: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

169

Synthesis

trite) have previously shown that the ratio of anammox to denitrification depends on the carbon to nitrogen (C/N) ratio, i.e. the activity of anammox bacteria increased with higher nitrogen relative to carbon content (lower C/N) ratio of the source substrate (Babbin et al., 2014). How-ever, in my studies, I have observed a difference in the potential activity of anammox bacteria with higher expression levels in Oyster Ground compared to Lake Grevelingen sediments. The C/N ratio is relatively high in sediments of the Oyster Ground (7−8) and Lake Grevelingen (9.2 ± 0.2; Rao et al., 2016), indicating that this does potentially not explain the differences in the potential activity of anammox bacteria in the sediments studied here. Generally, the activity of chemolithoautotrophic bacteria in Lake Grevelingen sediments was restricted to March un-der oxygenated bottom water concentrations, indicating that chemolithoautotrophic bacteria in general are affected by bottom water hypoxia/anoxia, the lack of bioturbation and rising sulfide concentrations (Chapter 5).

Additionally, the abundance and metabolic capacities of the chemolithoautotrophic aerobic ammonia oxidizing archaea (AOA), i.e. Thaumarchaeota, were also evaluated in other marine sediment systems, i.e. in sandy to muddy sediments of the Iceland shelf. Thaumarchaeota were detected at all stations but their metabolic activity was low based on lack of incorporation of 13C-labeled substrate in archaeal GDGT lipids. The low activity of Thaumarchaeota is in agree-ment with the low oxygen penetration into the sediment, supporting that the absence of oxy-gen is limiting the activity of Thaumarchaeota. Ammonia and nitrite concentrations showed no changes during incubations, indicating that other factors than the availability of ammonia might inhibit the activity of Thaumarchaeota.

7.2. Diversity of chemolithoautotrophic microorganisms in coastal sedimentsThe general diversity of the microbial community was studied in Lake Grevelingen surface sed-iments by applying 16S rRNA amplicon sequencing analysis (Chapters 5, 6), before and during summer hypoxia. In general, these sediments harbor a distinct bacterial community comparable to other surface sediments, such as coastal marine (North Sea), estuarine and Black Sea sedi-ments (Jessen, et al., 2016; Ye et al., 2016; Dyksma et al., 2016). Approximately, 50% of the bac-terial reads were assigned to the clades of Gammaproteobacteria, Deltaproteobacteria and Ep-silonbacteria, whereas the remaining reads were distributed among the orders of Bacteroidetes (14%), Planctomycetes (6%), Alphaproteobacteria (2%) and other orders (22%). The majority of reads assigned to the Gammaproteobacteria were closely related to the orders Alteromo-modales, Chromatiales and Thiotrichales, whereas most of the reads of Epsilonproteobacteria were assigned to the order of Campylobacterales. Within the Deltaproteobacteria, most reads were assigned to the orders of Desulfarculales and Desulfobacterales.

In March, under oxygenated bottom water conditions, the surface sediment was characterized by high relative abundances (percentage of 16S rRNA gene reads) of Gammaproteobacteria and Epsilonproteobacteria, which are known chemolithoautotrophic sulfur oxidizers, e.g. ca-pable of oxidizing sulfide, thiosulfate, elemental sulfur and polythionates, using oxygen and nitrate as electron acceptor. In August, the lack of electron acceptors in the bottom water, i.e. oxygen and nitrate, coincided with a decrease in the relative abundance of Gammaproteobac-teria and Epsilonproteobacteria. At the same time, the number of reads assigned to the order of Desulfobacterales increased in the top centimeter of the sediment and a shift in the bacterial

Page 178: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

170

community toward chemoorganoheterotrophic sulfate-reducing bacteria related to the genera Desulfococcus and Desulfosarcina was evident. In coastal sediments, these genera are characteristic in deeper sediment layers experiencing anaerobic mineralization (Llobet-Brossa et al., 2002; Mi-yatake, 2011; Muyzer and Stams, 2008). The DNA-based 16S rRNA gene amplicon sequencing was complemented with the analysis of bacterial phospholipid derived fatty acids (PLFA) anal-ysis in Lake Grevelingen sediments down to 6 cm sediment depth in all seasons and stations. In March, surface sediments were characterized by relative high concentrations of C16 and C18 monosaturated PLFAs found in Gammaproteobacteria and Epsilonproteobacteria (Vestal and White 1989), whereas in deeper sediment layers, ai15:0, i17:1ω7c, and Me16:0, indicating that sulfate-reducing Deltaproteobacteria were more abundant (Boschker et al., 2002; Taylor and Parks, 1983; Edlund et al., 1985). In August, surface sediment showed an increased contribution of iso, anteiso, and branched PLFAs, more similar to the deeper sediments from March, but showed higher abundances of 16:0, 14:0, and 18:1ω9c compared to March. These shifts in the PLFA patterns are in agreement with the temporal difference in the bacterial community shown by 16S rRNA gene amplicon sequencing. However, the results of the PLFA analysis have to be interpreted with caution because there is now evidence that the method used here (Guckert et al., 1985), i.e. the separation of phospholipids from glycolipids and neutral lipids using silica col-umn chromatography, does not result in a fraction only containing phospholipids but also other lipid classes (Heinzelmann et al., 2014). Thus, commonly reported PLFA composition might be not derived purely from phospholipids but from of other lipid classes.

Additionally, the archaeal diversity was estimated by 16S rRNA gene amplicon sequencing in surface sediments of Lake Grevelingen (Chapter 6). The archaeal community was dominated by chemolithoautotrophic Thaumarchaeota marine group I (MGI) in March (50−82%), when oxygen was still present and sulfide was absent, next to the presence of other archaeal groups, such as members of the DPANN Woesearchaeota, Thermoplasmatales and members of the Miscellaneous Crenarchaeotic group (MCG). In August, the percentage of reads, attributed to the DPANN Woesearchaeota increased notably (57−95%) at the expense of reads assigned to the Thaumarchaeota. This change in the archaeal community is compatible with the aerobic me-tabolism of Thaumarchaeota (Könneke et al., 2005) and the anaerobic metabolism of DPANN archaea (Castelle et al., 2015). Thaumarchaeota have been shown to be inhibited by sulfide (Berg et al., 2015), potentially explaining the decrease in the MGI population upon the increase of sulfide concentrations in the summer.

In summary, results of the community analysis revealed a diverse microbial community in Lake Grevelingen surface sediments consisting of chemolithoautotrophic and chemoorgano-hetero-trophic microorganisms, responding to changes in the availability of reduced nitrogen and sulfur compounds and electron acceptors (oxygen and nitrogen) induced by hypoxia. Nevertheless, the bacterial and archaeal community analysis was performed by 16S rRNA gene amplicon sequenc-ing, therefore I cannot completely exclude a possible primer bias (Sipos et al., 2010; Schloss et al., 2011). Additionally, the diversity analysis of the 16S rRNA gene does not give information about absolute abundances, the metabolism and more importantly, about the activity of sedimentary microorganisms and their role in biogeochemical cycling of carbon, nitrogen and sulfur.

To further estimate the diversity of microorganisms fixing CO2 via the Calvin-Benson-Bassh-am (CBB) and the reductive tricarboxylic acid (rTCA) pathways, surface sediments of Lake Grevelingen sediments (Chapter 5) were investigated by cloning and sequencing the cbbL and

Page 179: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

171

Synthesis

aclB gene, encoding for the key enzymes ribulose 1,5-bisphosphate carboxylase/oxygenase (Ru-bisCO) and ATP citrate lyase, respectively. All detected cbbL sequences clustered with uncultered gammaproteobacterial clones, assigned to the orders Chromatiales and Thiotrichales, reported in intertidal sediment from Lowes Cove, ME (Nigro and King, 2007). The aclB gene sequences in surface sediments were predominantly related to aclB sequences of Epsilonproteobacteria of the order Campylobacterales, macrofaunal symbionts and to sequences of Euryarchaeota. No clustering of cbbL and aclB gene sequences according to one of the stations or one of the seasons was observed. Results revealed a stable community of bacteria using the CBB cycle, as well as bacteria and archaea using the rTCA pathway in Lake Grevelingen surface sediments. Until today, there is no PCR-based approach available to identify microorganisms using other CO2 fixation pathways, such as the reductive acetyl-CoA, or Wood-Ljungdahl (WL) pathway utilized by Firmicutes, Planctomycetes, Deltaproteobacteria, Spirochaeta and Euryarchaeota, or the 3-hydroxypropionate/4-hydroxybutyrate (3-HP/4-HB) cycle and the dicarboxylate/4- hy-droxybutyrate (DC/4-HB) cycle operating in Crenarchaeota (Hügler and Sievert, 2011), mainly due to the diversity of key genes of these pathways which impairs the development of a PCR-based detection method. Thus, an underestimation of the diversity of chemolithoautotrophic microorganisms in the coastal marine sediments under study in this thesis is possible.

Another goal of this thesis was to further estimate the diversity of chemolithoautotrophic microorganisms involved in the sedimentary nitrogen and sulfur cycle by using group specific 16S rRNA genes and functional genes involved in reoxidation pathways. This approach was applied in sediments of the Oyster Grounds (North Sea) and Lake Grevelingen surface sedi-ments. In Oyster Grounds surface (0–1 cm) and deeper (9–10 cm) sediments, the diversity of ammonia oxidizing archaea (Thaumarchaeota, AOA) and ammonia oxidizing bacteria (AOB), was determined by targeting the functional amoA gene of AOA and AOB. Results showed that amoA gene sequences of AOA amplified from Oyster Grounds surface sediments fell into dis-tinct subclusters of Nitrosopumilus sp., whereas most of the archaeal amoA sequences obtained from Lake Grevelingen sediments were closely related to Nitrosopumilus maritimus (SCM1) and in minor proportion to “Candidatus Nitrosoarchaeum limnia” found in low-salinity habitats (Mosier et al., 2012), showing a different community composition of AOA in Lake Grevelingen surface sediments, which is most likely explained by the lower salinity (~30 psu) compared to Oyster Grounds sediments (~34.5 psu). AOB amoA gene sequences retrieved from Oyster Grounds sediments were closely related to Nitrospira sp. and Nitrosolobus multiformis.

Another important chemolithoautotrophic microbial group present in anoxic sediment and involved in the nitrogen cycle is anammox bacteria. One issue much in debate is the relevance of anammox bacteria vs. heterotrophic denitrifiers in the N2 removal process in the environment. In this thesis I tackled this issue by determining the diversity and role of anammox bacteria, heterotrophic denitrifiers and other microbial groups involved in nitrogen removal (e.g. sulfur oxidizers coupled to nitrate reduction) in Lake Grevelingen sediments (Chapter 4). The diversity of the anammox 16S rRNA gene, whereas the nirS gene of denitrifying and anammox bacteria, as well as the aprA gene of sulfur oxidizing and sulfate reducing bacteria, were determined in Lake Grevelingen surface sediments (chapter 4). The diversity of nirS-type denitrifiers in surface sediments (0−1 cm) revealed a temporally stable community of denitrifying bacteria involved in autotrophic denitrification coupled to sulfide or iron oxidation (e.g. Thiobacillus denitrificans, Sid-eroxidans lithotrophicus), and of diverse heterotrophic denitrifiers. Most of the anammox bacterial

Page 180: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

172

nirS gene sequences obtained from Lake Grevelingen surface sediments were closely related to “Candidatus Scalindua profunda”, whereas anammox bacterial 16S rRNA gene sequences were closely related to “Candidatus Scalindua marina” and “Candidatus Scalindua brodae” as found in Oyster Grounds sediments (0–1 cm and 9–10 cm sediment depth).

The analysis of the aprA gene also revealed a diverse sulfur oxidizing community; including affiliations to autotrophic sulfur-dependent denitrifying Beta- and Gammaproteobacteria (Thio-bacillus denitrificans and Sulfuricella denitrificans), chemolithoautotrophic sulfur-dependent nitrate reducing Betaproteobacteria (Thiobacillus thioparus), heterotrophic sulfate reducing Deltaproteo-bacteria (e.g. Desulfosarcina sp., Desulfobulbus propionicus) or phototrophic sulfur oxidizing bacteria (Lamprocystis purpurea). In addition, the diversity of genes involved in carbon fixation and reoxi-dation pathways were also estimated showing no clustering of sequences according to the depth, station or season. These results were interpreted as an indication of the presence of a stable population of ammonia oxidizing archaea and bacteria, anammox (Chapter 2) and denitrifying bacteria, as well as sulfate reducing and sulfur oxidizing bacteria including sulfide-dependent denitrifiers (Chapter 4) in coastal sediments experiencing decreasing bottom water oxygen con-centrations due to summer stratification.

In general, the microbial community in Lake Grevelingen surface sediments consisted of phototrophic, heterotrophic and autotrophic microorganisms involved in carbon-, nitrogen and sulfur cycling. Thaumarchaeota have been detected in Oyster Ground sediments and Lake Grev-elingen sediments, where they seem to play an important role in the reoxidation of reduced inorganic compounds in order to sustain biogeochemical cycling of carbon and nitrogen com-pounds especially under oxygenated bottom water conditions and/or in bioturbated sediments (Fig. 1). The diversity of anammox bacteria in Oyster Grounds sediment is comparable to the anammox bacterial diversity in Lake Grevelingen sediments by the 16S rRNA gene diversity, indicating that the diversity of anammox bacteria is less affected by seasonal hypoxia and rising sulfide concentrations compared to the archaeal population as shown in chapter 2 and 4.

7.3. Abundance and distribution of chemolithoautotrophic microorganisms in coastal sedimentsMicrobial processes such as nitrification, denitrification and anammox are important for ni-trogen and carbon cycling, and are tightly coupled in marine sediments (Zhang et al., 2013). The abundance and seasonal variations of archaeal and bacterial ammonia oxidizers, anammox and denitrifying, and sulfide dependent denitrifiers, as well as environmental factors affecting these groups, are not well studied in coastal marine sediments. In this thesis, the seasonal and depth distribution of the abundance of these groups was estimated in marine sediments of the Oyster Grounds (North Sea), in sediments of the Iceland shelf, as well as in Lake Grevelingen sediments to get insights of the importance of these groups and their impact in the carbon and nitrogen cycles in selected marine sediments.

In Oyster Grounds sediments (Chapter 2) this was achieved by quantifying specific intact po-lar lipids attributed to ammonia oxidizing Thaumarchaeota (AOA) (i.e. hexose phosphohexose (HPH) crenarchaeol) and anammox bacteria (i.e. phosphatidylcholine (PC)-monoether ladder-ane), as well as the abundance of the 16S rRNA genes, the ammonia monooxygenase subunit A (amoA) gene of AOA and AOB, and the hydrazine synthase (hzsA) of anammox bacteria, in February, May, and in August during decreasing bottom water oxygen concentrations. One of

Page 181: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

173

Synthesis

the highlights of the study in the Oyster Grounds sediments is the detection of aerobic am-monia oxidizing microorganisms at depths where the oxygen concentration was expected to be undetectable, which is counterintuitive with their aerobic metabolism. This was attributed to the intense bioturbation reported in these sediments which could provide enough oxygen to support the presence and activity of these microbial groups. This is further supported by other studies like the one of Beman et al. (2012), where AOA and AOB amoA genes were detected up to 10 cm sediment depth in bioturbated sediments of Catalina Island (CA, USA). My results, and those of Beman et al. (2012), suggest that aerobic ammonia oxidizers are indeed living and actively involved in the marine sedimentary nitrogen cycle in deeper layers of coastal marine sediments than previously thought, which is relevant as it extends the area of active dark carbon fixation and nitrification in marine sediments.

Another conclusion of the study performed in the Oyster Grounds sediments is that seasonal variation in the abundance of Thaumarchaeota (AOA) within the sediment given by quantifi-cation of 16S rRNA and amoA gene were not reflected in the specific Thaumarchaeota lipid biomarker HPH-crenarchaeol quantification, which was stable with depth in the first 12 cm. HPH-crenarchaeol has been shown to be more labile compared to crenarchaeol with a glycosidic head group, and its concentration corresponds well to the DNA-based abundance of Thaumar-chaeota (Pitcher et al., 2011a; Schouten et al., 2012; Lengger et al., 2012a). These results might suggest a possible preservation of this biomarker within the sediment vertical profile which would invalidate the use of it as indicator of abundance of living Thaumarchaeota in deep sed-iments.

The presence and potential activity of Thaumarchaeota were also assessed in Iceland shelf sediment incubations with 13C-labeled substrates (chapter 3). This study also included a direct analysis of IPL-crenarchaeol as specific lipid of Thaumarchaeota. The IPL-crenarchaeol with (HPH)- head group was between 80−90% in surface sediments of all stations and it decreased up to 10−25% in the deepest sediment layer. However, the presence of HPH-crenarchaeol in anoxic parts of the sediment, as seen in the Oyster Grounds sediments (Chapter 2), can again be due to preservation of this compound and cannot be then taken as a biomarker of a living Thaumarchaeota population.

In order to rule out a possible preservation effect of lipid biomarkers in deep sediments, the analysis or archaeal lipids was later restricted to surface sediments in Lake Grevelingen in or-der to determine the abundance and distribution of chemolithoautotrophic ammonia oxidizing Thaumarchaeota as well as other archaeal groups present. The methodological improvement of this study with respect to the one included in Chapter 2 and chapter 3 is that a new method was applied for the detection of the polar head groups of different IPL-GDGTs and was not only restricted to the specific GDGT crenarchaeol of Thaumarchaeota. With this, it was possible to extend the analysis of IPL-GDGT to the entire archaeal population present in the sediments. The detection of HPH-GDGTs in March in Lake Grevelingen sediment, coinciding with a higher presence of Thaumarchaeota detected by 16S rRNA gene amplicon sequence, indicated that the Thaumarchaeota population was actively living in those sediments. On the other hand, the switch to an archaea DPANN-dominated population in August with a similar IPL-GDGT profile, and the fact that DPANN archaea are not genetically capable of synthesizing their own lipids, suggest that this archaea group might be recycling the IPL-GDGTs already present in the sediment. This factor complicates the interpretations of the lipid sources in sedimentary settings

Page 182: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

174

Figu

re 1

: The

abu

ndan

ce a

nd a

ctiv

ity o

f A

OA

(gre

en) a

nd A

OB

(dar

k bl

ue),

as w

ell a

s of

ana

mm

ox b

acte

ria (y

ello

w) i

n bi

otur

bate

d se

dim

ents

of

the

Oys

ter G

roun

ds (N

orth

Sea

) (0–

4 cm

sedi

men

t dep

th).

Big

blue

arr

ow in

dica

tes o

xyge

n su

pply

by

biot

urba

tion.

Red

arr

ow in

dica

tes p

oten

tial a

bund

ance

an

d ac

tivity

of

othe

r m

icro

bial

gro

ups.

Plea

se n

ote

that

the

dash

ed li

nes

do n

ot r

eflec

t the

act

ual d

epth

pro

files

of

biom

arke

rs (fi

nd in

cha

pter

2),

but

repr

esen

t the

pro

file

dist

ribut

ion

(take

n to

geth

er th

e ab

unda

nce

and

pote

ntia

l act

ivity

); (th

in d

ashe

d lin

e) lo

wer

abu

ndan

ce a

nd a

ctiv

ity, (

thic

ker

dash

ed

lines

) hig

her a

bund

ance

and

act

ivity

.

Page 183: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

175

Synthesis

Figu

re 2

: Rep

rese

ntat

ion

of t

he p

rofil

e di

strib

utio

n (ta

ken

toge

ther

the

abu

ndan

ce a

nd p

oten

tial a

ctiv

ity)

(das

hed

lines

) an

d ch

emol

ithoa

utot

roph

ic

activ

ity (13

C P

LFA

-SIP

) (tri

angl

e) in

Lak

e G

reve

linge

n se

dim

ents

dur

ing

oxyg

enat

ed b

otto

m w

ater

con

ditio

ns (l

eft)

and

durin

g hy

poxi

a (ri

ght)

(0–5

cm

se

dim

ent d

epth

) (C

hapt

er 4

, 5 a

nd 6

). [c

olor

cod

e: (p

ink)

Tha

umar

chae

ota,

(dar

k bl

ue) h

eter

otro

phic

and

aut

otro

phic

den

itrify

ing

bact

eria

, (gr

een)

sulfa

te

redu

cing

and

sulfu

r oxi

dizi

ng m

icro

orga

nism

s inc

ludi

ng su

lfide

-dep

ende

nt d

enitr

ifier

s, (p

urpl

e) c

able

bac

teria

, (re

d) B

eggi

atoa

ceae

, (lig

ht b

lue)

sulfu

r ox-

idiz

ing

Gam

map

rote

obac

teria

, (da

rk g

reen

) sul

fur o

xidi

zing

Eps

ilonp

rote

obac

teria

, (or

ange

) DPA

NN

Woe

sear

chae

ota,

(ligh

t blu

e) s

ulfa

te re

duci

ng a

nd

sulfu

r disp

ropo

rtio

ning

Del

tapr

oteo

bact

eria

]. (b

lack

char

acte

rs) a

bund

ant a

nd p

oten

tially

activ

e, (g

rey c

hara

cter

s) ab

unda

nt an

d lo

w p

oten

tial a

ctiv

ity. O

nly

redu

ced

spec

ies a

nd e

lect

ron

acce

ptor

s are

show

n.

Page 184: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

176

and highlight the need of combining different methodological approaches (i.e. lipid biomarkers, gene-based detection, activity measurements) to determine the chemolithoautotrophic potential and biological sources. Also, this study highlights the dramatic change in the archaea population in Lake Grevelingen surface sediments upon summer hypoxia and increase of sulfide concentra-tion in the pore water. This can be explained by a toxic effect of sulfide on the Thaumarchaeota population, together with decrease in oxygen available for their aerobic metabolism. In fact, a similar toxic effect of sulfide has been observed in Thaumarchaeota enrichment cultures from Baltic Sea water (Berg et al., 2015).

In this thesis, I have also determined the abundance and distribution of anammox bacteria in different marine sediment systems both by using specific lipid biomarkers and quantification of specific 16S rRNA and metabolic genes. For example, anammox bacteria 16S rRNA and the hydrazine synthase (hzsA) gene abundance were detected throughout the Oyster Grounds sediment with more pronounced depth and seasonal differences than those observed for AOA and AOB. Both, anammox bacteria 16S rRNA and hzsA gene abundances were higher in August compared to February and May, which was not reflected in the abundance of the anammox spe-cific lipid biomarker PC-monoether ladderane. This can be explained by partial preservation of this biomarker lipid in the sediment due to a relatively low turnover rate (Brandsma et al., 2011; Bale et al., 2014). I also observed a ten-fold difference between abundances of the anammox 16S rRNA gene and the hzsA gene. This difference between the 16S rRNA gene and functional genes of anammox bacteria has been previously observed and is possibly explained by primer biases of the 16S rRNA gene primers, suggesting the amplification of 16S rRNA gene fragments of bacteria other than anammox bacteria (Harhangi et al., 2012; Bale et al., 2014).

The presence of anammox bacteria biomarkers in the upper layers of Oyster Grounds sed-iments (0−4 cm) is still remarkable as the metabolism of anammox bacteria is expected to be incompatible with the presence of oxygen. Some studies have reported that anammox bacteria can be present at high oxygen levels, by being dormant or using anaerobic niches (Kuypers et al., 2005; Woebken et al., 2007). It is also possible that the metabolic activity of AOA/AOB and heterotrophic bacteria support the presence of anammox bacteria by oxygen consumption al-lowing the presence of anammox bacteria even in the oxygenated sediments. My results suggest that anammox bacteria can tolerate the presence of oxygen (Kalvelage et al., 2011; Yan et al., 2012) and co-exist with aerobic ammonia oxidizers and potentially have an important role in the nitrogen cycle in marine sediments.

Similarly, anammox bacteria detected in Lake Grevelingen sediments showed a clear seasonal contrast in their nitrite reductase (nirS) gene abundance with higher copy numbers in August during summer hypoxia compared to March, similar to the 16S rRNA and hzsA gene abundance of anammox bacteria detected in Oyster Ground sediments (Chapter 2). The seasonal trend was as well not reflected in the abundance of the anammox bacteria lipid biomarker PC-monoether ladderane lipid, as it has been reported for Oyster Grounds sediments (Brandsma et al., 2011; Bale et al., 2014). In contrast, the concentration of ladderane lipid fatty acids correlated with the abundance of anammox bacteria determined by the 16S rRNA gene abundance, suggesting that the abundance of ladderane lipid fatty acids could be interpreted as a proxy for the abundance of anammox bacteria, together with specific gene quantification. A difference of three to four orders of magnitude between abundances of the anammox 16S rRNA gene and the “Candidatus Scalindua” specific nirS gene was detected. This discordance between the 16S rRNA gene and

Page 185: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

177

Synthesis

functional genes of anammox bacteria has been previously observed and has been discussed above in detail, thus the possibility of amplifying other bacteria has been ruled out by cloning and sequencing of the product generated during the quantitative PCR reaction for Lake Grev-elingen sediments.

Overall, the nirS gene abundance of denitrifiers was three to four orders higher compared to the anammox nirS gene abundance, suggesting that nirS-type denitrifiers have a more prominent role in the nitrogen removal in Lake Grevelingen sediments compared to anammox bacteria. The abundance of denitrifying bacteria did not vary substantially despite of differences in the oxygen and sulfide availability, whereas anammox bacteria were more abundant in the summer hypoxia, but in those sediments with lower sulfide concentrations, indicating that the abundance of anammox bacteria is impaired by high sulfide concentrations as previously suggested by oth-ers (Dalsgaard et al., 2003; Jensen et al., 2008). Anammox abundance was also highest in Oyster Grounds sediments (0.5 cm) in August under low bottom water oxygen conditions and decreas-ing oxygen penetration. These results are most likely explained by of the fact that anammox bacteria are favored by hypoxic conditions (i.e. Neubacher et al., 2013). Besides, the optimum temperature of anammox bacteria of 15°C has been reported in temperate shelf sediments (Dalsgaard et al., 2005) which is close to the summer temperature in Oyster Grounds and Lake Grevelingen sediments (~17°C). Additionally, bioturbation has been shown to extend the area of aerobic processes such as nitrate reduction and thus the availability of reduced compounds such as nitrite, which would fuel the anammox process (Meyer et al., 2005). In Lake Grevelingen sediments, anammox abundance was highest in bioturbated compared to non bioturbated sed-iments (S3), indicating that bioturbation might additionally favor the abundance of anammox bacteria. Bioturbation has also been reported in Oyster Grounds (Duineveld et al., 1991), which together with the reported high activity of AOA, AOB and/or other aerobic heterotrophs re-moving the available oxygen, could contribute to the proliferation of anammox bacteria.

In order to determine the abundance of other microorganisms involved in nitrogen removal in Lake Grevelingen sediments (Chapter 4), I also quantified the nitrite reductase (nirS) gene of chemoorganoheterotrophic denitrifying bacteria and archaea, performing the stepwise conver-sion of nitrate/nitrite to dinitrogen gas (Zumft, 1997), as well as the aprA gene of sulfate reduc-ing and sulfur oxidizing bacteria including autotrophic sulfide-dependent denitrifying members of the Alpha-, Beta-, Gamma-, and Epsilonproteobacteria (Shao et al., 2010).The nirS gene abundance has previously been show to correlate positively with potential denitrification rates (Mosier and Francis, 2010). Here, the nirS gene abundance of denitrifying bacteria was relatively constant with sediment depth, suggesting a stable community of heterotrophic and autotro-phic denitrifying bacteria actively involved in the denitrification process in Lake Grevelingen sediments. However, assumptions have to be made with caution, because foraminiferal denitri-fication might be as important as heterotrophic denitrification in marine sediments, as well as metal-dependent denitrification (Devol, 2015). Besides, and some denitrifying microorganisms use the copper-containing nitrite reductase encoded by the nirK gene which could have led to the underestimation of the diversity and abundance of total denitrifiers by exclusively targeting the nirS gene. However, nirS denitrifiers are more likely to have a complete denitrification pathway (including nor and nos genes), suggesting that nirS-containing denitrifiers may be more likely to reduce nitrite to N2 (Graf et al. 2014). In addition, the nirS gene has been shown to be more abundant and more diverse than nirK in estuarine sediments, although this pattern sometimes

Page 186: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

178

shifts in different estuarine regions (Nogales et al. 2002; Abell et al. 2010; Mosier and Francis 2010; Beman, 2014; Smith et al. 2015b; Lee and Francis, 2017). In general, the nirS gene abun-dance was slightly higher in March in all three stations in Lake Grevelingen sediments (Chapter 4), but the difference was more pronounced in station 1, which can be most likely explained by lower sulfide concentrations in March compared to August. The lower sulfide concentrations in March could favor the proliferation of heterotrophic denitrifiers, as sulfide has been shown to impair denitrification by inhibition of NO and N2O reductases (Sørensen et al., 1980).

The abundance of chemolithoautotrophic microorganisms utilizing the CBB and rTCA path-ways for carbon fixation was also estimated in Lake Grevelingen sediments (Chapter 5) by quan-tifying the cbbL and aclB gene, encoding for the key-enzymes ribulose 1,5 bisphosphate carbox-ylase/oxygenase (CBB) and ATP citrate lyase (rTCA). Significant differences were determined in the abundance of cbbL and aclB genes between the seasons but not between the stations, suggesting a stable community of community of autotrophic microorganisms using the CBB and rTCA pathway. The abundance of the cbbL gene was at least 2-fold higher than aclB gene abundance independent of the season or station, indicating a more prominent role of microor-ganisms using the CBB cycle compared to those using the rTCA pathway. The results indicated a more important role of autotrophic aerobic or facultative aerobic Gammaproteobacteria of the orders Chromatiales and Thiotrichales compared to Epsilonproteobacteria of the order Cam-pylobacterales and Euryarchaeota in Lake Grevelingen sediments. However, these assumptions have to be made with caution, as the primers used to amplify aclB genes are mainly based on sequences of Epsilonproteobacteria (Campbell et al., 2006), suggesting that potentially there is an underestimation of the abundance of microorganisms using the rTCA pathway in these sed-iments, as well as the abundance of microorganisms using other carbon fixation pathways than the CBB or rTCA pathway (Hügler and Sievert, 2011; Berg et al., 2010, 2011).

Taking all the considerations mentioned above, the determination of the abundance of mi-croorganisms based on DNA sequences and lipid biomarker molecules in marine coastal and estuarine sediments has to be interpreted with caution. DNA-based methods employed in this thesis depend on PCR amplification with primers, designed based on the sequences available in the databases, potentially causing an underestimation of the diversity as well as of the abundance of target organisms. As the abundance of functional and 16S rRNA genes was estimated by qPCR and 16S rRNA gene amplicon sequencing, I cannot completely rule out primer induced PCR biases (Schloss et al., 2011; Sipos et al., 2010). Additionally, DNA and lipid biomarkers can be partly preserved in anoxic sediments, which possibly leads to an overestimation of the abun-dance of the studied microorganisms (Torti et al., 2015; Lengger et al., 2012; Rush et al., 2012).

7.4. Activity of chemolithoautotrophic microorganisms in coastal sediments In my thesis, I have estimated activity by determining incorporation of 13C-labeled substrates in membrane lipids as well as the gene expression of 16S rRNA and functional genes as proxy of transcriptional activity. The abundance and gene expression of functional genes has been frequently interpreted as an indication of the process rate in which the protein-coding gene is involved (Nogales et al., 2002; Smith et al., 2007; Abell et al., 2010, 2011; Bowen et al., 2014). However, this is not always the case and it is in general dependent on the regulation of the specif-ic gene under study. As an example, the transcript levels of the coding gene for the dissimilatory sulfite reductase (dsrA), catalyzing the final step in sulfate reduction, has been seen to positively

Page 187: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

179

Synthesis

correlate with sulfate reduction rate (SRR) when growing under limited electron acceptor (sul-fate) while no correlation between SRR and dsrA gene expression was observed when grown under electron donor limitation (Villanueva et al., 2008). Similarly, a recent study by Carini et al. (2017) observed that ammonia limitation in the thaumarchaeon “Candidatus Nitrosopelagicus brevis” induced a decrease in the expression of the ammonia oxidation core proteins, which could be interpreted as a confirmation that lower ammonia oxidation activity is correlated to lower amoA gene expression levels. However, this study also concluded that the genes involved in ammonia oxidation constitute a large part of the total transcript pool irrespective of growth and environmental conditions, which makes it difficult to connect changes in gene expression to differences in the metabolic process.

Examples like this illustrate the need of studying, in pure cultures and under controlled con-ditions, the regulation of the expression of metabolic genes and its potential correlation to the process rates before making definite conclusions based on environmental samples. In this thesis, several metabolic genes involved in key re-oxidation pathways were analyzed for their gene ex-pression patterns in environmental samples. For some of these genes, previous information on the regulation of their gene expression was available (e.g. AOA amoA gene; Abell et al., 2011), which allowed us to interpret the results more confidently. For some other genes, e.g. anammox bacterial hzsA gene (Harhangi et al., 2011), the determination of their gene expression and comparison with their gene abundance and process rate in environmental samples in this thesis (Chapters 2, 4), has allowed to get a better understanding of the environmental conditions that can affect the activity of the metabolic process.

To estimate the activity of aerobic and anaerobic ammonia oxidizers and anammox bacteria in coastal sediments, the gene expression of their 16S rRNA gene, the amoA gene of AOA and AOB, and the expression of the anammox bacterial hzsA gene, were analyzed here by quantify-ing the transcript abundance by reverse transcription and subsequent qPCR, in sediments of the Oyster Grounds (North Sea) (Chapter 2) and in Lake Grevelingen (chapter 4). The detection of the transcriptional activity of the AOA 16S rRNA gene up to 12 cm depth in Oyster Grounds sediment supports that AOA are indeed active in subsurface sediments. The potential transcrip-tional activity, by the 16S rRNA gene RNA:DNA ratio, was higher during winter, whereas the abundance was slightly higher during the summer months. The AOA amoA gene RNA:DNA ratio has been detected in August throughout the core but only in the upper 3 cm in February, because the quantity of transcripts in other seasons and depths was under the detection limit of the qPCR assay. However, the relationship of amoA gene expression and in situ ammonia oxi-dation is still unclear (Lam et al., 2007; Nicol et al., 2008; Abell et al., 2011; Vissers et al., 2013). Recent studies have observed a good correlation between nitrification rates and the abundance of amoA genes and transcripts in coastal waters (Smith et al., 2014; Urakawa et al., 2014), indi-cating that the detection of AOA amoA transcripts suggest AOA were active in Oyster Grounds sediments in the summer throughout the sediment in comparison with the winter in which AOA would be only active in the upper part of the sediment (upper 2 cm).

The abundance of AOB amoA transcripts was higher than for AOA, and was detected up to 12 cm sediment depth in all seasons, which would suggest a higher AOB activity compared to AOA. However, a study by Urakawa et al. (2014) in estuarine waters did not observe a correlation between nitrification rates and AOB amoA transcripts. Additionally, previous studies have de-tected the presence of AOB amoA transcripts in non-active starved cells (Bollmann et al., 2005;

Page 188: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

180

Geets et al., 2006; Sayaveda-Soto and Arp, 2011) indicating that the AOB amoA transcriptional activity is not a good indicator of AOB nitrification activity due to a longer half-life of AOB amoA transcripts than expected for AOA amoA transcripts.

The metabolic capacities of sedimentary Thaumarchaeota are still under debate. To shed fur-ther light on this important chemolithoautotrophic metabolism in marine sediments, we also performed labeling experiments using 13C-labeled substrates, such as bicarbonate, pyruvate, ami-no acids and glucose in sediment cores of the Iceland Shelf (Chapter 3) and determined the de-gree of 13C-incorporation into the most abundant GDGT lipids, i.e. crenarchaeol and GDGT-0. These compounds have been suggested to be taken up by Thaumarchaeota in the environment (Ouverney and Fuhrman, 2000; Wuchter et al., 2003; Herndl et al., 2005; Takano et al., 2010; Pitcher et al., 2011c), or in enrichment cultures (Park et al., 2010; Jung et al., 2011; Tourna et al., 2011; Kim et al., 2012), and are possibly the most suitable substrates resulting in uptake by sedimentary Archaea. However, the incubation with labeled 13C substrates did not result in any substantial incorporation of the 13C into the biphytanyl chains of these lipids. No incorpora-tion of 13C in IPL-GDGT was detected in clay rich and organic rich sediments, whereas some enrichment in δ13C (1−4‰) of IPL-GDGTs recovered from sandy sediment was detected after incubation with bicarbonate and pyruvate. This apparent lack of uptake could be due to several reasons. It is possible that the Thaumarchaeota were active, but the substrates were not taken up. Secondly, it is possible that most of the IPLs measured are fossil, and thus only a small amount of these IPLs are part of living and active sedimentary Thaumarchaeota, as also discussed above. Another possibility is that the metabolism and growth rates are so slow that no incorporation could be detected over the timescales investigated. Thus, the incubation period of 96−144 h, may have been too short. These results suggested that sedimentary Thaumarchaeota grow slowly in contrast to other sedimentary organisms, and/or a slow turn-over of IPL-GDGTs compared to bacterial/eukaryotic PLFAs. Interestingly, enrichment culture experiments with a thaumarchaeal strain closely related (99%) to Nitrosopumilus maritimus (SCM1), discovered that α-keto acids, such as pyruvate and oxaloacetate, were not directly incorporated into AOA cells, but functioned as chemical scavengers that detoxified intracellularly produced H2O2 during ammonia oxidation (Kim et al., 2016), indicating that previous findings of pyruvate incorporation could have been misinterpreted as evidence for mixotrophic growth of AOA.

In order to determine the potential activity of anammox bacteria in Oyster Grounds and Lake Grevelingen sediments, the potential transcriptional activity of the 16S rRNA gene, as well as of the functional genes hzsA and nirS encoding for the key-enzymes of the anammox reaction, hydrazine synthase and nitrite reductase (Chapter 2 and 4, respectively). A previous study by Bale et al. (2014) observed a good correlation between anammox rate activity and anammox bacterial 16S rRNA transcript abundance in North Sea sediments, suggesting that the tran-scriptional activity of anammox bacteria 16S rRNA gene is a good proxy for anammox bacteria activity. In Oyster Grounds sediments, the transcriptional activity of anammox bacteria 16S rRNA and hzsA gene was detected throughout the sediment in all seasons, and the 16S rRNA RNA:DNA ratio was higher during summer as observed for the anammox bacteria abundance. In Lake Grevelingen sediments the anammox 16S rRNA gene transcript abundance was detect-ed throughout the sediment. However, the anammox bacteria 16S rRNA RNA:DNA ratio was low (0.01−1) compared to the values reported in sediments of the Oyster Grounds (1.6−34.5), further supporting a low anammox activity in Lake Grevelingen sediments.

Page 189: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

181

Synthesis

Nevertheless, quantification results of anammox bacteria 16S rRNA gene have to be inter-preted with caution because there is evidence that the ribosome content does not decrease during the period of starvation (Schmid et al., 2005; Li and Gu, 2011). Besides, the increase of anammox bacteria 16S rRNA gene transcriptional activity during summer in Oyster Grounds sediments was not observed for the hzsA gene, which remained stable with sediment depth and also for the different seasons. The low and constitutive transcriptional activity of the hzsA gene indicates that the hzsA gene transcript abundance does not reflect variations of anammox bac-teria activity in environmental sediment samples, thus the hzsA transcript abundance does not seem to be an adequate biomarker for the estimation of the activity of anammox bacteria. Fur-ther experiments, preferably with pure cultures and under controlled physiological conditions, should be performed to clarify the regulation of the hzsA gene expression.

Environmental conditions favoring the abundance and activity of anammox bacteria have been discussed above in detail. Bioturbation activity could explain the differences in the activity of anammox bacteria in sediments of the Oyster Grounds compared to the Lake Grevelingen sediments, where only in station S3 bioturbators have been recognized in March, coinciding with the highest anammox abundance compared to the other stations. Additionally, the activity of anammox bacteria might be inhibited by high carbon and sulfide concentrations, as discussed above, probably explaining the higher activity of anammox bacteria in Oyster Grounds (low TOC), compared to Lake Grevelingen sediments (high TOC).

To further estimate the potential activity of microbial groups involved in denitrification oth-er than anammox bacteria, i.e. denitrifying bacteria, sulfate reducing/sulfur oxidizing bacteria including sulfide-dependent denitrifying microorganisms, in seasonal hypoxic and sulfidic sedi-ments of Lake Grevelingen (Chapter 4), I also analyzed the potential transcriptional activity of these groups. This was achieved by quantifying the nirS transcript abundance of denitrifying bacteria, as well as the aprA transcript abundance of sulfate reducing and sulfur oxidizing mi-croorganisms. Unfortunately, there is a lack of studies correlating the nirS gene expression of denitrifying bacteria with denitrification rates, and some studies have shown a lack of direct relationship between nirS gene expression and modeled rates of denitrification in marine surface sediments (Bowen et al., 2014), whereas other studies have observed a correlation between a decrease of nirS transcript abundance and the reduction of denitrification rates coinciding with declining concentrations of nitrate in estuarine sediments (Dong et al., 2000, 2002; Smith et al., 2007). These studies would justify the use of nirS gene transcript abundance as a proxy for denitrification performed by nirS-type denitrifiers. In my study, the nirS transcript abundance of nirS-type denitrifiers was two or three orders lower compared to values previously reported for estuarine sediments, suggesting a low denitrification rate in surface sediments of all stations in March and in deeper sediment layers in August. Gene expression of the nirS gene was detected in surface sediments in March (all stations) and in deeper layers in August (stations S1 and S3), which is most likely explained by the availability of nitrate/nitrite in the bottom water. The potential transcriptional activity of sulfate reducing and sulfur oxidizing including chemolitho-autotrophic sulfide-dependent denitrifiers, determined by the aprA gene RNA:DNA ratio was relatively low (≤ 0.00088), which also indicated a low denitrification potential of these groups.

Although denitrification was more relevant than the anammox process as suggested by the gene abundance and potential transcriptional activity, and measured denitrification rates (be-tween 36 and 96 μM m−2 d−1; Seitaj, D. personal communication) in Lake Grevelingen sediments

Page 190: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

182

which were low compared to other marine sediments. Denitrification rates vary between 30 and 270 μM m−2 d−1 in marine arctic sediments (Rysgaard et al., 2004), whereas in estuarine sediments of the lower St. Lawrence estuary, denitrification rates of 270 ± 30 μM m−2 d−1 were measured (Crowe et al., 2012). The lower denitrification and anammox potential in Lake Grev-elingen sediments could be explained by the effect of high concentrations of sulfide, potentially inhibiting both heterotrophic denitrification and anammox. Another explanation of the low denitrification and anammox potential of Lake Grevelingen sediments can be found in the in-teractions of anammox and heterotrophic denitrifiers with sulfur-oxidizers also involved in the nitrogen cycle. Large filamentous sulfur oxidizing bacteria have been detected in Lake Grevelin-gen sediments, i.e. cable bacteria and Beggiatoaceae including the genera Beggiatoa, Thiomargarita and Thioploca (Salman et al., 2011, 2013). Cable bacteria perform electrogenic sulfur oxidation (e-SOx), in which the oxidation of the electron donor and the reduction of the electron accep-tor are separated over cm-scale distances and the redox coupling is ensured by long distance electron transport (Nielsen et al., 2010; Pfeffer et al., 2012). Cable bacteria can play a role in bioavailable nitrogen removal, as they have been shown to use oxygen but might as well use nitrite and nitrate (Meysman et al., 2015; Marzocchi et al., 2014; Risgaard-Petersen et al., 2014). Additionally, Beggiatoaceae couple the oxidation of sulfide to nitrate reduction resulting in N2 and/or NH4

+ (Mußmann et al., 2003). Previous studies have observed that in anoxic sediments both denitrification and anammox can be supported by intracellular nitrate transport, performed by sulfide-oxidizing Thioploca and Beggiatoa (Mußmann et al., 2007; Jorgensen, 2010; Prokopenko et al., 2013) down to deeper sediment layers and possibly supplying anammox and denitrifying bacteria with nitrite and/or ammonia produced by DNRA (Teske and Nelson, 2006; Proko-penko et al., 2011). Cable bacteria have been detected in S1 and S3 in March, whereas station S2 was dominated by Beggiatoaceae down to ~3 cm, but both groups were hardly detected in subsurface sediments (0.5 cm downwards) during summer hypoxia. The presence of sulfide-de-pendent denitrifiers, as shown by the aprA gene abundance and diversity, as well as the presence of cable bacteria and Beggiatoa-like bacteria could potentially support the denitrification and anammox processes in March. However, the nitrate and nitrite concentrations in the bottom water in March were reported to be relatively low (28 and 0.7 μM on average, respectively). This limitation could induce a strong competition for nitrate, nitrite and sulfide with Beggiatoaceae, cable bacteria and other nitrate-reducing bacteria, which may explain the limited denitrification potential observed in March. During summer hypoxia (August), sulfide inhibition and low nitrate and nitrite concentrations seem to limit the activity of heterotrophic and autotrophic denitrifiers and anammox bacteria. Further studies, involving denitrification rate measurements will be re-quired to further assess the effects of hypoxia and high sulfide concentrations in the sediments of Lake Grevelingen.

The chemolithoautotrophic activity of the bacterial community has also been determined by PLFA analysis combined with 13C-stable isotope probing in Grevelingen sediments (Chapter 5). In March, under oxygenated bottom water conditions, the chemolithoautotrophic rates were substantially higher than those in August, when oxygen in the bottom water was depleted to low levels. Moreover, in August, the chemolithoautotrophic rates showed a clear decrease in with water depth (S3 to S1), which is in line with the bottom water oxygen concentration over water depth. At the same time, the 16S rRNA gene amplicon sequencing revealed that the re-sponse of the chemolithoautotrophic bacterial community in surface sediments is associated

Page 191: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

183

Synthesis

with the seasonal changes in the bottom water oxygen concentrations. As described above, in March the microbial community in surface sediments of all stations was characterized by high relative abundances of chemolithoautotrophic sulfur oxidizing Gammaproteobacteria and Ep-silonproteobacteria using oxygen or nitrate as electron acceptor. In August, the lack of oxygen and nitrate in the bottom water was accompanied by a decrease in the relative abundances of Gamma- proteobacteria and Epsilonproteobacteria, and an increase of the relative abundance of heterotrophic sulfate-reducing Deltaproteobacteria using sulfate as electron acceptor. These results support the chemolithoautotrophic rates measured in March and August, indicating that the lack of electron acceptors such as oxygen and nitrate induced by hypoxia limiting the activity of chemolithoautotrophic bacteria in August. Recently, a study by Jessen et al. (2017) addressed the effects of hypoxia on the microbial community composition in Black Sea sediments, and showed that variations in oxygen supply to the seafloor influenced the composition of the bac-terial community in surface sediments. Similar to my results, they reported an increase of Del-taproteobacteria under variable to anoxic conditions, whereas Gammaproteobacteria reached highest abundances under variable hypoxic conditions while decreasing toward stable oxic and anoxic conditions (Jessen et al., 2017). Together, these results show the important impact of bottom water hypoxia on the microbial community composition and especially on chemolitho-autotrophs in sediments.

7.5. OutlookIn the last five years, new and improved methods have been developed and their cost reduced, which would now make them compatible with studies with a massive amount of samples. In the last two chapters of this thesis, I have already applied both 16S rRNA gene amplicon sequencing and lipid analysis with high resolution accurate mass/mass spectrometry that showed a quanti-tative and qualitative increase in the amount of data and the estimation of the diversity of not only chemolithoautotrophic microorganisms.

The studies included in this thesis regarding the gene expression and abundance of key met-abolic genes of chemolithoautotrophs under different environmental and physiological condi-tions have shed further light on the role of this group in the carbon and nitrogen cycle in coastal sediments. In this thesis, the diversity of chemolithoautotrophic microorganisms was mostly determined with gene-targeted approaches, either by amplifying and sequencing specific 16S rRNA gene or functional genes involved in the carbon fixation of re-oxidation pathways. In this synthesis, I have listed the disadvantages of this approach mainly due to the biases introduced by the primer design, lack of sequences in databases, and inherent PCR amplification biases. Currently, these limitations could have been overcome by metagenomic and metatranscriptomic approaches, which are now much more affordable, the data analysis is more feasible, and most importantly they are not PCR-biased. With this approaches, I could have determined both di-versity and the gene expression of chemolithoautotrophic key genes. However, the connection between gene expression and the activity of the actual process is still problematic. This infor-mation can only be validated by laboratory experiments in which the regulation of the gene of interest is thoroughly studied under different physiological and environmental conditions. Another recommendation is the combination of activity proxy measurements included in this thesis with direct rate measurements when possible to directly assess if the gene expression can be correlated to the activity measurement. Apart from using metatranscriptomics, the activity

Page 192: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

Chapter 7

184

of specific microorganisms could now also detected by using more sensitive approaches than the lipid-SIP approaches used in this thesis. For example, nano-scale secondary ion mass spec-trometry (NanoSIMS) in combination with catalyzed reporter deposition (CARD)- fluorescence in situ hybridization (FISH) has proven to be very effective for the determination of metabolic capacity linked to phylogenetic identity at the single cell level (Behrens et al., 2008; Musat et al., 2008; Morono et al., 2011). However, a problem that remains is that this method still relies on incubation with labeled substrates with the inherent limitations of labeling-based incubations including labeling incubation time and cross-feeding of 13C metabolites produced by the primary consumers of the 13C substrate.

Finally, in this thesis both specific lipid biomarkers for targeted groups (i.e. crenarchaeol for Thaumarchaeota, ladderane lipids for anammox) and bacterial and archaeal lipid characterization were performed in sediment samples with the aim of determining the abundance and activity (i.e. Lipid-SIP) of chemolithoautotrophs. However, I have seen that some of this biomarkers (i.e. HPH-crenarchaeol, PC-ladderane) are potentially preserved in deeper sediments invalidating their use as proxies of living biomass. In this regard, a recommendation would be to continue in the search of lipid biomarkers that can be more representative of living biomass and less prone to preservation. Laboratory cultures together with characterization of the microbial lipidome with methods similar to the ones applied in Chapter 6 (i.e. high resolution accurate mass spec-trometry) to detect intact polar lipids would be key in improving this issue. Also in chapter 6, I have seen the potential recycling of IPLs by archaeal groups like the DPANN. Further studies need to carefully assess the effect of such lipid recycling in sediment and the interpretations derived from those results.

Page 193: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 194: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 195: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

187

Summary

The role of chemolithoautotrophic microorganisms has been considered to be of minor impor-tance in coastal marine sediments although it has not been investigated in depth. Additionally, the impact of seasonal hypoxic/anoxic conditions on microbial chemolithoautotrophy in coastal marine sediments has not been examined. Therefore, in this thesis the diversity, abundance and activity of specific groups of chemolithoautotrophic microorganisms involved in the nitrogen and sulfur cycling, as well as the total microbial community, has been determined in coastal sed-iments by using DNA/RNA- and lipid-based biomarker approaches. Results show the presence and potential importance of chemolithoautotrophs in the biogeochemical cycling of carbon, nitrogen and sulfur. Temporal changes of the diversity, abundance and activity of chemolith-oautotrophic microorganisms coinciding with seasonal hypoxia in coastal sediments have been determined.

The diversity, abundance and activity of chemolithoautotrophic microorganisms, such as am-monia oxidizing archaea (AOA) and bacteria (AOB), anammox and denitrifying bacteria, as well as sulfate reducing and sulfur oxidizing microorganisms, including sulfide dependent denitrifiers, have been determined. Additionally, the diversity and activity of chemolithoautotrophic micro-organisms involved in the CBB and rTCA cycle as well as the metabolism of Thaumarchaeota and the abundance of Archaea have been addressed. In the following studies, we focused on coastal marine sediments, exposed to summer hypoxia, i.e. (1) the Oyster Grounds in the central North Sea, exposed to decreasing bottom water oxygen concentrations, (2) different sediment types (sand, clay, and mud) of the Iceland Shelf, and (3) seasonally hypoxic and sulfidic sedi-ments of saline Lake Grevelingen (The Netherlands).

In Oyster Grounds sediments (chapter 2), seasonal variations of archaeal and bacterial am-monia oxidizers (AOA and AOB) and anammox bacteria, as well as the environmental factors affecting these groups, were addressed. We examined the seasonal and depth distribution of the abundance and potential activity of these microbial groups in coastal marine sediments of the southern North Sea. We quantified specific intact polar lipids (IPLs) as well as the abundance and gene expression of their 16S rRNA gene, the ammonia monooxygenase subunit A (amoA) gene of AOA and AOB, and the hydrazine synthase (hzsA) gene of anammox bacteria. AOA, AOB and anammox bacteria were detected and transcriptionally active down to 12 cm sediment depth. This study unraveled the coexistence and metabolic activity of AOA, AOB, and anammox bacteria in bioturbated marine sediments of the North Sea, leading to the conclusion that the metabolism of these microbial groups is spatially coupled based on the rapid consumption of oxygen that allows the coexistence of aerobic and anaerobic ammonia oxidizers. AOA outnum-bered AOB throughout the year which may be caused by the higher oxygen affinity of AOA compared to AOB. Anammox bacterial abundance and activity were higher during summer, indicating that their growth and activity are favored by higher temperatures and lower oxygen available in the sediments due to summer stratifying conditions in the water column. During the summer, anammox bacteria are probably not in competition with AOA and AOB for ammonia as the concentrations were relatively high in the sediment pore water most likely as a result of ammonification processes and the activity of denitrifiers and DNRA.

In Iceland Shelf sediments (chapter 3), we investigated the activity of Thaumarchaeota in sediments by supplying different 13C-labeled substrates which have previously been shown to be

Page 196: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

184

incorporated into archaeal cells in water incubations and/or enrichment cultures. We determined the incorporation of 13C-label from bicarbonate, pyruvate, glucose and amino acids into thau-marchaeal intact polar lipid-glycerol dibiphytanyl glycerol tetraethers (IPL-GDGTs) during 4–6 day incubations of marine sediment cores from three sites on the Iceland Shelf. Thaumarchaeal IPLs were present in sediment cores recovered from the Iceland Shelf and the presence of labile IPL-biomarkers for Thaumarchaeota, such as hexose phosphohexose (HPH)-crenarchaeol, sug-gested the presence of living Thaumarchaeota. However, incubations with 13C-labeled substrates bicarbonate, pyruvate, glucose and amino acids did not result in any substantial incorporation of the 13C into the biphytanyl chains of these lipids. Turnover times and generation times of thau-marchaeal lipids were estimated to be at least several years, suggesting slow growth of sedimen-tary Thaumarchaeota in contrast to other sedimentary organisms, and/or a slow degradation of IPL-GDGTs, in contrast to bacterial or eukaryotic PLFAs.

In Lake Grevelingen sediments (chapter 4), we evaluated the abundance, diversity and poten-tial activity of denitrifying, anammox, and sulfide-dependent denitrifying bacteria in the sedi-ments of the seasonally hypoxic saline Lake Grevelingen, known to harbor an active microbial community involved in sulfur oxidation pathways. Depth distributions of 16S rRNA gene, nirS gene of denitrifying and anammox bacteria, aprA gene of sulfur-oxidizing and sulfate-reducing bacteria, and ladderane lipids of anammox bacteria were studied in sediments impacted by sea-sonally hypoxic bottom waters. This study unraveled the coexistence and potential activity of heterotrophic and autotrophic denitrifiers, anammox bacteria as well as SOB/SRB in seasonally hypoxic and sulfidic sediments of Lake Grevelingen. The nirS-type heterotrophic denitrifiers were a stable community regardless of changes in oxygen and sulfide concentrations in different seasons and with a similar abundance to that detected in other sediments not exposed to high sulfide concentrations. Besides, nirS-type denitrifiers outnumbered anammox bacteria leading to the conclusion that anammox does not contribute significantly to the N2 removal process in Lake Grevelingen sediments. The anammox bacteria population seemed to be affected by the physicochemical conditions between seasons and their abundance and activity was higher in lower sulfide concentrations and low carbon input, also supporting a possible inhibition of anammox bacteria by sulfide. The sulfide-dependent denitrifiers in Lake Grevelingen sediments have proven to be abundant, but their expected sulfide-detoxification potential did not seem to alleviate the predicted inhibition of organoheterotrophic denitrifiers and anammox by sulfide. Both denitrifiers and anammox bacteria activity could be inhibited by the competition with cable bacteria and Beggiatoaceae for electron donors and acceptors (such as sulfide, nitrate and nitrite) before summer hypoxia but might be still supported by the potential sulfide detoxification of the sediment through sulfide oxidation, supporting at least a low activity. During summer hypoxia, sulfide inhibition and low nitrate and nitrite concentrations seem to limit the activity of hetero-trophic and autotrophic (sulfide-dependent) denitrifiers and anammox bacteria.

Additionally, we examined the changes in activity and structure of the sedimentary chem-olithoautotrophic bacterial community in Lake Grevelingen under oxic (spring) and hypoxic (summer) conditions (Chapter 5). Combined 16S rRNA gene amplicon sequencing and analysis of phospholipid derived fatty acids indicated a major temporal shift in community structure. Chemolithoautotrophy rates in the surface sediment were three times higher in spring compared to summer. Besides, the depth distribution of chemolithoautotrophy was linked to the distinct sulfur oxidation mechanisms identified through microsensor profiling, i.e., canonical sulfur ox-

Page 197: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

189

idation, electrogenic sulfur oxidation by cable bacteria, and sulfide oxidation coupled to nitrate reduction by Beggiatoaceae. The metabolic diversity of the sulfur-oxidizing bacterial community suggests a complex niche partitioning within the sediment probably driven by the availability of reduced sulfur compounds and electron acceptors regulated by seasonal hypoxia.

In Lake Grevelingen sediments, we also compared the archaeal abundance and composition by DNA-based methods and the archaeal intact polar lipid (IPL) diversity in surface sediments to determine the potential biological sources of the archaeal IPLs detected in surface sedi-ments under changing environmental conditions (Chapter 6). We observed a dramatic change in the archaeal community composition and lipid abundance in surface sediments of a season-ally hypoxic marine lake which corresponded to a switch from a Thaumarchaeota-dominated to a DPANN-dominated archaeal community, while the total IPLs detected were significantly reduced. Considering the reduced genome of the members of the superphylum DPANN and their apparent inability to synthesize their own membrane lipids, we hypothesize that they use the CLs previously synthesized by the Thaumarchaeota to form their membrane. Results indi-cate an active recycling of fossil IPLs in the marine surface sediment.

Overall, the research presented in this thesis has provided new insights on the diversity, abun-dance and activity of chemolithoautotrophic microorganisms, such as AOA, AOB, anammox and denitrifying bacteria, as well as sulfate reducing and sulfur oxidizing microorganisms, includ-ing sulfide-dependent denitrifiers in coastal marine sediments. Hypoxia and increasing sulfide concentration have proven to be key factors restricting the presence and activity of the above mentioned groups. In addition, the interactions with other microbial groups and bioturbation seem to favor the presence of specific chemolithoautotrophs in unexpected sediment depths, such as the presence of AOA in depths where the oxygen concentration was expected to be undetectable, or the presence of anammox bacteria in bioturbated sediments where nitrate re-ducers could thrive and provide nitrite to fuel the anammox process. In general, the methods employed in these studies regarding the gene expression and abundance of key metabolic genes of chemolithoautotrophs under different environmental and physiological conditions have shed further light on the role of this group in the carbon and nitrogen cycle in coastal sediments. However, the connection between gene expression and the activity of the actual process is still be problematic and future studies should address the regulation of specific metabolic genes un-der different physiological and environmental conditions. Finally, the application of specific lipid biomarkers to detect the presence and abundance of specific chemolithoautotrophs in coastal sediments (i.e. crenarchaeol for Thaumarchaeota, ladderane lipids for anammox) has revealed a potential preservation effect of these biomarkers in deeper sediments potentially invalidating their use as proxies of living biomass. Further studies are needed to fully assess this effect as well as the recycling of intact polar lipids by specific archaeal groups.

Page 198: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 199: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

191

Samenvatting

De rol die chemolithoautotrofe micro-organismen spelen in sedimenten van kustzeeën is nog steeds onduidelijk. Met name is het effect van het optreden van seizoensgebonden hypoxische/anoxische condities op de microbiële chemolithoautotrofe gemeenschappen niet onderzocht. In dit proefschrift is de diversiteit, celaantallen en activiteit van specifieke groepen chemolithoau-totrofe micro-organismen die actief zijn in de stikstof- en zwavelcyclus, in relatie tot de totale microbiële gemeenschap, in kustsedimenten onderzocht door gebruik te maken van op DNA/RNA en lipiden gebaseerde biomarker benaderingen. Specifiek gaat het om ammonia-oxideren-de archaea (AOA) en bacteriën (AOB), anammox en denitrificerende bacteriën, sulfaat-reducer-ende en zwavel-oxiderende micro-organismen (waaronder sulfide-afhankelijke denitrificerende bacteriën). Ook werden de diversiteit en activiteit van chemolithoautotrofe micro-organismen betrokken bij de CBB- en rTCA-cycli bestudeerd. De studie richtte zich op sedimenten van kustzeeën, d.w.z. (1) de Oestergronden in de centrale Noordzee, die gedurende de zomer bloot-gesteld zijn aan verlaagde zuurstofconcentraties, (2) verschillende soorten sedimenten (zand, klei en modder) van de IJslandse kustzeeën, en (3) hypoxische en sulfidische sedimenten van het zoute Grevelingenmeer (Nederland), blootgesteld aan seizoensgebonden hypoxia. De resul-taten laten het belang van chemolithoautotrofe micro-organismen in de biogeochemische cycli van koolstof, stikstof en zwavel zien. Seizoensgebonden veranderingen in zuurstofconcentratie leiden tot grote tijdelijke veranderingen in het aantal, de diversiteit, en activiteit van chemolitho-autotrofe micro-organismen.

In de gebioturbeerde sedimenten van de Oestergronden in de zuidelijke Noordzee zijn de seizoensvariaties en diepteverdeling in concentratie van AOA en AOB en anammox-bacteriën in relatie tot omgevingsfactoren bestudeerd (hoofdstuk 2). Specifieke intacte polaire lipiden (IPL) zijn gekwantificeerd. Daarnaast werden specifieke 16S rRNA-genen, het ammoniakmonoox-ygenase subeenheid A (amoA) gen van AOA en AOB, en het hydrazine synthase (hzsA) gen van anammox bacteriën en hun expressie gekwantificeerd. AOA, AOB en anammox bacteriën werden gedetecteerd en waren actief tot 12 cm diepte in het sediment. Het metabolisme van deze microbiële groepen is ruimtelijk gekoppeld op basis van zuurstofverbruik, hetgeen de coëx-istentie van aërobe en anaërobe ammonia-oxiderende bacteriën mogelijk maakt. AOA waren gedurende het hele jaar meer dominant aanwezig dan AOB, hetgeen waarschijnlijk veroorzaakt wordt door de hogere zuurstofaffiniteit van AOA in vergelijking met AOB. De celaantallen en activiteit van anammox bacteriën waren hoger in de zomer. Dit wordt waarschijnlijk veroorzaakt door hogere temperaturen en lagere zuurstofbeschikbaarheid in de sedimenten gedurende de periode van zomerstratificatie. Gedurende de zomer zijn de anammox bacteriën waarschijnlijk niet in concurrentie met AOA en AOB voor ammonia omdat de concentraties betrekkelijk hoog waren in het poriewater van het sediment, waarschijnlijk als gevolg van ammonificatie processen.

In sedimenten van IJslandse kustzeeën werden de activiteit van ammonia-oxiderende Thau-marchaeota onderzocht met behulp van verschillende 13C-gelabelde substraten, die in eerdere studies met incubatie met verrijkingsculturen en zeewater succesvol bleken (hoofdstuk 3). Sedi-mentkernen werden gedurende 4-6 dagen geïncubeerd met 13C-gelabeled bicarbonaat, pyruvaat, glucose en aminozuren en opname van het label in Thaumarchaeale in de membraanlipiden (intacte polaire glyceroldibiphytanylglyceroltetraethers; IPL-GDGTs) werd bepaald. De aan-wezigheid van labiele IPL-GDGT membraanlipiden karakteristiek voor Thaumarchaeota, zoals

Page 200: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

192

hexosefosfoshexose (HPH) -crenarchaeol, duidde op de aanwezigheid van levende Thaumar-chaeota. Incubatie met de 13C-gemerkte substraten leidde echter niet tot een substantiële inbouw van het 13C label in de bifytanylketens van deze membraanlipiden. Omzettijden en generatieti-jden van thaumarchaeale lipiden werden geschat op ten minste enkele jaren, hetgeen duidt op een trage groei van sedimentaire Thaumarchaeota of een langzame afbraak van IPL-GDGTs. Dit contrasteert met de bacteriële of eukaryotische membraanlipiden (vetzuren afkomstig van polaire lipiden) waar het label wel in terecht kwam.

In het sediment van het Grevelingenmeer werden de diversiteit, celaantallen en potentiële activiteit van denitrificerende, anammox en sulfide-afhankelijke denitrificerende bacteriën be-studeerd (hoofdstuk 4). De concentratie van 16S rRNA-genen, het nirS-gen van denitrificeren-de en anammox bacteriën, het aprA-gen van zwavel-oxiderende (SOB) en sulfaat-reducerende (SRB) bacteriën en ladderaanlipiden van anammox bacteriën werd bepaald in sedimenten tot 5 cm diepte op drie verschillende plaatsen en op twee verschillende momenten in het jaar. De condities in de sedimenten van het Grevelingenmeer worden sterk beïnvloed door het seizoens-gebonden optreden van hypoxia in het bodemwater. Deze studie toont de coëxistentie en po-tentiële activiteit van heterotrofe en autotrofe denitrificerende bacteriën, anammox bacteriën en SOB/SRB aan. De heterotrofe denitrificerende bacteriën van het nirS-type vormen een sta-biele gemeenschap, die onafhankelijk is van de veranderingen in zuurstof- en sulfideconcentratie gedurende de verschillende seizoenen. Hun celaantallen zijn even hoog als in sedimenten die niet gekenmerkt werden door hoge sulfideconcentraties. NirS-type denitrificerende bacteriën overtroffen in aantal de anammox bacteriën, hetgeen suggereert dat anammox niet significant bijdraagt aan stikstofverwijdering in sedimenten van het Grevelingenmeer. De gentranscriptie activiteit van denitrificerende en anammox bacteriën evenals van zwaveloxiderende, waaronder sulfide-afhankelijke denitrificerende en sulfaat-reducerende bacteriën, werden mogelijk door de accumulatie van sulfide in de zomer geïnhibeerd. Zowel de denitrificerende en anammox bac-teriën kunnen worden belemmerd door de concurrentie voor NO3

- en NO2- met kabel- en Beg-

giatoa-achtige bacteriën voor zomerhypoxia. Resultaten suggereren dat sulfide verwijdering door zwavel-oxiderende bacteriën onvoldoende is om de activiteit van denitrificerende en anammox bacteriën in de sedimenten van het Grevelingenmeer te ondersteunen.

In hoofdstuk 5 is het effect van seizoensgebonden hypoxia op de activiteit en de gemeen-schapsstructuur van chemolithoautotrofe bacteriën onderzocht in sedimenten van het Grevelin-genmeer voor (Maart) en tijdens (Augustus) het optreden van zomerhypoxia. Gecombineerde 16S rRNA gen amplicon sequencing en PLFA analyse onthulde een belangrijke temporale ver-schuiving in de gemeenschapsstructuur. Aërobe zwavel-oxiderende Gamma- en Epsilonproteo-bacteriën overheersten de chemolithoautotrofe gemeenschap in het voorjaar, terwijl Deltapro-teobacteriën veel meer voorkwamen tijdens de zomer. De bijdrage van chemolithoautotrofe bacteriën, bepaald m.b.v. PLFA-SIP, was drie maal hoger in de lente dan tijdens de zomer, vooral in oppervlaktesedimenten. De dominantie van cbbL (RuBisCO) en aclB (ATP-citraat lyase) genen liet een overheersing, onafhankelijk van het seizoen en plaats, van micro-organismen die de CBB cyclus gebruiken in vergelijking met de rTCA cyclus zien. De diepteverdeling in het sediment van chemolithoautotrofie was gekoppeld aan de verschillende zwavel-oxidatiemechanismen, die werden geïdentificeerd door middel van microsensorprofielen, d.w.z. kanonische zwavel-oxi-datie, elektrogenische zwavel-oxidatie door kabelbacteriën en sulfide-oxidatie gekoppeld aan nitraatreductie door Beggiatoaceae. De metabolische diversiteit van de zwavel-oxiderende bacte-

Page 201: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

193

riële gemeenschap suggereert een complexe verdeling van niches in het sediment, waarschijnlijk veroorzaakt door de beschikbaarheid van gereduceerde zwavelverbindingen (H2S, S0 en S2O3

2-) en elektronacceptoren (O2 en NO3

-), gereguleerd door het optreden van seizoensgebonden hy-poxia.

In de oppervlakte sedimenten van het Grevelingenmeer is de op DNA-gebaseerde archaeale diversiteit en distributie vergeleken met die op basis van archaeale intacte membraanlipiden om de potentiële biologische bronnen van de archaeale IPLs te detecteren (hoofdstuk 6). De ar-chaeale levensgemeenschap veranderde ten gevolge van de hypoxia van het bodemwater van een door Thaumarchaeota-gedomineerde in een door DPANN-gedomineerde archaeale levens-gemeenschap. Tegelijkertijd nam de concentratie IPL-GDGTs met een orde van grootte af. Archaea in het superphylum DPANN worden gekenmerkt door een klein genoom en bezitten geen genen die coderen voor eiwitten die de synthese van membraanlipiden katalyseren. Op ba-sis hiervan is de hypothese geponeerd dat zij zogenaamde core lipiden gebruiken die eerder door de Thaumarchaeota zijn gesynthetiseerd om hun membraan te vormen. Dit zou duiden op een actieve recycling van fossiele IPL-GDGTs in oppervlaktesedimenten.

Samenvattend heeft het onderzoek in dit proefschrift nieuwe inzichten verschaft met betrek-king tot de diversiteit, celaantallen en activiteit van chemolithoautotrofe micro-organismen. Hy-poxia en sulfideconcentratie zijn belangrijke factoren die hun aanwezigheid en activiteit beperken. Bovendien lijken de interacties met andere microbiële groepen en bioturbatie de aanwezigheid van specifieke chemolithoautotrofe microben op onverwachte sedimentdiepten te bevorderen, zoals de aanwezigheid van AOA in sedimentlagen waar zuurstof niet detecteerbaar was. In het algemeen hebben de methoden die in deze studies worden toegepast met betrekking tot de gen-expressie en de overvloed aan belangrijke metabolische genen van chemolithoautotrophs onder verschillende milieu- en fysiologische omstandigheden, de rol van deze groep in de koolstof- en stikstofcyclus in kustsedimenten verder benadrukt. De directe koppeling tussen genexpressie en activiteit is echter nog steeds niet eenduidig en toekomstige studies zullen de regulering van spec-ifieke metabolische genen onder verschillende fysiologische en milieuomstandigheden verder moeten bestuderen. Tenslotte heeft het onderzoek aangetoond dat specifieke lipide biomarkers de potentie hebben om specifieke chemolithoautotrofe microben in sedimenten van kustzeeën te detecteren (bijv. crenarchaeol voor Thaumarchaeota, ladderaan lipiden voor anammox bacte-riën). Omdat deze lipiden bewaard kunnen blijven in sedimenten hebben zij potentie om gebrui-kt te worden in de reconstructie van vroegere microbiële activiteit. De mogelijke recycling van intacte polaire lipiden door specifieke archaea is een proces dat dit gebruik mogelijk zou kunnen compliceren.

Page 202: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 203: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

195

References

Abell, G.C.J., Revill, A.T., Smith, C., Bissett, A.P., Volkman, J.K., Robert, S.S., 2010. Archaeal ammonia oxidizers and nirS-type denitrifiers dominate sediment nitrifying and denitrifying populations in a subtropical macrotidal estuary. The ISME Journal 4, 286–300.

Abell, G.C.J., Banks, J., Ross, D.J., Keane, J.P., Robert, S.S., Revill, A.T., Volkman, J.K., 2011. Effects of estuarine sediment hypoxia on nitrogen fluxes and ammonia oxidizer gene tran-scription. FEMS Microbiology Ecology 75, 111–122.

Aelion, C.M., Warttinger, U., 2010. Sulfide inhibition of nitrate removal in coastal sediments. Estuaries and Coasts 33, 798–803.

Agogué, H., Brink, M., Dinasquet, J., Herndl, G.J., 2008. Major gradients in putatively nitrifying and non-nitrifying Archaea in the deep North Atlantic. Nature 456, 788–791.

Alonso-Sáez, L., Waller, A.S., Mende, D.R., Bakker, K., Farnelid, H., Yager, P.L., Lovejoy, C., Tremblay, J.-É., Potvin, M., Heinrich, F., Estrada, M., Riemann, L., Bork, P., Pedrós-Alió, C., Bertilsson, S., 2012. Role for urea in nitrification by polar marine Archaea. Proceedings of the National Academy of Sciences of the USA 109, 17989–17994.

Altschul, S.F., Gish, W., Miller, W., Myers, E.W., Lipman, D.J., 1990. Basic local alignment search tool. Journal of Molecular Biology 215, 403–10.

Arp, D.J., Chain, P.S.G., and Klotz, M.G, 2007. The impact of genome analyses on our under-standing of ammonia- oxidising bacteria. Annual Review of Microbiology 61: 503–528.

Babbin, A.R., Keil, R.G., Devol, A.H., Ward, B.B., 2014. Organic matter stoichiometry, flux, and oxygen control nitrogen loss in the Ocean. Sciencexpress Reports 50. doi:10.1007/s002530051340.

Baker, B.J., Lesniewski, R.A., Dick, G.J., 2012. Genome-enabled transcriptomics reveals archaeal populations that drive nitrification in a deep-sea hydrothermal plume. The ISME Journal 6, 2269–2279.

Bale, N.J., Villanueva, L., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2013. Differ-ent seasonality of pelagic and benthic Thaumarchaeota in the North Sea. Biogeosciences 10, 7195–7206. doi:10.5194/bg-10-7195-2013.

Bale, N.J., Villanueva, L., Fan, H., Stal, L.J., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2014. Occurrence and activity of anammox bacteria in surface sediments of the south-ern North Sea. FEMS Microbiology Ecology 89, 99–110.

Bauersachs, T., Speelman, E.N., Hopmans, E.C., Reichart, G.-J., Schouten, S., Sinninghe Dam-sté, J.S., 2010. Fossilized glycolipids reveal past oceanic N2 fixation by heterocystous cya-nobacteria. Proceedings of the National Academy of Sciences of the USA 107, 1–5. doi:10.1073/pnas.1007526107.

Becker, P.T., Samadi, S., Zbinden, M., Hoyoux, C., Compère, P., Ridder, C.D.E., 2009. First insights into the gut microflora associated with an echinoid from wood falls environments. Cahiers De Biologie Marine 50, 343–352.

Behrens, S., Lösekann, T., Pett-Ridge, J., Weber, P.K., Ng, W.-O., Stevenson, B.S., Hutcheon, I.D., Relman, D.A., Spormann, A.M., 2008. Linking microbial phylogeny to metabolic ac-tivity at the single-cell level by using enhanced element labeling-catalyzed reporter deposi-tion fluorescence in situ hybridization (EL-FISH) and NanoSIMS. Applied and Environmental Microbiology 74, 3143–3150.

Beman, J.M., Francis, C.A., 2006. Diversity of ammonia-oxidizing archaea and bacteria in the sediments of a hypernutrified subtropical estuary: Bahía del Tóbari, Mexico. Applied and Environmental Microbiology 72, 7767–7777.

Page 204: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

196

Beman, J.M., Popp, B.N., Francis, C.A., 2008. Molecular and biogeochemical evidence for ammonia oxidation by marine Crenarchaeota in the Gulf of California. The ISME Journal 2, 429–441.

Beman, J.M., Bertics, V.J., Braunschweiler, T., and Wilson, J.M., 2012. Quantification of am-monia oxidation rates and the distribution of ammonia-oxidizing Archaea and Bacteria in marine sediment depth profiles from Catalina Island, California. Frontiers in Microbiology 3, 263. doi:10.3389/fmicb.2012.00263.

Beman, J.M., 2014. Activity, abundance, and diversity of nitrifying archaea and denitrifying bacteria in sediments of a subtropical estuary: Bahía del Tóbari, Mexico. Estuaries and Coasts 37, 1343–1352.

Berg, C., Vandieken, V., Thamdrup, B., Jürgens, K., 2015. Significance of archaeal nitrification in hypoxic waters of the Baltic Sea. The ISME Journal 9, 1319–1332.

Berg, I.A., Kockelkorn, D., Buckel, W., Fuchs, G., 2007. A 3-hydroxypropionate/4-hydroxybu-tyrate autotrophic carbon dioxide assimilation pathway in Archaea. Science 318, 1782–1786.

Berg, I.A., Kockelkorn, D., Vera, W.H.R., Say, R.F., Zarzycki, J., Hügler, M., Alber, B.E., Fuchs, G., 2010. Autotrophic carbon fixation in archaea. Nature Reviews Microbiology 8, 447–460.

Berg, I.A., 2011. Ecological aspects of the distribution of different autotrophic CO2 fixation pathways. Applied and Environmental Microbiology 77, 1925–1936.

Bernhard, A.E., Landry, Z.C., Blevins, A., de la Torre, J.R., Giblin, A.E., and Stahl, D.A., 2010 . Abundance of ammonia-oxidizing archaea and bacteria along an estuarine salinity gradient in relation to potential nitrification rates. Applied and Environmental Microbiology 76, 1285–1289.

Biddle, J.F., Lipp, J.S., Lever, M.A., Lloyd, K.G., Sørensen, K.B., Anderson, R., Fredricks, H.F., Elvert, M., Kelly, T.J., Schrag, D.P., Sogin, M.L., Brenchley, J.E., Teske, A., House, C.H., Hinrichs, K.-U., 2006. Heterotrophic Archaea dominate sedimentary subsurface ecosystems off Peru. Proceedings of the National Academy of Sciences of the USA 103, 3846–3851.

Blaga, C.I., Reichart, G.J., Heiri, O., Sinninghe Damsté, J.S., 2009. Tetraether membrane lipid distributions in water-column particulate matter and sediments: A study of 47 European lakes along a north-south transect. Journal of Paleolimnology 41, 523–540.

Blazejak, A., Schippers, A., 2011. Real-Time PCR Quantification and diversity analysis of the functional genes aprA and dsrA of sulfate-reducing prokaryotes in marine sediments of the Peru continental margin and the Black Sea. Frontiers in Microbiology 2, 253. doi:10.3389/fmicb.2011.00253.

Bligh, E.G., Dyer, W.J., 1959. A rapid method of total lipid extraction and purification. Canadi-an Journal of Biochemistry and Physiology 37, 911–917.

Blumer, M., Chase, T., Watson, S.W., 1969. Fatty acids in the lipids of marine and terrestrial nitrifying bacteria. Journal of Bacteriology 99, 366–370.

Böttcher, M.E., Thamdrup, B., Gehre, M., Theune, A., 2005. 34S/32S and 18O/16O fraction-ation during sulfur disproportionation by Desulfobulbus propionicus. Geomicrobiology Journal 22, 219–226.

Bollmann, A., Schmidt, I., Saunders, A.M., Nicolaisen, M.H., 2005. Influence of starvation on potential ammonia-oxidizing activity and amoA mRNA levels of Nitrosospira briensis. Applied and Environmental Microbiology 71, 1276–1282.

Bowman, J.P., McMeekin, T.A., 2005. Order X, Alteromodales ord. nov., p 443–591, in: Bren-ner, D., Krieg, N., Staley, J., Garrity, G. (ed) Bergey’s manual of systematic bacteriology, 2nd ed. vol 2, Springer New York.

Page 205: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

197

Boschker, H.T.S., Nold, S.C., Wellsbury, P., Bos, D., de Graaf, W., Pel, R., Parkes, R.J., Cap-penberg, T.E., 1998. Direct linking of microbial populations to specific biogeochemical processes by 13C-labelling of biomarkers. Letters to Nature 392, 801–805.

Boschker, H.T.S., Middelburg, J.J., 2002. Stable isotopes and biomarkers in microbial ecology. FEMS Microbiology Ecology 40, 85–95.

Boschker, H.T.S., Vasquez-Cardenas, D., Bolhuis, H., Moerdijk-Poortvliet, T.W.C., Moodley, L., 2014. Chemoautotrophic carbon fixation rates and active bacterial communities in intertidal marine sediments. PLoS ONE 9, e101443. doi:10.1371/journal.pone.0101443.

Boumann, H.A., Hopmans, E.C., van de Leemput, I., Op den Camp, H.J.M., van de Vossen-berg, J., Strous, M., Jetten, M.S.M., Sinninghe Damsté, J.S., Schouten, S., 2006. Ladderane phospholipids in anammox bacteria comprise phosphocholine and phosphoethanolamine headgroups. FEMS Microbiology Letters 258, 297–304.

Bowen, J.L., Babbin, A.R., Kearns, P.J., Ward, B.B., 2014. Connecting the dots: linking nitrogen cycle gene expression to nitrogen fluxes in marine sediment mesocosms. Frontiers in Microbi-ology 5, 429. doi:10.3389/fmicb.2014.00429.

Bowles, M.W., Nigro, L.M., Teske, A.P., Joye, S.B., 2012. Denitrification and environmental factors influencing nitrate removal in Guaymas Basin hydrothermally altered sediments. Frontiers in Microbiology 3, 377. doi:10.3389/fmicb.2012.00377.

Braker, G., Fesefeldt, A., 1998. Development of PCR primer systems for amplification of ni-trite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Applied and Environmental Microbiology 64, 3769–3775.

Braker, G., Zhou, J., Wu, L., Devol, A.H., Tiedje, J.M., 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific North-west marine sediment communities. Applied and Environmental Microbiology 66, 2096–2104.

Brandsma, J., van de Vossenberg, J., Risgaard-Petersen, N., Schmid, M.C., Engström, P., Eure-nius, K., Hulth, S., Jaeschke, A., Abbas, B., Hopmans, E.C., Strous, M., Schouten, S., Jetten, M.S.M., Sinninghe Damsté, J.S., 2011. A multi-proxy study of anaerobic ammonium oxida-tion in marine sediments of the Gullmar Fjord, Sweden. Environmental Microbiology Reports 3, 360–366.

Brochier-Armanet, C., Boussau, B., Gribaldo, S., Forterre, P., 2008. Mesophilic Crenarchaeota: proposal for a third archaeal phylum, the Thaumarchaeota. Nature Reviews Microbiology 6, 245–252.

Brunet, R.C., Garcia-Gil, L.J., 1996. Sulfide-induced dissimilatory nitrate reduction to ammonia in anaerobic freshwater sediments. FEMS Microbiology Ecology 21, 131–138.

Brüchert, V., Jørgensen, B.B., Neumann, K., Riechmann, D., Schlösser, M., Schulz, H., 2003. Regulation of bacterial sulfate reduction and hydrogen sulfide fluxes in the central Namib-ian coastal upwelling zone. Geochimica et Cosmochimica Acta 67, 4505–4518.

Bulow, S.E., Francis, C.A., Jackson, G.A., Ward, B.B., 2008. Sediment denitrifier community composition and nirS gene expression investigated with functional gene microarrays. Envi-ronmental Microbiology 10, 3057–3069.

Burgin, A.J., Hamilton, S.K., 2007. Have we overemphasized the role of denitrification in aquatic ecosystems? A review of nitrate removal pathways. Frontiers in Ecology and the Envi-ronment 5, 89–96.

Campbell, B.J., Stein, J.L., Cary, S.C., 2003. Evidence of chemolithoautotrophy in the bacterial community associated with Alvinella pompejana, a hydrothermal vent polychaete. Applied and Environmental Microbiology 69, 5070–5078.

Page 206: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

198

Campbell, B.J., Engel, A.S., Porter, M.L., Takai, K., 2006. The versatile ε-proteobacteria : key players in sulphidic habitats. Nature Reviews Microbiology 4, 458–468.

Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F.D., Costello, E.K., Fierer, N., Peña, A.G., Goodrich, J.K., Gordon, J.I., Huttley, G.A., Kelley, S.T., Knights, D., Koenig, J.E., Ley, R.E., Lozupone, C.A., Mcdonald, D., Muegge, B.D., Pirrung, M., Reeder, J., Sevinsky, J.R., Turnbaugh, P.J., Walters, W.A., Widmann, J., Yatsunenko, T., Zaneveld, J., Knight, R., 2010. Correspondence QIIME allows analysis of high- throughput communi-ty sequencing data Intensity normalization improves color calling in SOLiD sequencing. Nature Methods 7, 335–336.

Castelle, C.J., Wrighton, K.C., Thomas, B.C., Hug, L.A., Brown, C.T., Wilkins, M.J., Frischkorn, K.R., Tringe, S.G., Singh, A., Markillie, L.M., Taylor, R.C., Williams, K.H., Banfield, J.F., 2015. Genomic expansion of domain archaea highlights roles for organisms from new phyla in anaerobic carbon cycling. Current Biology 25, 690–701.

Church, M.J., Wai, B., Karl, D.M., DeLong, E.F., 2010. Abundances of crenarchaeal amoA genes and transcripts in the Pacific Ocean. Environmental Microbiology 12, 679–688.

Cline, J.D., 1969. Spectrophotometric determination of hydrogen sulfide. Limnology and Oceanog-raphy 14, 454–458.

Comolli, L.R., Baker, B.J., Downing, K.H., Siegerist, C.E., Banfield, J.F., 2009. Three-dimen-sional analysis of the structure and ecology of a novel, ultra-small archaeon. The ISME Journal 3, 159–167.

Coolen, M.J.L., Abbas, B., van Bleijswijk, J., Hopmans, E.C., Kuypers, M.M.M., Wakeham, S.G., Sinninghe Damsté, J.S., 2007. Putative ammonia-oxidizing Crenarchaeota in suboxic waters of the Black Sea: a basin-wide ecological study using 16S ribosomal and functional genes and membrane lipids. Environmental Microbiology 9, 1001–1016.

Crowe, S.A., Canfield, D.E., Mucci, A., Sundby, B., Maranger, R., 2012. Anammox, denitri-fication and fixed-nitrogen removal in sediments from the Lower St. Lawrence Estuary. Biogeosciences 9, 4309–4321. doi:10.5194/bg-9-4309-2012.

Cruaud, P., Vigneron, A., Lucchetti-Miganeh, C., Ciron, P.E., Godfroy, A., Cambon-Bonavi-ta, M.A., 2014. Influence of DNA extraction method, 16S rRNA targeted hypervariable regions, and sample origin on microbial diversity detected by 454 pyrosequencing in marine chemosynthetic ecosystems. Applied and Environmental Microbiology 80, 4626–4639.

Cypionka, H., 2006. Lithotrophie-Verwertung anorganischer Elektronendonatoren, 193–202, in: Grundlagen der Mikrobiologie, 3rd ed., Springer-Verlag Berlin Heidelberg New York.

Dale, O.R., Tobias, C.R., Song, B., 2009. Biogeographical distribution of diverse anaerobic am-monium oxidizing (anammox) bacteria in Cape Fear River Estuary. Environmental Microbiology 11, 1194–1207.

Dalsgaard, T., Thamdrup, B., 2002. Factors controlling anaerobic ammonium oxidation with nitrite in marine sediments. Applied and Environmental Microbiology 68, 3802–3808.

Dalsgaard, T., Canfield, D.E., Petersen, J., Thamdrup, B., Acuña-González, J., Acuna-Gonza-lez, J., 2003. N2 production by the anammox reaction in the anoxic water column of Golfo Dulce, Costa Rica. Nature 422, 606–608.

Dalsgaard, T., Thamdrup, B., Canfield, D.E., 2005. Anaerobic ammonium oxidation (anam-mox) in the marine environment. Research in Microbiology 156, 457–464.

Dang, H., Wang, C., Li, J., Li, T., Tian, F., Jin, W., Ding, Y., Zhang, Z., 2009. Diversity and distribution of sediment nirS-encoding bacterial assemblages in response to environmental gradients in the eutrophied Jiaozhou Bay, China. Microbial Ecology 58, 161–169.

Page 207: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

199

Dang, H., Chen, R., Wang, L., Guo, L., Chen, P., Tang, Z., Tian, F., Li, S., Klotz, M.G., 2010. Environmental factors shape sediment anammox bacterial communities in hypernutrified Jiaozhou Bay, China. Applied and Environmental Microbiology 76, 7036–7047.

Dang, H., Zhou, H., Yang, J., Ge, H., Jiao, N., Luan, X., Zhang, C., Klotz, M.G., 2013. Thau-marchaeotal signature gene distribution in sediments of the northern South China Sea: an indicator of the metabolic intersection of the marine carbon, nitrogen, and phosphorus cycles? Applied and Environmental Microbiology 79, 2137–2147.

DeLong, E.F., 1992. Archaea in coastal marine environments. Proceedings of the National Academy of Sciences of the USA 89, 5685–5689.

DeSantis, T.Z., Hugenholtz, P., Larsen, N., Rojas, M., Brodie, E.L., Keller, K., Huber, T., Dalevi, D., Hu, P., Andersen, G.L., 2006. Greengenes, a chimera-checked 16S rRNA gene database and workbench compatible with ARB. Applied and Environmental Microbiology 72, 5069–5072.

Devol, A.H., 2015. Denitrification, anammox and N2 production in marine sediments. Annual Review of Marine Science 7, 403–423.

De Wilde, P.A.W.J., Berghuis, E.M., Kok, A., 1984. Structure and energy demand of the ben-thic community of the Oyster Ground, central North Sea. Netherlands Journal of Sea Research 18, 143–159.

Diaz, R.J., Rosenberg, R., 2008. Spreading dead zones and consequences for marine ecosys-tems. Science 321, 926–929.

Donachie, S.P., Bowman, J.P., On, S.L.W., Alam, M., 2005. Arcobacter halophilus sp. nov., the first obligate halophile in the genus Arcobacter. International Journal of Systematic and Evolutionary Microbiology 55, 1271–1277.

Donders, T.H., Guasti, E., Bunnik, F.P.M., van Aken, H., 2012. Impact van de Brouwersdam op Zuurstofcondities in de Grevelingen; Reconstructies uit natuurlijke Sediment Archieven. TNO-060-UT-2011-02116, Geological Survey of the Netherlands.

Dong, L.F., Nedwell, D.B., Underwood, G.J.C., Thornton, D.C.O., Rusmana, I., 2002. Nitrous oxide formation in the Colne Estuary, England: the central role of nitrite. Applied and Envi-ronmental Microbiology 68, 1240–1249.

Dong, L.F., Thornton, D.C.O., Nedwell, D.B., Underwood, G.J.C., 2000. Denitrification in sedi-ments of the River Colne estuary, England. Marine Ecology Progress Series 203, 109–122.

Duineveld, G.C.A., Konitzer, A., Niermann, U., De Wilde, P.A.W.J., and Gray, J.S., 1991. The macrobenthos of the North Sea. Netherlands Journal of Sea Research 28, 53–65.

Dyksma, S., Bischof, K., Fuchs, B.M., Hoffmann, K., Meier, D., Meyerdierks, A., Pjevac, P., Probandt, D., Richter, M., Stepanauskas, R., Mußmann, M., 2016. Ubiquitous Gammapro-teobacteria dominate dark carbon fixation in coastal sediments. The ISME Journal 10, 1939–1953.

Edgar, R.C., 2004. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research 32, 1792–1797.

Edgar, R.C., 2010. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461.

Edlund, A., Nichols, P.D., Roffey, R., White, D.C., 1985. Extractable and lipopolysaccharide fatty acid and hydroxy acid profiles from Desulfovibrio species. Journal of Lipid Research 26, 982–988.

Engström, P., Dalsgaard, T., Hulth, S., and Aller, R.C., 2005. Anaerobic ammonium oxidation by nitrite (anammox): Implications for N2 production in coastal marine sediments. Geochimi-ca et Cosmochimica Acta 69, 2057–2065.

Page 208: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

200

Erguder, T.H., Boon, N., Wittebolle, L., Marzorati, M., and Verstraete, W., 2009. Environmen-tal factors shaping the ecological niches of ammonia-oxidizing archaea. FEMS Microbiology Reviews 33, 855–869.

Ellen, A.F., Albers, S.V., Huibers, W., Pitcher, A., Hobel, C.F. V, Schwarz, H., Folea, M., Schouten, S., Boekema, E.J., Poolman, B., Driessen, A.J.M., 2009. Proteomic analysis of secreted membrane vesicles of archaeal Sulfolobus species reveals the presence of endosome sorting complex components. Extremophiles 13, 67–79.

Elling, F.J., Könneke, M., Lipp, J.S., Becker, K.W., Gagen, E.J., Hinrichs, K.-U., 2014. Effects of growth phase on the membrane lipid composition of the thaumarchaeon Nitrosopumilus maritimus and their implications for archaeal lipid distributions in the marine environment. Geochimica et Cosmochimica Acta 141: 579–597.

Enoksson, V., Samuelsson, M.-O., 1987. Nitrification and dissimilatory ammonium production and their effects on nitrogen flux over the sediment-water interface in bioturbated coastal sediments. Marine Ecology Progress Series 36, 181–189.

Finster, K., Liesack, W., Thamdrup, B., 1998. Elemental sulfur and thiosulfate disproportion-ation by Desulfocapsa sulfoexigens sp. nov., a new anaerobic bacterium isolated from marine surface sediment. Applied and Environmental Microbiology 64, 119–125.

Francis, C.A., Roberts, K.J., Beman, J.M., Santoro, A.E., Oakley, B.B., 2005. Ubiquity and diver-sity of ammonia-oxidizing archaea in water columns and sediments of the ocean. Proceedings of the National Academy of Sciences of the USA 102, 14683–14688.

Fry, B., 2006. Stable isotope ecology. Springer New York. Fuhrman, J.A., McCallum, K., Davis, A.A., 1992. Novel major archaebacterial group from

marine plankton. Letters to Nature 356, 472–475. Fuhrman, J.A., Davis, A.A., 1997. Widespread Archaea and novel Bacteria from the deep sea as

shown by 16S rRNA gene sequences. Marine Ecology Progress Series 150, 275–285. Gattinger, A., Schloter, M., Munch, J.C., 2002. Phospholipid etherlipid and phospholipid fatty

acid fingerprints in selected euryarchaeotal monocultures for taxonomic profiling. FEMS Microbiology Letters 213, 133–139.

Geets, J., Boon, N., Verstraete, W. 2006. Strategies of aerobic ammonia-oxidizing bacteria for coping with nutrient and oxygen fluctuations. FEMS Microbiology Ecology 58, 1–13.

Gilbert, J.A., Dupont, C.L., 2011. Microbial Metagenomics: Beyond the Genome. Annual Re-view of Marine Science 3, 347–371.

Graf, D.R.H., Jones, C.M., Hallin, S., 2014. Intergenomic comparisons highlight modularity of the denitrification pathway and underpin the importance of community structure for N2O emissions. PLoS ONE 9, 1–20. doi:10.1371/journal.pone.0114118

Grasshoff, K., Kremlin, K., Ehrhardt, M., 1983. Methods of seawater analysis. Verlag Chemie GmbH, Weiheim (Germany).

Greenwood, N., Parker, E.R., Fernand, L., Sivyer, D.B., Weston, K., Painting, S.J., Kröger, S., Forster, R.M., Lees, H.E., Mills, D.K., Laane, R.W.P.M., 2010. Detection of low bottom water oxygen concentrations in the North Sea; implications for monitoring and assessment of ecosystem health. Biogeosciences 7, 1357–1373. doi:10.5194/bg-7-1357-2010.

Grote, J., Schott, T., Bruckner, C.G., Glöckner, F.O., Jost, G., Teeling, H., Labrenz, M., Jürgens, K., 2012. Genome and physiology of a model Epsilonproteobacterium responsible for sulfide detoxification in marine oxygen depletion zones. Proceedings of the National Academy of Sciences of the USA 109, 506–510.

Gruber, N., Galloway, J.N., 2008. An Earth-system perspective of the global nitrogen cycle. Nature 451, 293–296.

Page 209: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

201

Guilini, K., Oevelen, D. Van, Soetaert, K., Middelburg, J.J., Vanreusel, A., 2010. Nutritional importance of benthic bacteria for deep-sea nematodes from the Arctic ice margin: Results of an isotope tracer experiment. Limnology and Oceanography 55, 1977–1989.

Guckert, J.B., Antworth, C.P., Nichols, P.D., White, D.C., 1985. Phospholipid, ester-linked fatty acid profiles as reproducible assays for changes in prokaryotic community structure of estu-arine sediments. FEMS Microbiology Ecology 31: 147–158.

Hafenbradl, D., Keller, M., Dirmeier, R., Rachel, R., Roßnagel, P., Burggraf, S., Huber, H., Stetter, K.O., 1996. Ferroglobus placidus gen. nov., sp. nov., a novel hyperthermophilic archae-um that oxidizes Fe2+ at neutral pH under anoxic conditions. Archives of Microbiology 166, 308–314.

Hagen, K.D., Nelson, D.C., 1996. Organic carbon utilization by obligately and facultatively au-totrophic Beggiatoa strains in homogeneous and gradient cultures. Applied and Environmental Microbiology 62, 947–953.

Hagens, M., Slomp, C.P., Meysman, F.J.R., Seitaj, D., Harlay, J., Borges, A.V., Middelburg, J.J., 2015. Biogeochemical processes and buffering capacity concurrently affect acidification in a seasonally hypoxic coastal marine basin. Biogeosciences 12, 1561–1583. doi:10.5194/bg-12-1561-2015.

Hallam, S.J., Mincer, T.J., Schleper, C., Preston, C.M., Roberts, K., Richardson, P.M., Delong, E.F., 2006. Pathways of carbon assimilation and ammonia oxidation suggested by environ-mental genomic analyses of marine Crenarchaeota. PLoS Biology 4, e95. doi:10.1371/jour-nal.pbio.0040095.

Harhangi, H.R., Le Roy, M., van Alen, T., Hu, B.-L., Groen, J., Kartal, B., Tringe, S.G., Quan, Z.-X., Jetten, M.S.M., Op den Camp, H.J.M., 2012. Hydrazine synthase, a unique phylo-marker with which to study the presence and biodiversity of anammox bacteria. Applied and Environmental Microbiology 78, 752–758.

Harvey, H.R., Fallon, R.D., Patton, J.S., 1986. The effect of organic matter and oxygen on the degradation of bacterial membrane lipids in marine sediments. Geochimica et Cosmochimica Acta 50, 795–804.

Hatzenpichler, R., 2012. Diversity, physiology, and niche differentiation of ammonia-oxidizing archaea. Applied and Environmental Microbiology 78, 7501–7510.

Hedges, J.I., Keil, R.G., 1995. Sedimentary organic matter preservation: an assessment and speculative synthesis. Marine Chemistry 49, 81–115.

Hedges, J.I., Keil, R.G., 1999. Sedimentary organic matter preservation: A test for selective degradation under oxic conditions. American Journal of Science 299, 529–555.

Heinzelmann, S.M., Bale, N.J., Hopmans, E.C., Sinninghe Damsté, J.S., Schouten, S., van der Meer, M.T.J., 2014. Critical assessment of glyco- and phospholipid separation by using silica chromatography. Applied and Environmental Microbiology 80, 360–365.

Helder, W., de Vries, R., 1979. An automatic phenol-hypochlorite method for the determina-tion of ammonia in sea- and brackish waters. Netherlands Journal of Sea Reasearch 13, 154–160.

Herbert, R.A., 1999. Nitrogen cycling in coastal marine ecosystems. FEMS Microbiology Reviews 23, 563–590.

Herfort, L., Schouten, S., Abbas, B., Veldhuis, M.J.W., Coolen, M.J.L., Wuchter, C., Boon, J.P., Herndl, G.J., Sinninghe Damsté, J.S., 2007. Variations in spatial and temporal distribution of Archaea in the North Sea in relation to environmental variables. FEMS Microbiology Ecology 62, 242–257.

Page 210: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

202

Herndl, G.J., Reinthaler, T., Teira, E., Aken, H. Van, Veth, C., Pernthaler, A., Pernthaler, J., 2005. Contribution of archaea to total prokaryotic production in the deep Atlantic Ocean. Applied and Environmental Microbiology 71, 2303–2309.

Hershberger, K.L., Barns, S.M., Reysenbach, A.-L., Dawson, S.C., Pace, N.R., 1996. Wide diver-sity of Crenarchaeota. Nature 384, 420.

Hoefs, M.J.L., Schouten, S., de Leeuw, J.W., King, L.L., 1997. Ether lipids of planktonic Ar-chaea in the marine water column. Applied and Environmental Microbiology 63, 3090–3095.

Hofmann, A.F., Soetaert, K., Middelburg, J.J., Meysman, F.J.R., 2010. AquaEnv: An aquatic acid-base modelling environment in R. Aquatic Geochemistry 16, 507-546.

Holmes, D.E., Nevin, K.P., Lovley, D.R., Holmes, E.T., 2004. In situ expression of nifD in Geo-bacteraceae in subsurface sediments. Applied and Environmental Microbiology 70, 7251–7259.

Hopmans, E.C., Schouten, S., Pancost, R.D., van der Meer, M.T., Sinninghe Damsté, J.S., 2000. Analysis of intact tetraether lipids in archaeal cell material and sediments by high perfor-mance liquid chromatography/atmospheric pressure chemical ionization mass spectrome-try. Rapid Communications in Mass Spectrometry 14, 585–589.

Hopmans, E.C., Kienhuis, M.V.M., Rattray, J.E., Jaeschke, A., Schouten, S., Sinninghe Damsté, J.S., 2006. Improved analysis of ladderane lipids in biomass and sediments using high-per-formance liquid chromatography/atmospheric pressure chemical ionization tandem mass spectrometry. Rapid Communications in Mass Spectrometry 20, 2099–2103.

Hornek, R., Pommerening-Röser, A., Koops, H.-P., Farnleitner, A.H., Kreuzinger, N., Kirschner, A., Mach, R.L., 2006. Primers containing universal bases reduce multiple amoA gene specific DGGE band patterns when analysing the diversity of beta-ammonia oxidizers in the environment. Journal of Microbiological Methods 66, 147–155.

Howarth, R. W., 1984. The ecological significance of sulfur in the energy dynamics of salt marsh and coastal marine sediments, Biogeochemistry 1, 5–27, doi:10.1007/BF02181118.

Howarth, R.W., Marino, R., 2006. Nitrogen as the limiting nutrient for eutrophication in coastal marine ecosystems: Evolving views over three decades. Limnology and Oceanography 51, 364–376.

Huang, S., Chen, C., Yang, X., Wu, Q., Zhang, R., 2011. Distribution of typical denitrifying functional genes and diversity of the nirS-encoding bacterial community related to envi-ronmental characteristics of river sediments. Biogeosciences 8, 3041–3051. doi:10.5194/bg-8-3041-2011.

Huber, R., Dyba, D., Huber, H., Burggraf, S., Rachel, R., 1998. Sulfur-inhibited Thermosphaera aggregans sp. nov., a new genus of hyperthermophilic archaea isolated after its prediction from environmentally derived 16S rRNA sequences. International Journal of Systematic Bacteri-ology 48, 31–38.

Hügler, M., Wirsen, C.O., Fuchs, G., Craig, D., Sievert, S.M., Hu, M., Taylor, C.D., 2005. Evi-dence for autotrophic CO2 fixation via the reductive tricarboxylic acid cycle by members of the ε subdivision of Proteobacteria. Journal of Bacteriology 187, 3020–3027.

Hügler, M., Sievert, S.M., 2011. Beyond the Calvin Cycle: Autotrophic Carbon Fixation in the Ocean. Annual Review of Marine Science 3, 261–289.

Huguet, C., Hopmans, E.C., Febo-Ayala, W., Thompson, D.H., Sinninghe Damsté, J.S., Schouten, S., 2006a. An improved method to determine the absolute abundance of glycerol dibiphytanyl glycerol tetraether lipids. Organic Geochemistry 37, 1036–1041.

Huguet, C., Cartes, J.E., Sinninghe Damsté, J.S., Schouten, S., 2006b. Marine crenarchaeotal membrane lipids in decapods: Implications for the TEX86 paleothermometer. Geochemistry, Geophysics, Geosystems 7, Q11010.

Page 211: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

203

Huguet, C., Schimmelmann, A., Thunell, R., Lourens, L.J., Sinninghe Damsté, J.S., Schouten, S., 2007. A study of the TEX86 paleothermometer in the water column and sediments of the Santa Barbara Basin, California. Paleoceanography 22, PA3203. doi:10.1029/2006PA001310.

Huguet, C., Martens-Habbena, W., Urakawa, H., Stahl, D.A., Ingalls, A.E., 2010. Comparison of extraction methods for quantitative analysis of core and intact polar glycerol dialkyl glycerol tetraethers (GDGTs) in environmental samples. Limnology and Oceanography Methods 8, 127–145.

Inagaki, F., Takai, K., Kobayashi, H., Nealson, K., Horikoshi, K., 2003. Sulfurimonas autotrophica gen. nov., sp. nov., a novel sulfur-oxidizing ε -proteobacterium isolated from hydrothermal sediments in the Mid-Okinawa Trough. International Journal of Systematic and Evolutionary Microbiology 53, 1801–1805.

Ingalls, A.E., Shah, S.R., Hansman, R.L., Aluwihare, L.I., Santos, G.M., Druffel, E.R.M., Pear-son, A., 2006. Quantifying archaeal community autotrophy in the mesopelagic ocean using natural radiocarbon. Proceedings of the National Academy USA 103, 6442–6447.

Ishii, K., Mußmann, M., Macgregor, B.J., Amann, R., 2004. An improved fluorescence in situ hybridization protocol for the identification of bacteria and archaea in marine sediments. FEMS Microbiology Ecology 50, 203–212.

Jahn, U., Summons, R., Sturt, H., Grosjean, E., Huber, H., 2004. Composition of the lipids of Nanoarchaeum equitans and their origin from its host Ignicoccus sp. strain KIN4/I. Archives of Microbiology 182, 404–413.

Jaeschke, A., Rooks, C., Trimmer, M., Nicholls, J.C., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2009. Comparison of ladderane phospholipid and core lipids as indicators for anaerobic ammonium oxidation (anammox) in marine sediments. Geochimica et Cosmochimica Acta 73, 2077–2088.

Jaeschke, A., Abbas, B., Zabel, M., Hopmans, E.C., Schouten, S., Damsté, J.S., 2010. Molecular evidence for anaerobic ammonium – oxidizing (anammox) bacteria in continental shelf and slope sediments off northwest Africa. Limnology and Oceanography 55, 365–376.

Jarrell, K.F., Walters, A.D., Bochiwal, C., Borgia, J.M., Dickinson, T., Chong, J.P.J., 2011. Major players on the microbial stage: Why Archaea are important. Microbiology 157, 919–936.

Jensen, M.M., Kuypers, M.M.M., Lavik, G., 2008. Rates and regulation of anaerobic ammoni-um oxidation and denitrification in the Black Sea. Limnology and Oceanography 53, 23–36.

Jessen, G.L., Lichtschlag, A., Struck, U., Boetius, A., 2016. Distribution and composition of thiotrophic mats in the hypoxic zone of the Black Sea (150–170 m water depth, Crimea margin). Frontiers in Microbiology 7, 1011. doi:10.3389/fmicb.2016.01011.

Jessen, G.L., Lichtschlag, A., Ramette, A., Pantoja, S., Rossel, P.E., Schubert, C.J., Struck, U., Boetius, A., 2017. Hypoxia causes preservation of labile organic matter and changes sea-floor microbial community composition (Black Sea). Science Advances 3, e1601897.

Jetten, M.S.M., van Niftrik, L., Strous, M., Kartal, B., Keltjens, J.T., Op den Camp, H.J.M., 2009. Biochemistry and molecular biology of anammox bacteria. Critical Reviews in Biochemistry and Molecular Biology, 1–20. doi:10.1080/10409230902722783.

Jørgensen, B.B., 1978. Comparison of methods for the quantification of bacterial sulphate reduction in coastal marine-sediments I. Measurement with radio- tracer techniques. Geomi-crobiology Journal 1: 11–27.

Jørgensen, B.B., 1982a. Ecology of the bacteria of the sulphur cycle with special reference to anoxic-oxic interface environments. Biological Sciences 298, 543–561.

Jørgensen, B.B., 1982b. Mineralization of organic matter in the sea bed – the role of sulphate reduction. Nature 296, 643–645.

Page 212: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

204

Jørgensen, B.B., 2006. Bacteria and Marine Biogeochemistry, 169–206, in: Schulz, H.D., Zabel, M. (ed) Marine Geochemistry. Springer Berlin Heidelberg.

Jørgensen, B.B., Kasten, S., 2006. Sulfur Cycling and Methane Oxidation, 271–309, in: Schulz, H.D., Zabel, M. (ed) Marine Geochemistry. Springer Berlin Heidelberg.

Jørgensen, B.B., Nelson, D.C., 2004. Sulfide oxidation in marine sediments: Geochemistry meets microbiology, 63–81, in: Amend, J.P., Edwards, K.J., Lyons, T.W. (ed) Sulfur Biogeo-chemistry-Past and Present. Geological Society of America.

Jørgensen, B.B., 2010. Big sulfur bacteria. The ISME Journal 4, 1083–1084. Jørgensen, B.B., Dunker, R., Grünke, S., Røy, H., 2010. Filamentous sulfur bacteria, Beggiatoa

spp., in arctic marine sediments (Svalbard, 79 degrees N). FEMS Microbiology Ecology 73, 500–513.

Joye, S.B., Anderson, I.C., 2008. Nitrogen cycling in coastal sediments, 867–915, in: Capone, D.G., Bronk, D.A., Mulholland, M.R., Carpenter, E.J. (ed) Nitrogen in the Marine Environ-ment, 2nd ed., Elsevier Inc., Amsterdam.

Joye, S.B., Hollibaugh, J.T., 1995. Influence of Sulfide Inhibition of Nitrification on Nitrogen Regeneration in Sediments. Science 270, 623–625.

Jung, M., Park, S., Min, D., Kim, J., Rijpstra, W.I.C., Kim, G., Madsen, E.L., Rhee, S.-K., 2011. Enrichment and characterization of an autotrophic ammonia-oxidizing archaeon of meso-philic crenarchaeal group I.1a from an agricultural soil. Applied and Environmental Microbiology 77, 8635–8647.

Jukes, T.H., and Cantor, C.R., 1969. Evolution of protein molecules, 21-132, in: Munro, H.N. (ed) Mammalian Protein Metabolism, Academic Press, New York.

Kaneko, M., Kitajima, F., Naraoka, H., 2011. Stable hydrogen isotope measurement of archaeal ether-bound hydrocarbons. Organic Geochemistry 42, 166–172.

Karner, M.B., DeLong, E.F., Karl, D.M., 2001. Archaeal dominance in the mesopelagic zone of the Pacific Ocean. Nature 409, 507–510.

Kartal, B., Maalcke, W.J., de Almeida, N.M., Cirpus, I., Gloerich, J., Geerts, W., Op den Camp, H.J.M., Harhangi, H.R., Janssen-Megens, E.M., Francoijs, K.-J., Stunnenberg, H.G., Kelt-jens, J.T., Jetten, M.S.M., Strous, M., 2011. Molecular mechanism of anaerobic ammonium oxidation. Nature 479, 127–130.

Kalvelage, T., Jensen, M.M., Contreras, S., Revsbech, N.P., Lam, P., Günter, M. et al., 2011. Oxygen sensitivity of anammox and coupled N-cycle processes in oxygen minimum zones. PLoS One 6, e29299. doi:10.1371/journal.pone.0029299.

Keldermann, P., Nieuwenhuize, J., Meerman-van der Repe, A.M., Van Liere, J.M., 1984. Chang-es of sediment distribution patterns in Lake Grevelingen, an enclosed estuary in the SW Netherlands. Netherlands Journal of Sea Research 18, 273–285.

Keil, R.G., Montlucon, D.B., Prahl, F.G., Hedges, J.I., 1994. Sorptive preservation of organic matter in marine sediments. Letters to Nature 370, 549–551.

Kim, B.H., Gadd, G.M., 2008. Chemolithotrophy, 354–386, in: Bacterial Physiology and Me-tabolism. Cambridge University Press.

Kim, J.H., van der Meer, J., Schouten, S., Helmke, P., Willmott, V., Sangiorgi, F., Koç, N., Hop-mans, E.C., Sinninghe Damsté, J.S., 2010. New indices and calibrations derived from the distribution of crenarchaeal isoprenoid tetraether lipids: Implications for past sea surface temperature reconstructions. Geochimica et Cosmochimica Acta 74, 4639–4654.

Page 213: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

205

Kim, J.-G., Jung, M.-Y., Park, S.-J., Rijpstra, W.I.C., Sinninghe Damsté, J.S., Madsen, E.L., Min, D., Kim, J.-S., Kim, G.-J., Rhee, S.-K., 2012. Cultivation of a highly enriched ammonia-ox-idizing archaeon of thaumarchaeotal group I.1b from an. Environmental Microbiology 14, 1528–1543.

Kim, J.-G., Park, S.-J., Sinninghe Damsté, J.S., Schouten, S., Rijpstra, W.I.C., Jung, M.-Y., Kim, S.-J., Gwak, J.-H., Hong, H., Si, O.-J., Lee, S., Madsen, E.L., Rhee, S.-K., 2016. Hydrogen peroxide detoxification is a key mechanism for growth of ammonia-oxidizing archaea. Pro-ceedings of the National Academy of Sciences of the USA 113, 7888–7893.

Klappenbach, J.A., Saxman, P.R., Cole, J.R., Schmidt, T.M., 2001. rrndb: the ribosomal RNA operon copy number database. Nucleic Acids Research 29, 181–184.

Kleindienst, S., Ramette, A., Amann, R., Knittel, K., 2012. Distribution and in situ abundance of sulfate-reducing bacteria in diverse marine hydrocarbon seep sediments. Environmental Microbiology 14, 2689–2710.

Klindworth, A., Pruesse, E., Schweer, T., Peplies, J., Quast, C., Horn, M., Glöckner, F.O., 2013. Evaluation of general 16S ribosomal RNA gene PCR primers for classical and next-gener-ation sequencing-based diversity studies. Nucleic Acids Research 41, 1–11. doi:10.1093/nar/gks808.

Knap, A., A. Michaels, A. Close, H.D., Dickson, A. (Eds.), 1996. Protocols for the Joint Global Ocean Flux Study (JGOFS) Core Measurements. JGOFS Report 19, 1–170.

Knief, C., Altendorf, K., Lipski, A., 2003. Linking autotrophic activity in environmental samples with specific bacterial taxa by detection of 13C-labelled fatty acids. Environmental Microbiology 5, 1155–1167.

Koga, Y., Kyuragi, T., Nishihara, M., Sone, N., 1998. Did archaeal and bacterial cells arise inde-pendently from noncellular precursors? A hypothesis stating that the advent of membrane phospholipid with enantiomeric glycerophosphate backbones caused the separation of the two lines of descent. Journal of Molecular Evolution 46, 54–63.

Koga, Y., Morii, H., 2007. Biosynthesis of ether-type polar lipids in archaea and evolutionary considerations. Microbiology and Molecular Biology Reviews 71, 97–120.

Koga, Y., Nakano, M., 2008. A dendrogram of archaea based on lipid component parts composition and its relationship to rRNA phylogeny. Systematic and Applied Microbiology 31, 169–182.

Könneke, M., Bernhard, A.E., de la Torre, J.R., Walker, C.B., Waterbury, J.B., Stahl, D.A., 2005. Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437, 543–546.

Könneke, M., Lipp, J.S., Hinrichs, K.-U., 2012. Carbon isotope fractionation by the marine ammonia-oxidizing archaeon Nitrosopumilus maritimus. Organic Geochemistry 21–24.

Kowalchuk, G.A., Stephen, J.R., De Boer, W., Prosser, J.I., Embley, T.M., Woldendorp, J.W., 1997 Analysis of ammonia-oxidizing bacteria of the beta subdivision of the class Proteo-bacteria in coastal sand dunes by denaturing gradient gel electrophoresis and sequencing of PCR-amplified 16S ribosomal DNA fragments. Applied and Environmental Microbiology 63, 1489–1497.

Kristensen, E., 2000. Organic matter diagenesis at the oxic/anoxic interface in coastal marine sediments, with emphasis on the role of burrowing animals. Hydrobiologia 426, 1–24.

Kubo, K., Lloyd, K.G., F Biddle, J., Amann, R., Teske, A., Knittel, K., 2012. Archaea of the Miscellaneous Crenarchaeotal Group are abundant, diverse and widespread in marine sedi-ments. The ISME Journal 6, 1949–1965.

Kuenen, J.G., 2008. Anammox bacteria: from discovery to application. Nature Reviews Microbiol-ogy 6, 320–326.

Page 214: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

206

Kuypers, M.M.M., Sliekers, A.O., Lavik, G., Schmid, M., Jørgensen, B.B., Kuenen, J.G., Sin-ninghe Damsté, J.S., Strous, M., Jetten, M.S.M., 2003. Anaerobic ammonium oxidation by anammox bacteria in the Black Sea. Letters to Nature 422, 2–5.

Kuypers, M.M.M., Lavik, G., Woebken, D., Schmid, M., Fuchs, B.M., Amann, R., Jørgensen, B.B., Jetten, M.S.M., 2005. Massive nitrogen loss from the Benguela upwelling system through anaerobic ammonium oxidation. Proceedings of the National Academy of Sciences of the USA 102, 6478–6483.

Labrenz, M., Grote, J., Mammitzsch, K., Boschker, H.T.S., Laue, M., Jost, G., Glaubitz, S., Jürgens, K., 2013. Sulfurimonas gotlandica sp. nov., a chemoautotrophic and psychrotolerant epsilonproteobacterium isolated from a pelagic redoxcline, and an emended description of the genus Sulfurimonas. International Journal of Systematic and Evolutionary Microbiology 63, 4141–4148.

Lam, P., Jensen, M.M., Lavik, G., McGinnis, D.F., Müller, B., Schubert, C.J., Amann, R., Thamdrup, B., Kuypers, M.M.M., 2007. Linking crenarchaeal and bacterial nitrification to anammox in the Black Sea. Proceedings of the National Academy of Sciences of the USA 104, 7104–7109.

Lam, P., Lavik, G., Jensen, M.M., van de Vossenberg, J., Schmid, M., Woebken, D., Gutiérrez, D., Amann, R., Jetten, M.S.M., Kuypers, M.M.M., 2009. Revising the nitrogen cycle in the Peruvian oxygen minimum zone. Proceedings of the National Academy of Sciences of the USA 106, 4752–4757.

Lattuati, A., Guezennec, J., Metzger, P., Largeau, C., 1998. Lipids of Thermococcus hydrother-malis, an archaea isolated from a deep- sea hydrothermal vent. Lipids 33, 319–326.

Laverock, B., Smith, C., Tait, K., Osborn, M., Widdicombe, S., Gilbert, J., 2010. Bioturbating shrimp alter the structure and diversity of bacterial communities in coastal marine sedi-ments. The ISME Journal 4, 1531–1544.

Laverock, B., Tait, K., Gilbert, J.A., Osborn, A.M., Widdicombe, S., 2013. Impacts of biotur-bation on temporal variation in bacterial and archaeal nitrogen-cycling gene abundance in coastal sediments. Environmental Microbiology Reports 6, 113–121.

Lavik, G., Stührmann, T., Brüchert, V., Van der Plas, A., Mohrholz, V., Lam, P., Mußmann, M., Fuchs, B.M., Amann, R., Lass, U., Kuypers, M.M.M., 2009. Detoxification of sulphidic African shelf waters by blooming chemolithotrophs. Nature 457, 581–584.

Laws, E.A., 1984. Improved estimates of phytoplankton carbon based on 14C Incorporation into chlorophyll a. Journal of Theoretical Biology 110, 425–434.

Lee, J.A., Francis, C.A., 2017. Spatiotemporal characterization of San Francisco Bay denitrify-ing communities: a comparison of nirK and nirS diversity and abundance. Microbial Ecology 73, 271–284.

Lengger, S.K., Hopmans, E.C., Reichart, G.J., Nierop, K.G.J., Sinninghe Damsté, J.S., Schouten, S., 2012. Intact polar and core glycerol dibiphytanyl glycerol tetraether lipids in the Arabian Sea oxygen minimum zone. Part II: Selective preservation and degradation in sediments and consequences for the TEX86. Geochimica et Cosmochimica Acta 98, 228–243.

Lengger, S.K., Hopmans, E.C., Sinninghe Damsté, J.S., Schouten, S., 2012. Comparison of ex-traction and work up techniques for analysis of core and intact polar tetraether lipids from sedimentary environments. Organic Geochemistry 47, 34–40.

Lenk, S., Arnds, J., Zerjatke, K., Musat, N., Amann, R., Mumann, M., 2011. Novel groups of Gammaproteobacteria catalyse sulfur oxidation and carbon fixation in a coastal, intertidal sediment. Environmental Microbiology 13, 758–774.

Page 215: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

207

Li, M., Hong, Y., Klotz, M.G., Gu, J.-D., 2010. A comparison of primer sets for detecting 16S rRNA and hydrazine oxidoreductase genes of anaerobic ammonium-oxidizing bacteria in marine sediments. Applied Microbiology and Biotechnology 86, 781–790.

Li, M., Ford, T., Li, X., Gu, J., 2011. Cytochrome cd1 -containing nitrite reductase encoding gene nirS as a new functional biomarker for detection of anaerobic ammonium oxidizing (anammox) bacteria. Environmental Science & Technology 45, 3547–3553.

Li, M., Gu, J.-D., 2011. Advances in methods for detection of anaerobic ammonium oxidizing (anammox) bacteria. Applied microbiology and biotechnology 90, 1241–1252.

Li, J., Li, F., Yu, S., Qin, S., Wang, G., 2013. Impacts of mariculture on the diversity of bacterial communities within intertidal sediments in the Northeast of China. Microbial Ecology 66, 861–870.

Li, Y., Peacock, A., White, D., Geyer, R., Zhang, C., 2007. Spatial patterns of bacterial signature biomarkers in marine sediments of the Gulf of Mexico. Chemical Geology 238, 168–179.

Lin, Y., Lipp, J.S., Elvert, M., Holler, T., Hinrichs, K., 2013. Assessing production of the ubiq-uitous archaeal diglycosyl tetraether lipids in marine subsurface sediment using intramolecu-lar stable isotope probing. Environmental Microbiology 15, 1634–1646.

Lipp, J.S., Morono, Y., Inagaki, F., Hinrichs, K.-U., 2008. Significant contribution of Archaea to extant biomass in marine subsurface sediments. Nature 454, 991–994.

Lipp, J.S., Hinrichs, K.-U., 2009. Structural diversity and fate of intact polar lipids in marine sediments. Geochimica et Cosmochimica Acta 73, 6816–6833.

Lipski, A., Spieck, E., Makolla, A., Altendorf, K., 2001. Fatty acid profiles of nitrite-oxidiz-ing bacteria reflect their phylogenetic heterogeneity. Systematic and Applied Microbiology 24, 377–384.

Liu, X., Lipp, J.S., Hinrichs, K.-U., 2011. Distribution of intact and core GDGTs in marine sediments. Organic Geochemistry 42, 368–375.

Llobet-Brossa, E., Rabus, R., Böttcher, M.E., Könneke, M., Finke, N., Schramm, A., Meyer, R.L., Grötzschel, S., Rosselló-Mora, R., Amann, R., 2002. Community structure and activity of sulfate-reducing bacteria in an intertidal surface sediment : a multi-method approach. Aquatic Microbial Ecology 29, 211–226.

Lloyd, K.G., May, M.K., Kevorkian, R.T., Steen, D., 2013. Meta-Analysis of quantification methods shows that Archaea and Bacteria have similar abundances in the subseafloor. Ap-plied and Environmental Microbiology 79, 7790–7799.

Logemann, J., Graue, J., Köster, J., Engelen, B., Rullkötter, J., Cypionka, H., 2011. A laboratory experiment of intact polar lipid degradation in sandy sediments. Biogeosciences 8, 2547–2560. doi:10.5194/bg-8-2547-2011.

Lohse, L., Epping, E., Helder, W., van Raaphorst, W., 1996. Oxygen pore water profiles in continental shelf sediments of the North Sea: turbulent versus molecular diffusion. Marine Ecology Progress Series 145, 63–75.

Lopes dos Santos, R.A., Prange, M., Castaneda, I.S., Schefuß, E., Mulitza, S., Schulz, M., Nie-dermeyer, E.M., Sinninghe Damsté, J.S., Schouten, S., 2010. Glacial-interglacial variability in Atlantic meridional overturning circulation and thermocline adjustments in the tropical North Atlantic. Earth and Planetary Science Letters 300, 407–414.

Ludwig, W., Strunk, O., Westram, R., Richter, L., Meier, H., Yadhukumar, Buchner, A., Lai, T., Steppi, S., Jobb, G., Förster, W., Brettske, I., Gerber, S., Ginhart, A.W., Gross, O., Grumann, S., Hermann, S., Jost, R., König, A., Liss, T., Lüssmann, R., May, M., Nonhoff, B., Reichel, B., Strehlow, R., Stamatakis, A., Stuckmann, N., Vilbig, A., Lenke, M., Ludwig, T., Bode, A., Schleifer, K.-H., 2004. ARB: a software environment for sequence data. Nucleic Acids Research 32, 1363–1371.

Page 216: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

208

Ma, Y., Liu, F., Kong, Z., Yin, J., Kou, W., Wu, L., Ge, G., 2016. The distribution pattern of sediment archaea community of the Poyang Lake, the largest freshwater lake in China. Archaea. Article ID 9278929. doi:10.1155/2016/9278929.

Madsen, E.L., 2008. Microbial Diversity: Who is here and how do we know? 162–224 in: Envi-ronmental Microbiology: From Genomes to Biogeochemistry. Wiley-Blackwell, UK.

Malkin, S.Y., Rao, A.M., Seitaj, D., Vasquez-Cardenas, D., Zetsche, E.-M., Hidalgo-Martinez, S., Boschker, H.T., Meysman, F.J., 2014. Natural occurrence of microbial sulphur oxidation by long-range electron transport in the seafloor. The ISME Journal 8, 1843–1854.

Martens-Habbena, W., Berube, P.M., Urakawa, H., de la Torre, J.R., Stahl, D.A., 2009. Ammo-nia oxidation kinetics determine niche separation of nitrifying Archaea and Bacteria. Nature 461, 976–979.

Marzocchi, U., Trojan, D., Larsen, S., Louise Meyer, R., Peter Revsbech, N., Schramm, A., Peter Nielsen, L., Risgaard-Petersen, N., 2014. Electric coupling between distant nitrate reduction and sulfide oxidation in marine sediment. The ISME Journal 8, 1682–1690.

McDonald, D., Price, M.N., Goodrich, J., Nawrocki, E.P., DeSantis, T.Z., Probst, A., Andersen, G.L., Knight, R., Hugenholtz, P., 2012. An improved Greengenes taxonomy with explicit ranks for ecological and evolutionary analyses of Bacteria and Archaea. The ISME Journal 6, 610–618.

Meng, J., Xu, J., Qin, D., He, Y., Xiao, X., Wang, F., 2014. Genetic and functional properties of uncultivated MCG archaea assessed by metagenome and gene expression analyses. The ISME Journal 8, 650–659.

Meyer, B., Kuever, J., 2007a. Phylogeny of the alpha and beta subunits of the dissimilatory ad-enosine-5’-phosphosulfate (APS) reductase from sulfate-reducing microorganisms - origin and evolution of the dissimilatory sulfate-reduction pathway. Microbiology 153, 2026–2044.

Meyer, B., Kuever, J., 2007b. Molecular analysis of the distribution and phylogeny of dissimila-tory adenosine-5’-phosphosulfate reductase-encoding genes (aprBA) among sulfur- oxidiz-ing prokaryotes. Microbiology 153, 3478–3498.

Meyer, B., Kuever, J., 2007c. Molecular analysis of the diversity of sulfate-reducing and sul-fur-oxidizing prokaryotes in the environment, using aprA as functional marker gene. Applied and Environmental Microbiology 73, 7664–7679.

Meyer, R.L., Risgaard-Petersen, N., Allen, D.E., 2005. Correlation between anammox activity and microscale distribution of nitrite in a subtropical mangrove sediment. Applied and Envi-ronmental Microbiology 71, 6142–6149.

Meysman, F.J.R., Middelburg, J.J., Heip, C.H.R., 2006. Bioturbation: a fresh look at Darwin’s last idea. Trends in Ecology & Evolution 21, 688–695.

Meysman, F.J.R., Risgaard-Petersen, N., Malkin, S.Y., Nielsen, L.P., 2015. The geochemical fingerprint of microbial long-distance electron transport in the seafloor. Geochimica et Cosmo-chimica Acta 152, 122–142.

Middelburg, J.J., Barranguet, C., Boschker, H.T.S., Herman, P.M.J., Moens, T., Heip, C.H.R., 2000. The fate of intertidal microphytobenthos carbon: An in situ 13C-labeling study. Limnol-ogy and Oceanography 45, 1224–1234.

Middelburg, J.J., 2011. Chemoautotrophy in the ocean. Geophysical Research Letters 38, 94–97.Middelburg, J.J., Levin, L.A., 2009. Coastal hypoxia and sediment biogeochemistry. Biogeosciences

6, 3655–3706. doi:10.5194/bgd-6-3655-2009. Middelburg, J.J., 2014. Stable isotopes dissect aquatic food webs from the top to the bottom.

Biogeosciences 11, 2357–2371. doi:10.5194/bg-11-2357-2014.

Page 217: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

209

Miyatake, T., 2011. Linking microbial community structure to biogeochemical function in coastal marine sediments. PhD dissertation. University of Amsterdam, Amsterdam, The Netherlands.

Mohan, S.B., Schmid, M., Jetten, M., Cole, J., 2004. Detection and widespread distribution of the nrfA gene encoding nitrite reduction to ammonia, a short circuit in the biological nitro-gen cycle that competes with denitrification. FEMS Microbiology Ecology 49, 433–443.

Moodley, L., Middelburg, J.J., Boschker, H.T.S., Duineveld, G.C.A., Pel, R., Herman, P.M., Heip, C.H.R., 2017. Bacteria and Foraminifera: key players in a short-term deep-sea benthic response to phytodetritus. Marine Ecology Progress Series 236, 23–29.

Moore, E.K., Villanueva, L., Hopmans, E.C., Rijpstra, W.I.C., Mets, A., Dedish, S.N., Sinninghe Damsté, J.S., 2015. Abundant trimethylornithine lipids and specific gene sequences indicate Planctomycete importance at the oxic/anoxic interface in Sphagnum -dominated northern wetlands. Applied and Environmental Microbiology 81, 6333–6344.

Morono, Y., Terada, T., Nishizawa, M., Ito, M., Hillion, F., Takahata, N., Sano, Y., Inagaki, F., 2011. Carbon and nitrogen assimilation in deep subseafloor microbial cells. Proceedings of the National Academy of Sciences of the USA 108, 18295–18300.

Mosier, A., Francis, C., 2008. Relative abundance and diversity of ammonia-oxidizing archaea and bacteria in the San Francisco Bay estuary. Environmental Microbiology 10, 3002–3016.

Mosier, A.C., Francis, C.A., 2010. Denitrifier abundance and activity across the San Francisco Bay estuary. Environmental Microbiology Reports 2, 667–76.

Mosier, A.C., Lund, M.B., Francis, C.A., 2012. Ecophysiology of an ammonia-oxidizing archae-on adapted to low-salinity habitats. Microbial Ecology 64, 955–963.

Murphy, J., Riley, J.P., 1962. A modified single solution method for the determination of phos-phate in natural waters. Analytica Chimica Acta 27, 31–36.

Musat, N., Halm, H., Winterholler, B., Hoppe, P., Peduzzi, S., Hillion, F., Horreard, F., Amann, R., Jørgensen, B.B., Kuypers, M.M.M., 2008. A single-cell view on the ecophysiology of anaerobic phototrophic bacteria. Proceedings of the National Academy of Sciences of the USA 105, 17861–17866.

Mußmann, M., Schulz, H.N., Strotmann, B., Kjaer, T., Nielsen, L.P., Rosselló-Mora, R.A., Amann, R.I., Jørgensen, B.B., 2003. Phylogeny and distribution of nitrate-storing Beggiatoa spp. in coastal marine sediments. Environmental Microbiology 5, 523–533.

Mußmann, M., Hu, F.Z., Richter, M., De Beer, D., Preisler, A., Jørgensen, B.B., Huntemann, M., Glöckner, F.O., Amann, R., Koopman, W.J.H., Lasken, R.S., Janto, B., Hogg, J., Stoodley, P., Boissy, R., Ehrlich, G.D., 2007. Insights into the genome of large sulfur bacteria revealed by analysis of single filaments. PLoS Biology 5, e230. doi:10.1371/journal.pbio.0050230.

Mußmann, M., Brito, I., Pitcher, A., Sinninghe Damsté, J.S., Hatzenpichler, R., Richter, A., Nielsen, J.L., Nielsen, P.H., Müller, A., Daims, H., Wagner, M., Head, I.M., 2011. Thau-marchaeotes abundant in refinery nitrifying sludges express amoA but are not obligate autotrophic ammonia oxidizers. Proceedings of the National Academy of Sciences of the USA 108, 16771–16776.

Muyzer, G., Stams, A.J.M., 2008. The ecology and biotechnology of sulphate-reducing bacteria. Nature Reviews Microbiology 6, 441–454.

Nanba, K., King, G.M., Dunfield, K., 2004. Analysis of facultative lithotroph distribution and diversity on volcanic deposits by use of the large subunit of ribulose 1,5-bisphosphate car-boxylase/oxygenase. Applied and Environmental Microbiology 70, 2245–2253.

Nelson, D.C., Jannasch, H.W., 1983. Chemoautotrophic growth of a marine Beggiatoa in sul-fide-gradient cultures. Archives of Microbiology 136, 262–269.

Page 218: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

210

Neubacher, E.C., Parker, R.E., Trimmer, M., 2011. Short-term hypoxia alters the balance of nitrogen cycle in coastal sediments. Limnology and Oceanography 56, 651–665.

Neubacher, E.C., Parker, R.E., Trimmer, M., 2013. The potential effect of sustained hypoxia on nitrogen cycling in sediment from the southern North Sea: a mesocosm experiment. Biogeochemistry 113, 69–84.

Neufeld, J.D., Dumont, M.G., Vohra, J., Murrell, J.C., 2006. Methodological considerations for the use of stable isotope probing in microbial ecology. Microbial Ecology 53, 435–442.

Nicholls, J., and Trimmer, M., 2009. Widespread occurrence of the anammox reaction in estua-rine sediments. Aquatic Microbial Ecology 55, 105–113.

Nicol, G.W., Leininger, S., Schleper, C., Prosser, J.I., 2008. The influence of soil pH on the diversity, abundance and transcriptional activity of ammonia oxidizing archaea and bacteria. Environmental Microbiology 10, 2966–2978.

Nielsen, L.P., Christensen, P.B., Revsbech, N.P., Sørensen, J., 1990. Denitrification and photo-synthesis in stream sediment studied with microsensor and whole core techniques. Limnology and Oceanography 35, 1135–1144.

Nielsen, L.P., 1992. Denitrification in sediment determined from nitrogen isotope pairing. FEMS Microbiology Letters 86, 357–362.

Nielsen, L.P., Risgaard-Petersen, N., Fossing, H., Christensen, P.B., Sayama, M., 2010. Electric currents couple spatially separated biogeochemical processes in marine sediment. Nature 463, 1071–1074.

Nienhuis, P.H., De Bree, B.H.H., 1984. Carbon fixation and chlorophyll in bottom sediments of brackish Lake Grevelingen, The Netherlands. Netherlands Journal of Sea Research 18, 337–359.

Nieuwenhuize, J., Maas, Y.E.M., Middelburg, J.J., 1994. Rapid analysis of organic carbon and nitrogen in particulate materials. Marine Chemistry 45, 217-224.

Niftrik van, L., Geerts, W.J.C., van Donselaar, E.G., Humbel, B.M., Webb, R.I., Fuerst, J.A., Verkleij, A.J., Jetten, M.S.M., Strous, M., 2008. Linking ultrastructure and function in four genera of anaerobic ammonium-oxidizing bacteria: Cell plan, glycogen storage, and local-ization of cytochrome c proteins. Journal of Bacteriology 190, 708–717.

Nigro, L.M., King, G.M., 2007. Disparate distributions of chemolithotrophs containing form IA or IC large subunit genes for ribulose-1,5-bisphosphate carboxylase/oxygenase in inter-tidal marine and littoral lake sediments. FEMS Microbiology Ecology 60, 113–125.

Nogales, B., Timmis, K.N., Nedwell, D.B., Osborn, A.M., 2002. Detection and diversity of ex-pressed denitrication genes in estuarine sediments after reverse transcription-PCR amplica-tion from mRNA. Applied and Environmental Microbiology 68, 5017–5025.

Nomaki, H., Ogawa, N.O., Takano, Y., Suga, H., Ohkouchi, N., Kitazato, H., 2011. Differing utilization of glucose and algal particulate organic matter by deep-sea benthic organisms of Sagami Bay, Japan. Marine Ecology Progress Series 431, 11–24.

Oakley, B.B., Francis, C.A., Roberts, K.J., Fuchsman, C.A., Srinivasan, S., Staley, J.T., 2007. Analysis of nitrite reductase (nirK and nirS) genes and cultivation reveal depauperate com-munity of denitrifying bacteria in the Black Sea suboxic zone. Environmental Microbiology 9, 118–130.

Oba, M., Sakata, S., Tsunogai, U., 2006. Polar and neutral isopranyl glycerol ether lipids as bio-markers of archaea in near-surface sediments from the Nankai Trough. Organic Geochemistry 37, 1643–1654.

Ortiz-Alvarez, R., Casamayor, E.O., 2016. High occurrence of Pacearchaeota and Woesearchaeota (Archaea superphylum DPANN) in the surface waters of oligotrophic high-altitude lakes. Environmental Microbiology Reports 8, 210–217.

Page 219: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

211

Ouverney, C.C., Fuhrman, J.A., 2000. Marine planktonic archaea take up amino acids. Applied and Environmental Microbiology 66, 4829–4833.

Pagani, I., Lapidus, A., Nolan, M., Lucas, S., Hammon, N., Deshpande, S., Cheng, J.-F., Chert-kov, O., Davenport, K., Tapia, R., Han, C., Goodwin, L., Pitluck, S., Liolios, K., Mavroma-tis, K., Ivanova, N., Mikhailova, N., Pati, A., Chen, A., Palaniappan, K., Land, M., Hauser, L., Chang, Y.-J., Jeffries, C.D., Detter, J.C., Brambilla, E., Kannan, K.P., Djao, O.D.N., Rohde, M., Pukall, R., Spring, S., Göker, M., Sikorski, J., Woyke, T., Bristow, J., Eisen, J.A., Markowitz, V., Hugenholtz, P., Kyrpides, N.C., Klenk, H.-P., 2011. Complete genome se-quence of Desulfobulbus propionicus type strain (1pr3). Standards in Genomic Sciences 4, 100–110.

Pancost, R.D., Hopmans, E.C., Sinninghe Damsté, J.S., 2001. Archaeal lipids in mediterranean cold seeps: Molecular proxies for anaerobic methane oxidation. Geochimica et Cosmochimica Acta 65, 1611–1627.

Park, B.-J., Park, S.-J., Yoon, D.-N., Schouten, S., Sinninghe Damsté, J.S., Rhee, S.-K., 2010. Cultivation of autotrophic ammonia-oxidizing archaea from marine sediments in coculture with sulfur-oxidizing bacteria. Applied and Environmental Microbiology 76, 7575–7587.

Pearson, A., McNichol, A.P., Benitez-Nelson, B.C., Hayes, J.M., Eglington, T.I., 2001. Origins of lipid biomarkers in Santa Monica Basin surface sediment: A case study using com-pound-specific 14C analysis. Geochimica et Cosmochimica Acta 65, 3123–3137.

Pearson, A., Huang, Z., Ingalls, a. E., Romanek, C.S., Wiegel, J., Freeman, K.H., Smittenberg, R.H., Zhang, C.L., 2004. Nonmarine crenarchaeol in Nevada hot springs. Applied and Envi-ronmental Microbiology 70, 5229–5237.

Pearson, A., Ingalls, A.E., 2013. Assessing the use of archaeal lipids as marine environmental proxies. Annual Review of Earth and Planetary Sciences 41, 359–384.

Pernthaler, A., Pernthaler, J., Amann, R., 2002. Fluorescence in situ hybridization and catalyzed reporter deposition for the identification of marine bacteria. Applied and Environmental Micro-biology 68, 3094–3101.

Pester, M., Schleper, C., Wagner, M., 2011. The Thaumarchaeota : an emerging view of their phylogeny and ecophysiology. Current Opinion in Microbiology 14, 300–306.

Pester, M., Rattei, T., Flechl, S., Gröngröft, A., Richter, A., Overmann, J., Reinhold-Hurek, B., Loy, A., Wagner, M., 2012. amoA-based consensus phylogeny of ammonia-oxidizing archaea and deep sequencing of amoA genes from soils of four different geographic regions. Envi-ronmental Microbiology 14, 525–539.

Pfeffer, C., Larsen, S., Song, J., Dong, M., Besenbacher, F., Meyer, R.L., Kjeldsen, K.U., Sch-reiber, L., Gorby, Y.A., El-Naggar, M.Y., Leung, K.M., Schramm, A., Risgaard-Petersen, N., Nielsen, L.P., 2012. Filamentous bacteria transport electrons over centimetre distances. Nature 491, 218–221.

Pitcher, A., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2009. Separation of core and intact polar archaeal tetraether lipids using silica columns: Insights into living and fossil biomass contributions. Organic Geochemistry 40, 12–19.

Pitcher, A., Hopmans, E.C., Mosier, A.C., Park, S.-J., Rhee, S.-K., Francis, C.A., Schouten, S., Sinninghe Damsté, J.S., 2011a. Core and intact polar glycerol dibiphytanyl glycerol tetra-ether lipids of ammonia-oxidizing archaea enriched from marine and estuarine sediments. Applied and Environmental Microbiology 77, 3468–3477.

Pitcher, A., Villanueva, L., Hopmans, E.C., Schouten, S., Reichart, G.-J., Sinninghe Damsté, J.S., 2011b. Niche segregation of ammonia-oxidizing archaea and anammox bacteria in the Arabian Sea oxygen minimum zone. The ISME Journal 5, 1896–1904.

Page 220: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

212

Pitcher, A., Wuchter, C., Siedenberg, K., Schouten, S., Sinninghe Damsté, J.S., 2011c. Crenar-chaeol tracks winter blooms of ammonia-oxidizing Thaumarchaeota in the coastal North Sea. Limnology and Oceanography 5, 2308–2318.

Pjevac, P., Kamyshny, A., Dyksma, S., Mußmann, M., 2014. Microbial consumption of zero-va-lence sulfur in marine benthic habitats. Environmental Microbiology 16, 3416–3430.

Plumb, J.J., Haddad, C.M., Gibson, J.A.E., Franzmann, P.D., 2007. Acidianus sulfidivorans sp. nov., an extremely acidophilic, thermophilic archaeon isolated from a solfatara on Lihir Island, Papua New Guinea, and emendation of the genus description. International Journal of System-atic and Evolutionary Microbiology 57, 1418–1423.

Porubsky, W.P., Weston, N.B., Joye, S.B., 2009. Benthic metabolism and the fate of dissolved inorganic nitrogen in intertidal sediments. Estuarine, Coastal and Shelf Science 83, 392–402.

Priemé, A., Braker, G., Tiedje, J.M., 2002. Diversity of Nitrite Reductase (nirK and nirS) gene fragments in forested upland and wetland soils. Applied and Environmental Microbiology 68, 1893–1900.

Prokopenko, M.G., Sigman, D.M., Berelson, W.M., Hammond, D.E., Barnett, B., Chong, L., Townsend-Small, A., 2011. Denitrification in anoxic sediments supported by biological nitrate transport. Geochimica et Cosmochimica Acta 75, 7180–7199.

Prokopenko, M.G., Hirst, M.B., De Brabandere, L., Lawrence, D.J.P., Berelson, W.M., Granger, J., Chang, B.X., Dawson, S., Crane, E.J., Chong, L., Thamdrup, B., Townsend-Small, A., Sigman, D.M., 2013. Nitrogen losses in anoxic marine sediments driven by Thioploca-ana-mmox bacterial consortia. Nature 500, 194–198.

Popp, B.N., Bidigare, R.R., Deschenes, B., Laws, E.A., Prahl, F.G., Tanimoto, J.K., 2006. A new method for estimating growth rates of alkenone-producing haptophytes. Limnology and Oceanography: Methods 4, 114–129.

Prosser, J.I., Nicol, G.W., 2008. Relative contributions of archaea and bacteria to aerobic am-monia oxidation in the environment. Environmental Microbiology 10, 2931–2941.

Quast, C., Klindworth, A., Pruesse, E., Schweer, T., Horn, M., Glo, F.O., 2013. Evaluation of general 16S ribosomal RNA gene PCR primers for classical and next-generation sequenc-ing-based diversity studies. Nucleic Acids Research 41, D590–D596.

Raaphorst, W. van, Kloosterhuis, H.T., Berghuis, E.M., Gieles, A.J.M., van Noort, G.J., 1992. Nitrogen cycling in two types of sediments of the southern North Sea (Frisian front, Broad Fourteens): field data and mesocosm results. Netherlands Journal of Sea Research 28, 293–316.

Raaphorst, W. van, Malschaert, H., van Haren, H. 1998. Tidal resuspension and deposition of particulate matter in the Oyster Grounds, North Sea. Marine Research 56, 257–291.

Radajewski, S., Ineson, P., Parekh, N.R., Murrell, J.C., 2000. Stable-isotope probing as a tool in microbial ecology. Nature 403, 646–649.

Radajewski, S., McDonald, I.R., Murrel, J.C., 2003. Stable-isotope probing of nucleic acids: a window to the function of uncultured microorganisms. Current Opinion in Biotechnology 14, 296–302.

Radajewski, S., McDonald, I.R., Murrell, J.C., 2003. Stable-isotope probing of nucleic acids: A window to the function of uncultured microorganisms. Current Opinion in Biotechnology 14, 296–302.

Rao, A.M.F., Malkin, S.Y., Hidalgo-Martinez, S., Meysman, F.J.R., 2016. The impact of electro-genic sulfide oxidation on elemental cycling and solute fluxes in coastal sediment. Geochimica et Cosmochimica Acta 172, 265–286.

Rappé, M.S., Giovannoni, S.J., 2003. The uncultured microbial majority. Annual Review of Micro-biology 57, 369–394.

Page 221: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

213

Rattray, J.E., van De Vossenberg, J., Hopmans, E.C., Kartal, B., van Niftrik, L., Rijpstra, W.I.C., Strous, M., Jetten, M.S.M., Schouten, S., Sinninghe Damsté, J.S., 2008. Ladderane lipid dis-tribution in four genera of anammox bacteria. Archives of Microbiology 190, 51–66.

Revsbech, N.P., 1989. An oxygen microsensor with a guard cathode. Limnology and Oceanography 34, 474–478.

Rinke, C., Schwientek, P., Sczyrba, A., Ivanova, N.N., Anderson, I.J., Cheng, J.-F., Darling, A., Malfatti, S., Swan, B.K., Gies, E.A., Dodsworth, J.A., Hedlund, B.P., Tsiamis, G., Sievert, S.M., Liu, W.-T., Eisen, J.A., Hallam, S.J., Kyrpides, N.C., Stepanauskas, R., Rubin, E.M., Hugenholtz, P., Woyke, T., 2013. Insights into the phylogeny and coding potential of micro-bial dark matter. Nature 499, 431–437.

Risgaard-Petersen, N., 2003. Coupled nitrification-denitrification in autotrophic and heterotro-phic estuarine sediments: On the influence of benthic microalgae. Limnology and Oceanography 48, 93–105.

Risgaard-Petersen, N., Meyer, R.L., Schmid, M., Jetten, M.S.M., Enrich-Prast, A., Rysgaard, S., Revsbech, N.P., 2004. Anaerobic ammonium oxidation in an estuarine sediment. Aquatic Microbial Ecology 36, 293–304.

Risgaard-Petersen, N., Meyer, R.L., Revsbech, N.P., 2005. Denitrification and anaerobic am-monium oxidation in sediments: effects of microphytobenthos and NO3

-. Aquatic Microbial Ecology 40, 67–76.

Risgaard-Petersen, N., Langezaal, A.M., Ingvardsen, S., Schmid, M.C., Jetten, M.S.M., Op den Camp, H.J.M., Derksen, J.W.M., Piña-Ochoa, E., Eriksson, S.P., Peter Nielsen, L., Peter Revsbech, N., Cedhagen, T., van der Zwaan, G.J., 2006. Evidence for complete denitrifica-tion in a benthic foraminifer. Nature 443, 93–96.

Risgaard-Petersen, N., Damgaard, L.R., Revil, A., Nielsen, L.P., 2014. Mapping electron sources and sinks in a marine biogeobattery. Journal of Geophysical Research Biogeosciences 119, 1475-1486. doi:10.1002/2014JG002673.

Rossel, P.E., Lipp, J.S., Fredricks, H.F., Arnds, J., Boetius, A., Elvert, M., Hinrichs, K.U., 2008. Intact polar lipids of anaerobic methanotrophic archaea and associated bacteria. Organic Geochemistry 39, 992–999.

Rooks, C., Schmid, M.C., Mehsana, W., Trimmer, M., 2012. The depth-specific significance and relative abundance of anaerobic ammonium-oxidizing bacteria in estuarine sediments (Medway Estuary, UK). FEMS Microbiology Ecology 80, 19–29.

Rotthauwe, J.H., Witzel, K.P., Liesack, W., 1997. The ammonia monooxygenase structural gene amoA as a functional marker: molecular fine-scale analysis of natural ammonia-oxidizing populations. Applied and environmental microbiology 63, 4704–12.

Ruehland, C., Blazejak, A., Lott, C., Loy, A., Erséus, C., Dubilier, N., 2008. Multiple bacterial symbionts in two species of co-occurring gutless oligochaete worms from Mediterranean sea grass sediments. Environmental Microbiology 10, 3404–3416.

Rusch, D.B., Halpern, A.L., Sutton, G., Heidelberg, K.B., Williamson, S., Yooseph, S., Wu, D., Eisen, J. a, Hoffman, J.M., Remington, K., Beeson, K., Tran, B., Smith, H., Baden-Tillson, H., Stewart, C., Thorpe, J., Freeman, J., Andrews-Pfannkoch, C., Venter, J.E., Li, K., Krav-itz, S., Heidelberg, J.F., Utterback, T., Rogers, Y.-H., Falcón, L.I., Souza, V., Bonilla-Rosso, G., Eguiarte, L.E., Karl, D.M., Sathyendranath, S., Platt, T., Bermingham, E., Gallardo, V., Tamayo-Castillo, G., Ferrari, M.R., Strausberg, R.L., Nealson, K., Friedman, R., Frazier, M., Venter, J.C., 2007. The Sorcerer II Global Ocean Sampling expedition: northwest Atlantic through eastern tropical Pacific. PLoS Biology 5, e77. doi:10.1371/journal.pbio.0050077.

Rush, D., Jaeschke, A., Hopmans, E.C., Geenevasen, J.A.J., Schouten, S., Sinninghe Damsté, J.S., 2011. Short chain ladderanes: Oxic biodegradation products of anammox lipids. Geo-chimica et Cosmochimica Acta 75, 1662–1671.

Page 222: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

214

Rush, D., Wakeham, S.G., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2012a. Bio-marker evidence for anammox in the oxygen minimum zone of the Eastern Tropical North Pacific. Organic Geochemistry 53, 80–87.

Rush, D., Hopmans, E.C., Wakeham, S.G., Schouten, S., Sinninghe Damsté, J.S., 2012b. Oc-currence and distribution of ladderane oxidation products in different oceanic regimes. Biogeosciences 9, 2407–2418. doi:10.5194/bg-9-2407-2012.

Russ, L., Kartal, B., Op den Camp, H.J.M., Sollai, M., Le Bruchec, J., Caprais, J.-C., Godfroy, A., Sinninghe Damsté, J.S., Jetten, M.S.M., 2013. Presence and diversity of anammox bacteria in cold hydrocarbon-rich seeps and hydrothermal vent sediments of the Guaymas Basin. Frontiers in Microbiology 4, 219. doi:10.3389/fmicb.2013.00219.

Russ, L., Speth, D.R., Jetten, M.S.M., Op den Camp, H.J.M., Kartal, B., 2014. Interactions between anaerobic ammonium and sulfur-oxidizing bacteria in a laboratory scale model system. Environmental Microbiology 16, 3487–3498.

Rütters, H., Sass, H., Cypionka, H., Rullkötter, J., 2002. Phospholipid analysis as a tool to study complex microbial communities in marine sediments. Journal of Microbiological Methods 48, 149–160.

Rysgaard, S., Glud, R.N., Risgaard-Petersen, N., Dalsgaard, T., 2004. Denitrification and anam-mox activity in Arctic marine sediments. Limnology and Oceanography 49, 1493–1502.

Rysgaard, S., Risgaard-Petersen, N., Nielsen, L.P., Revsbech, N.P., 1993. Nitrification and denitrification in lake and estuarine sediments measured by the 15N dilution technique and isotope pairing. Applied and Environmental Microbiology 59, 2093–2098.

Sahan, E., Muyzer, G., 2008. Diversity and spatio-temporal distribution of ammonia-oxidizing Archaea and Bacteria in sediments of the Westerschelde estuary. FEMS Microbiology Ecology 64, 175–186.

Saitou, N., Nei, M., 1987. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Molecular Biology and Evolution 4, 406–425.

Sakami, T., 2012. Distribution of ammonia-oxidizing archaea and bacteria in the surface sedi-ments of Matsushima Bay in relation to environmental variables. Microbes and Environments 27, 61–66.

Salman, V., Amann, R., Girnth, A.C., Polerecky, L., Bailey, J.V., Høgslund, S., Jessen, G., Pan-toja, S., Schulz-Vogt, H.N., 2011. A single-cell sequencing approach to the classification of large, vacuolated sulfur bacteria. Systematic and Applied Microbiology 34, 243–259.

Salman, V., Bailey, J.V., Teske, A., 2013. Phylogenetic and morphologic complexity of giant sul-phur bacteria. Antonie van Leeuwenhoek, International Journal of General and Molecular Microbiology 104, 169–186.

Santoro, A.E., Francis, C.A., de Sieyes, N.R., Boehm, A.B., 2008. Shifts in the relative abun-dance of ammonia-oxidizing bacteria and archaea across physicochemical gradients in a subterranean estuary. Environmental Microbiology 10, 1068–1079.

Sayama, M., Risgaard-petersen, N., Nielsen, L.P., Fossing, H., Christensen, P.B., 2005. Impact of bacterial NO3

− transport on sediment biogeochemistry. Applied and Environmental Microbi-ology 71, 7575–7577.

Sayavedra-Soto A., Arp D.J. (2011). Ammonia-oxidizing bacteria: their biochemistry and mo-lecular biology. Nitrification, 11–37, In: Ward B., Arp D., Klotz M (ed), American Society Microbiology. ASM Press: Washington, DC, USA.

Schauer, R., Risgaard-Petersen, N., Kjeldsen, K.U., Tataru Bjerg, J.J., Jørgensen, B.B., Schramm, A., Nielsen, L.P., 2014. Succession of cable bacteria and electric currents in marine sedi-ment. The ISME Journal 8, 1314–22.

Page 223: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

215

Schmid, M.C., Maas, B., Dapena, A., Pas-Schoonen, K. van de, Vossenberg, J. Van De, Kartal, B., Niftrik, L. van, Schmidt, I., Cirpus, I., Kuenen, J.G., Wagner, M., Sinninghe Damsté, J.S., Kuypers, M., Revsbech, N.P., Mendez, R., Jetten, M.S.M., Strous, M., 2005. Biomarkers for in situ detection of anaerobic ammonium-oxidizing (Anammox) bacteria. Applied and Envi-ronmental Microbiology 71, 1677–1684.

Schleper, C., Nicol, G.W., 2010. Ammonia-oxidizing archaea-physiology, ecology and evolution, Advances in Microbial Physiology 57. doi:10.1016/B978-0-12-381045-8.00001-1.

Schloss, P.D., Gevers, D., Westcott, S.L., 2011. Reducing the effects of PCR amplification and sequencing artifacts on 16S rRNA-based studies. PLoS ONE 6, e27310. doi:10.1371/jour-nal.pone.0027310.

Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann, M., Hollister, E.B., Lesniewski, R.A., Oakley, B.B., Parks, D.H., Robinson, C.J., Sahl, J.W., Stres, B., Thallinger, G.G., Van Horn, D.J., Weber, C.F., 2009. Introducing mothur: Open-source, platform-independent, community-supported software for describing and comparing microbial communities. Ap-plied and Environmental Microbiology 75, 7537–7541.

Schouten, S., Hoefs, M.J.L., Koopmans, M.P., Bosch, H., Sinninghe Damsté, J.S., 1998. Struc-tural characterization, occurrence and fate of archaeal ether- bound acyclic and cyclic biphytanes and corresponding diols in sediments. Organic Geochemistry 29, 1305–1319.

Schouten, S., Hopmans, E.C., Pancost, R.D., Sinninghe Damsté, J.S., 2000. Widespread occur-rence of structurally diverse tetraether membrane lipids: evidence for the ubiquitous pres-ence of low-temperature relatives of hyperthermophiles. Proceedings of the National Academy of Sciences of the USA 97, 14421–14426.

Schouten, S., Hopmans, E.C., Schefuß, E., Sinninghe Damsté, J.S., 2002. Distributional varia-tions in marine crenarchaeotal membrane lipids: A new tool for reconstructing ancient sea water temperatures? Earth and Planetary Science Letters 204, 265–274.

Schouten, S., Huguet, C., Hopmans, E.C., Kienhuis, M.V.M., Sinninghe Damsté, J.S., 2007. Analytical Methodology for TEX86 Paleothermometry by High-Performance Liquid Chro-matography/Atmospheric Pressure Chemical Ionization-Mass Spectrometry. Analytical Chemistry 79, 2940–2944.

Schouten, S., Hopmans, E.C., Baas, M., Boumann, H., Standfest, S., Könneke, M., Stahl, D.A., Sinninghe Damsté, J.S., 2008. Intact membrane lipids of “Candidatus Nitrosopumilus mari-timus”, a cultivated representative of the cosmopolitan mesophilic group I crenarchaeota. Applied and Environmental Microbiology 74, 2433–2440.

Schouten, S., Middelburg, J.J., Hopmans, E.C., Sinninghe Damsté, J.S., 2010. Fossilization and degradation of intact polar lipids in deep subsurface sediments: A theoretical approach. Geochimica et Cosmochimica Acta 74, 3806–3814.

Schouten, S., Pitcher, A., Hopmans, E.C., Villanueva, L., van Bleijswijk, J., Sinninghe Damsté, J.S., 2012. Intact polar and core glycerol dibiphytanyl glycerol tetraether lipids in the Arabian Sea oxygen minimum zone: I. Selective preservation and degradation in the water column and consequences for the TEX86. Geochimica et Cosmochimica Acta 98, 228–243.

Schouten, S., Hopmans, E.C., Sinninghe Damsté, J.S., 2013. The organic geochemistry of glyc-erol dialkyl glycerol tetraether lipids : A review. Organic Geochemistry 54, 19–61.

Schramm, A., Larsen, L.H., Revsbech, N.P., Ramsing, N.B., Amann, R., Schleifer, K.H., 1996. Structure and function of a nitrifying biofilm as determined by in situ hybridization and the use of microelectrodes. Applied and Environmental Microbiology 62, 4641–4647.

Schrenk, M.O., Huber, J.A., Edwards, K.J., 2010. Microbial Provinces in the Subseafloor. Annu-al Review of Marine Science 2, 279–304.

Page 224: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

216

Schubotz, F., Lipp, J.S., Elvert, M., Hinrichs, K.U., 2011. Stable carbon isotopic compositions of intact polar lipids reveal complex carbon flow patterns among hydrocarbon degrading microbial communities at the Chapopote asphalt volcano. Geochimica et Cosmochimica Acta 75, 4399–4415.

Schulz, H.N., Jørgensen, B.B., 2001. Big Bacteria. Annual Review Micriobiology 55, 105–137.Seitaj, D., Schauer, R., Sulu-Gambari, F., Hidalgo-Martinez, S., Malkin, S.Y., Burdorf, L.D.W.,

Slomp, C.P., Meysman, F.J.R., 2015. Cable bacteria generate a firewall against euxinia in seasonally hypoxic basins. Proceedings of the National Academy of Sciences of the USA 112, 13278–13283.

Seitzinger, S.P., 1988. Denitrification in freshwater and coastal marine ecosystems: Ecological and geochemical significance. Limnology and Oceanography 33, 702–724.

Seitzinger, S., Harrison, J.A., Böhlke, J.K., Bouwman, A.F., Lowrance, R., Peterson, B., Tobias, C., Van Drecht, G., 2006. Denitrification across landscapes and waterscapes: A synthesis. Ecological Applications 16, 2064–2090.

Shah, S.R., Mollenhauer, G., Ohkouchi, N., Eglinton, T.I., Pearson, A., 2008. Origins of archaeal tetraether lipids in sediments: Insights from radiocarbon analysis. Geochimica et Cosmochimica Acta 72, 4577–4594.

Shao, M.-F., Zhang, T., Fang, H.H.-P., 2010. Sulfur-driven autotrophic denitrification: diversity, biochemistry, and engineering applications. Applied Microbiology and Biotechnology 88, 1027–1042.

Sinninghe Damsté, J.S., Schouten, S., Hopmans, E.C., van Duin, A.C.T., Geenevasen, J.A.G., 2002a. Crenarchaeol: the characteristic core glycerol dibiphytanyl glycerol tetraether mem-brane lipid of cosmopolitan pelagic crenarchaeota. The Journal of Lipid Research 43, 1641–1651.

Sinninghe Damsté, J.S., Strous, M., Rijpstra, W.I.C., Geenevasen, J.A.G., van Duin, A.C.T., van Niftrik, L.A., Jetten, M.S.M., 2002b. Linearly concatenated cyclobutane lipids form a dense bacterial membrane. Letters to Nature 419, 8–12.

Sinninghe Damsté, J.S., Ossebaar, J., Schouten, S., Verschuren, D., 2012a. Distribution of tetra-ether lipids in the 25-ka sedimentary record of Lake Challa: Extracting reliable TEX86 and MBT/CBT palaeotemperatures from an equatorial African lake. Quaternary Science Reviews 50, 43–54.

Sinninghe Damsté, J.S., Rijpstra, W.I.C., Hopmans, E.C., Jung, M.Y., Kim, J.G., Rhee, S.K., Stieglmeier, M., Schleper, C., 2012b. Intact polar and core glycerol dibiphytanyl glycerol tetraether lipids of group I.1a and I.1b Thaumarchaeota in soil. Applied and Environmental Microbiology 78, 6866–6874.

Sipos, R., Székely, A., Révész, S., Márialigeti, K., 2010. Addressing PCR biases in environmental microbiology studies, 37–58, in: Cummings, S.P. (ed), Bioremediation Methods and Proto-cols, Series Methods in Molecular Biology, 599. Humana press.

Smith, C.J., Nedwell, D.B., Dong, L.F., Osborn, A.M., 2007. Diversity and abundance of nitrate reductase genes (narG and napA), nitrite reductase genes (nirS and nrfA), and their tran-scripts in estuarine sediments. Applied and Environmental Microbiology 73, 3612–3622.

Smith, J.M., Mosier, A.C., Francis, C.A., 2015. Spatiotemporal relationships between the abun-dance, distribution, and potential activities of ammonia-oxidizing and denitrifying microor-ganisms in intertidal sediments. Microbial Ecology 69, 13–24.

Soetaert, K; Herman, P.M.J.; Middelburg, J.J., 1996. A model of diagenetic processes from the shelf to abyssal depths. Geochimica et Cosmochimica Acta 60, 1019–1040.

Page 225: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

217

Sokoll, S., Holtappels, M., Lam, P., Collins, G., Schlüter, M., Lavik, G., Kuypers, M.M.M., 2012. Benthic nitrogen loss in the arabian sea off pakistan. Frontiers in Microbiology 3, 395. doi:10.3389/ fmicb.2012.00395.

Sørensen, J., Tiedje, J.M., Firestone, R.B., 1980. Inhibition by sulfide of nitric and nitrous-ox-ide reduction by denitrifying Pseudomonas fluorescens. Applied and Environmental Microbiology 39, 105–108.

Sorokin, D.Y., Lysenko, a. M., Mityushina, L.L., Tourova, T.P., Jones, B.E., Rainey, F.A., Robert-son, L.A., Kuenen, G.J., 2001. Thioalkalimicrobium aerophilum gen. nov., sp. nov. and Thioal-kalimicrobium sibericum sp. nov., and Thioalkalivibrio versutus gen. nov., sp. nov., Thioalkalivibrio nitratis sp. nov. and Thioalkalivibrio denitrificans sp. nov., novel obligately alkaliphilic. Internation-al Journal of Systematic and Evolutionary Microbiology 51, 565–580.

Sorokin, D.Y., Kuenen, J.G., Muyzer, G., 2011. The microbial sulfur cycle at extremely haloal-kaline conditions of soda lakes. Frontiers in Microbiology 2, 44. doi:10.3389/fmicb.2011.00044.

Spang, A., Hatzenpichler, R., Brocier-Armanet, C., Rattei, T., Tischler, P., Spieck, E., Streit, W., Stahl, D.A., Wagner, M., Schleper, C., 2010. Distinct gene set in two different lineages of ammonia-oxidizing archaea supports the phylum Thaumarchaeota. Trends in Microbiology 18, 331–340.

Strous, M., Kuenen, J.G., Jetten, M.S., 1999. Key physiology of anaerobic ammonium oxida-tion. Applied and Environmental Microbiology 65, 3248–3250.

Strous, M., Pelletier, E., Mangenot, S., Rattei, T., Lehner, A., Taylor, M.W., Horn, M., Daims, H., Bartol-Mavel, D., Wincker, P., Barbe, V., Fonknechten, N., Vallenet, D., Segurens, B., Schenowitz-Truong, C., Médigue, C., Collingro, A., Snel, B., Dutilh, B.E., Op den Camp, H.J.M., van der Drift, C., Cirpus, I., van de Pas-Schoonen, K.T., Harhangi, H.R., van Nif-trik, L., Schmid, M., Keltjens, J., van de Vossenberg, J., Kartal, B., Meier, H., Frishman, D., Huynen, M.A., Mewes, H.-W., Weissenbach, J., Jetten, M.S.M., Wagner, M., Le Paslier, D., 2006. Deciphering the evolution and metabolism of an anammox bacterium from a com-munity genome. Nature 440, 790–794.

Sturt, H.F., Summons, R.E., Smith, K., Elvert, M., Hinrichs, K.-U., 2004. Intact polar mem-brane lipids in prokaryotes and sediments deciphered by high-performance liquid chroma-tography/electrospray ionization multistage mass spectrometry-new biomarkers for biogeo-chemistry and microbial ecology. Rapid Communications in Mass Spectrometry 18, 617–628.

Sulu-Gambari, F., Seitaj, D., Meysman, F.J.R., Schauer, R., Polerecky, L., Slomp, C.P., 2016. Cable bacteria control iron-phosphorus dynamics in sediments of a coastal hypoxic basin. Environmental Science and Technology 50, 1227–1233.

Suzuki, D., Ueki, A., Amaishi, A., Ueki, K., 2007. Desulfobulbus japonicus sp. nov., a novel Gram-negative propionate-oxidizing, sulfate-reducing bacterium isolated from an estuarine sediment in Japan. International Journal of Systematic and Evolutionary Microbiology 57, 849–855.

Takai, K., Suzuki, M., Nakagawa, S., Miyazaki, M., Suzuki, Y., Inagaki, F., Horikoshi, K., 2006. Sulfurimonas paralvinellae sp. nov., a novel mesophilic, hydrogen- and sulfur-oxidizing chem-olithoautotroph within the Epsilonproteobacteria isolated from a deep-sea hydrothermal vent polychaete nest, reclassification of Thiomicrospira denitrificans as Sulfurimonas denitrificans comb. nov. and emended description of the genus Sulfurimonas. International Journal of System-atic and Evolutionary Microbiology 56, 1725–1733.

Takano, Y., Chikaraishi, Y., Ogawa, N.O., Nomaki, H., Morono, Y., Inagaki, F., Kitazato, H., Hinrichs, K.-U., Ohkouchi, N., 2010. Sedimentary membrane lipids recycled by deep-sea benthic archaea. Nature Geoscience 3, 858–861.

Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M., and Kumar, S., 2011. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution 28, 2731–2739.

Page 226: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

218

Tamura, K., Stecher, G., Peterson, D., Filipski, A., Kumar, S., 2013. MEGA6: Molecular evolu-tionary genetics analysis version 6.0. Molecular Biology and Evolution 30, 2725–2729.

Taylor, J., Parkes, R.J., 1983. The cellular fatty-acids of the sulfate-reducing bacteria, Desulfo-bacter sp., Desulfobulbus sp. and Desulfovibrio desulfuricans. Journal of General Microbiology 129, 3303–3309.

Teske, A., Nelson, D.C., 2006. The genera Beggiatoa and Thioploca, 784–810, in: Dworkin, M. (ed), The Prokaryotes 6: Gamma subclass. Springer New York.

Thamdrup, B., Dalsgaard, T., 2002. Production of N2 through anaerobic ammonium oxidation coupled to nitrate reduction in marine sediments. Applied and Environmental Microbiology 68, 1312–1318.

Thomsen, U., Kristensen, E., 1997. Dynamics of ∑CO2 in a surficial sandy marine sediment: The role of chemoautotrophy. Aquatic Microbial Ecology 12, 165–176.

Throbäck, I.N., Enwall, K., Jarvis, A., Hallin, S., 2004. Reassessing PCR primers targeting nirS, nirK and nosZ genes for community surveys of denitrifying bacteria with DGGE. FEMS Microbiology Ecology 49, 401–417.

Tiedje, J.M., Sexstone, A.J., Myrold, D.D., Robinson, J.A., 1982. Denitrification: ecological nich-es, competition and survival. Antonie van Leeuwenhoek 48, 569–583.

Tiquia, S.M., Masson, S.A., Devol, A., 2006. Vertical distribution of nitrite reductase genes (nirS) in continental margin sediments of the Gulf of Mexico. FEMS Microbiology Ecolo-gy 58, 464–475.

Torti, A., Lever, M.A., Jørgensen, B.B., 2015. Origin, dynamics, and implications of extracellu-lar DNA pools in marine sediments. Marine Genomics 24, 185–196.

Tourna, M., Stieglmeier, M., Spang, A., Könneke, M., Schintlmeister, A., Urich, T., 2011. Nitrososphaera viennensis, an ammonia oxidizing archaeon from soil. Proceedings of the National Academy of Sciences of the USA 108, 8420–8425.

Treusch, A.H., Leininger, S., Kletzin, A., Schuster, S.C., Klenk, H.-P., Schleper, C., 2005. Novel genes for nitrite reductase and Amo-related proteins indicate a role of uncultivated meso-philic crenarchaeota in nitrogen cycling. Environmental Microbiology 7, 1985–1995.

Urakawa, H., Martens-Habbena, W., Huguet, C., de la Torre, J.R., Ingalls, A.E., Devol, A.H., Stahl, D.A. 2014. Ammonia availability shapes the seasonal distribution and activity of archaeal and bacterial ammonia oxidizers in the Puget Sound Estuary. Limnology and Oceanog-raphy 59, 1321–1335.

Valentine, D.L., 2007. Adaptations to energy stress dictate the ecology and evolution of the Ar-chaea. Nature Reviews Microbiology 5, 1070–1077.

Vaquer-Sunyer, R., Duarte, C.M., 2008. Thresholds of hypoxia for marine biodiversity. Proceed-ings of the National Academy of Sciences of the USA 105, 15452–15457.

Vasquez-Cardenas, D., 2016. Bacterial chemoautotrophy in coastal sediments. PhD Disserta-tion. University of Amsterdam, Amsterdam, The Netherlands.

Vasquez-Cardenas, D., van de Vossenberg, J., Polerecky, L., Malkin, S.Y., Schauer, R., Hidal-go-Martinez, S., Confurius, V., Middelburg, J.J., Meysman, F.J., Boschker, H.T., 2015. Micro-bial carbon metabolism associated with electrogenic sulphur oxidation in coastal sediments. The ISME Journal 9, 1966–1978.

Venter, J.C., Remington, K., Heidelberg, J.F., Halpern, A.L., Rusch, D., Eisen, J.A., Wu, D., Paulsen, I., Nelson, K.E., Nelson, W., Fouts, D.E., Levy, S., Knap, A.H., Lomas, M.W., Ne-alson, K., White, O., Peterson, J., Hoffman, J., Parsons, R., Baden-Tillson, H., Pfannkoch, C., Rogers, Y.-H., Smith, H.O., 2004. Environmental genome shotgun sequencing of the Sargasso Sea. Science 304, 66–74.

Page 227: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

219

Vestal, J.R., White, D.C., 1989. Lipid Analysis in Microbial Ecology. BioScience 39, 535–541. Villanueva, L., Haveman, S.A., Summers, Z.M., Lovley, D.R., 2008. Quantification of Desul-

fovibrio vulgaris dissimilatory sulfite reductase gene expression during electron donor- and electron acceptor-limited growth. Applied and Environmental Microbiology 74, 5850–5853.

Villanueva, L., Sinninghe Damsté, J.S., Schouten, S., 2014. A re-evaluation of the archaeal membrane lipid biosynthetic pathway. Nature Reviews Microbiology 12, 438–448.

Villanueva, L., Schouten, S., Sinninghe Damsté, J.S., 2017. Phylogenomic analysis of lipid biosynthetic genes of Archaea shed light on the “lipid divide.” Environmental Microbiology 19, 54–69.

Vissers, E.W., Anselmetti, F.S., Bodelier, P.L.E., Muyzer, G., Schleper, C., Tourna, M., Laan-broek, H.J., 2013. Temporal and spatial coexistence of archaeal and bacterial amoA genes and gene transcripts in Lake Lucerne. Hindawi Publishing Corpoeration - Archaea, Article ID 289478. doi:10.1155/2013/289478.

Větrovský, T., Baldrian, P. 2013. The variability of the 16S rRNA gene in bacterial genomes and its consequences for bacterial community analyses. PLoS ONE 8, e57923. doi:10.1371/journal.pone. 0057923.

Vossenberg, van de, J., Woebken, D., Maalcke, W.J., Wessels, H.J.C.T., Dutilh, B.E., Kartal, B., Janssen-Megens, E.M., Roeselers, G., Yan, J., Speth, D., Gloerich, J., Geerts, W., van der Biezen, E., Pluk, W., Francoijs, K.-J., Russ, L., Lam, P., Malfatti, S. a, Tringe, S.G., Haaijer, S.C.M., Op den Camp, H.J.M., Stunnenberg, H.G., Amann, R., Kuypers, M.M.M., Jetten, M.S.M., 2013. The metagenome of the marine anammox bacterium “Candidatus Scalindua profunda” illustrates the versatility of this globally important nitrogen cycle bacterium. Environmental Microbiology 15, 1275–1289.

Wakeham, S.G., Lewis, C.M., Hopmans, E.C., Schouten, S., Sinninghe Damsté, J.S., 2003. Archaea mediate anaerobic oxidation of methane in deep euxinic waters of the Black Sea. Geochimica et Cosmochimica Acta 67, 1359–1374.

Walker, C.B., de la Torre, J.R., Klotz, M.G., Urakawa, H., Pinel, N., Arp, D.J., Brochier-Ar-manet, C., Chain, P.S.G., Chan, P.P., Gollabgir, A., Hemp, J., Hügler, M., Karr, E.A., Kön-neke, M., Shin, M., Lawton, T.J., Lowe, T., Martens-Habbena, W., Sayavedra-Soto, L.A., Lang, D., Sievert, S.M., Rosenzweig, A.C., Manning, G., Stahl, D.A., 2010. Nitrosopumilus maritimus genome reveals unique mechanisms for nitrification and autotrophy in globally distributed marine crenarchaea. Proceedings of the National Academy of Sciences of the USA 107, 8818–8823.

Ward, B., 2011. Nitrification: An Introduction and Overview of the State of the Field, p 3–8, in: Arp, D., Klotz, M. (ed) Nitrification, ASM Press, Washington, DC.

Waters, E., Hohn, M.J., Ahel, I., Graham, D.E., Adams, M.D., Barnstead, M., Beeson, K.Y., Bibbs, L., Bolanos, R., Keller, M., Kretz, K., Lin, X., Mathur, E., Ni, J., Podar, M., Richard-son, T., Sutton, G.G., Simon, M., Soll, D., Stetter, K.O., Short, J.M., Noordewier, M., 2003. The genome of Nanoarchaeum equitans: Insights into early archaeal evolution and derived parasitism. Proceedings of the National Academy of Sciences of the USA 100, 12984–12988.

Webster, G., Watt, L.C., Rinna, J., Fry, J.C., Evershed, R.P., Parkes, R.J., Weightman, A.J., 2006. A comparison of stable-isotope probing of DNA and phospholipid fatty acids to study prokaryotic functional diversity in sulfate-reducing marine sediment enrichment slurries. Environmental Microbiology 8, 1575–1589.

Weijers, J.W.H., Schouten, S., Spaargaren, O.C., Sinninghe Damsté, J.S., 2006. Occurrence and distribution of tetraether membrane lipids in soils: Implications for the use of the TEX86 proxy and the BIT index. Organic Geochemistry 37, 1680–1693.

Page 228: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

220

Wenk, C.B., Blees, J., Zopfi, J., Veronesi, M., Bourbonnais, A., Schubert, C.J., Niemann, H., Lehmann, M.F., 2013. Anaerobic ammonium oxidation (anammox) bacteria and sulfide-de-pendent denitrifiers coexist in the water column of a meromictic south-alpine lake. Limnolo-gy and Oceanography 58, 1–12.

Werner, J.J., Koren, O., Hugenholtz, P., DeSantis, T.Z., Walters, W.A., Caporaso, J.G., Angenent, L.T., Knight, R., Ley, R.E., 2012. Impact of training sets on classification of high-through-put bacterial 16S rRNA gene surveys. The ISME Journal 6, 94–103.

Weston, K., Fernand, L., Nicholls, J., Marca-Bell, A., Mills, D., Sivyer, D., Trimmer, M., 2008. Sedimentary and water column processes in the Oyster Grounds: a potentially hypoxic region of the North Sea. Marine Environmental Research 65, 235–249.

Wetsteyn, L.P.M.J., 2011. Grevelingenmeer: meer kwetsbaar? Een Beschrijving van de Ecologi-sche Ontwikkelingen voor de Periode 1999 t/m 2008-2010 in Vergelijking met de Periode 1990 t/m 1998. Rijkswaterstaat, Ministerie van Infrastructuur en Milieu, 1–163.

White, D.C., Davis, W.M., Nickels, J.S., King, J.D., Bobbie, R.J., 1979. Determination of the sedimentary microbial biomass by extractible lipid phosphate. Oecologica 62, 51–62.

White, D.C., Hedrick, D.B., Peacock, A.D., 2007. Lipid analyses for viable microbial biomass, community composition, metabolic status, and in situ metabolism, 112–125, in: Hurst, C.,Crawford, R., Garland, J., Lipson, D., Mills, A., Stetzenbach, L. (ed) Manual of Environ-mental Microbiology, 3rd ed., ASM Press, Washington, DC.

Woebken, D., Fuchs, B.M., Kuypers, M.M.M., Amann, R., 2007. Potential interactions of particle-associated anammox bacteria with bacterial and archaeal partners in the Namibian upwelling system. Applied and Environmental Microbiology 73, 4648–4657.

Woebken, D., Lam, P., Kuypers, M.M.M., Naqvi, S.W.A., Kartal, B., Strous, M., Jetten, M.S.M., Fuchs, B.M., Amann, R., 2008. A microdiversity study of anammox bacteria reveals a novel Candidatus Scalindua phylotype in marine oxygen minimum zones. Environmental Microbiology 10, 3106–3119.

Woese, C.R., Fox, G.E., 1977. Phylogenetic structure of the prokaryotic domain: The primary kingdoms. Proceedings of the National Academy of Sciences of the USA 74, 5088–5090.

Wuchter, C., Schouten, S., Boschker, H.T.S., Sinninghe Damsté, J.S., 2003. Bicarbonate uptake by marine Crenarchaeota. FEMS Microbiology Letters 219, 203–207.

Wuchter, C., Schouten, S., Coolen, M.J.L., Sinninghe Damsté, J.S., 2004. Temperature-depen-dent variation in the distribution of tetraether membrane lipids of marine Crenarchaeota: Implications for TEX86 paleothermometry. Paleoceanography 19, PA4028. doi:10.1029/2004PA001041.

Wuchter, C., Schouten, S., Wakeham, S.G., Sinninghe Damsté, J.S., 2005. Temporal and spatial variation in tetraether membrane lipids of marine Crenarchaeota in particulate organic mat-ter: Implications for TEX86 paleothermometry. Paleoceanography 20, PA3013. doi:10.1029/2004PA001110.

Wuchter, C., Schouten, S., Wakeham, S.G., Sinninghe Damsté, J.S., 2006. Archaeal tetraether membrane lipid fluxes in the northeastern Pacific and the Arabian Sea: Implications for TEX86 paleothermometry. Paleoceanography 21, PA4208. doi:10.1029/2006PA001279.

Xie, S., Lipp, J.S., Wegener, G., Ferdelman, T.G., Hinrichs, K., 2013. Turnover of microbial lipids in the deep biosphere and growth of benthic archaeal populations. Proceedings of the National Academy of Sciences of the USA 110, 6010–6014.

Yakimov, M.M., La Cono, V., Denaro, R., 2009. A first insight into the occurrence and expres-sion of functional amoA and accA genes of autotrophic and ammonia-oxidizing bathypelag-ic Crenarchaeota of Tyrrhenian Sea. Deep Sea Research Part II 56, 748–754.

Page 229: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

221

Yakimov, M.M., La Cono, V., Smedile, F., DeLuca, T.H., Juárez, S., Ciordia, S., Fernández, M., Albar, J.P., Ferrer, M., Golyshin, P.N., Giuliano, L., 2011. Contribution of crenarchaeal autotrophic ammonia oxidizers to the dark primary production in Tyrrhenian deep waters (Central Mediterranean Sea). The ISME Journal 5, 945–961.

Yan, J., Haaijer, S.C.M., Op den Camp, H.J.M., van Niftrik, L., Stahl, D.A., Könneke, M., Rush, D., Sinninghe Damsté, J.S., Hu, Y.Y., Jetten, M.S.M., 2012. Mimicking the oxygen minimum zones: stimulating interaction of aerobic archaeal and anaerobic bacterial ammonia oxidiz-ers in a laboratory-scale model system. Environmental Microbiology 14, 3146–3158.

Ye, Q., Wu, Y., Zhu, Z., Wang, X., Li, Z., Zhang, J., 2016. Bacterial diversity in the surface sedi-ments of the hypoxic zone near the Changjiang Estuary and in the East China Sea. Microbi-ologyOpen 5, 323–339. doi:10.1002/mbo3.330.

Zhang, C.L., Huang, Z., Cantu, J., Pancost, R.D., Brigmon, R.L., Lyons, T.W., Sassen, R., 2005. Lipid biomarkers and carbon isotope signatures of a microbial (Beggiatoa) mat associated with gas hydrates in the Gulf of Mexico. Applied and Environmental Microbiology 71, 2106–2112.

Zhang, C.L., Pearson, A., Li, Y.L., Mills, G., Wiegel, J., 2006. Thermophilic temperature optimum for crenarchaeol synthesis and its implication for archaeal evolution. Applied and Environmental Microbiology 72, 4419–4422.

Zhang, X., Agogué, H., Dupuy, C., Gong, J. 2013. Relative abundance of ammonia oxidizers, denitrifiers, and anammox bacteria in sediments of hyper-nutrified estuarine tidal flats and in relation to environmental conditions. CLEAN - Soil, Air, Water 41 (9999), 1–9. doi:10.1002/clen.201300013.

Zhang, Y.G., Zhang, C.L., Liu, X.L., Li, L., Hinrichs, K.U., Noakes, J.E., 2011. Methane Index: A tetraether archaeal lipid biomarker indicator for detecting the instability of marine gas hydrates. Earth and Planetary Science Letters 307, 525–534.

Zhang, Y., Xie, X., Jiao, N., Hsiao, S.S.-Y., Kao, S.-J., 2014. Diversity and distribution of amoA-type nitrifying and nirS-type denitrifying microbial communities in the Yangtze River estu-ary. Biogeosciences 11, 2131–2145. doi:10.5194/bg-11-2131-2014.

Zink, K.G., Wilkes, H., Disko, U., Elvert, M., Horsfield, B., 2003. Intact phospholipids - Micro-bial “life markers” in marine deep subsurface sediments. Organic Geochemistry 34, 755–769.

Zuckerkandl, E., Pauling, L. 1965. Evolutionary divergence and convergence in proteins, 97-166, in: Bryson, V., Vogel, H.J. (ed) Evolving Genes and Proteins, Academic Press, New York.

Zumft, W.G., 1997. Cell biology and molecular basis of denitrification. Microbiology and Molecular Biology Reviews 61, 533–616.

Page 230: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well
Page 231: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

223

Acknowledgements

During the past years, I have made experiences that I would have missed, if I would not have worked at the NIOZ on Texel. There are many people I’d like to thank for making it a good experience.

First of all, to my supervisors, Laura, thank you so much for everything and especially for your outstanding support and patience over the past years. It has been a rocky road and I have learnt a lot from you scientifically but also about living far away of your family on a small island in the North Sea. Thank you very much for your open ears whenever I needed you for assistance or just to motivate myself again. You have been the best supervisor that I can imagine! This is not self-evident in this rushing times! Jaap, thanks a lot for your time and effort, reviewing the publications and the thesis, especially in the last phase of the process. It was a great pleasure to work with you. Stefan, thank you very much for your constructive comments and discussions especially during my first phase, I have learnt a lot from you, especially how to deal with scientific articles during our literature discussions. Thank you all for always leaving your door open!

Elda, Harry, Judith and Anneke, thanks a lot for hosting and teaching me in the molecular lab. It was a great pleasure to work with you and as well to chat and sing along during long hours in the lab. You made me feel very welcome in the old and brown part of the building. Elda, thank you for offering your desk to me. It was my favorite spot!

Ellen and Nicole, thanks a lot for sharing your expertise and teaching me in the lipid lab, where I was completely unexperienced before I came to the NIOZ. Jort, Monique, Denise, Anchelique and Irene, thanks for your help with extraction methods, machines and data analysis as well for cheering me up a lot! Marianne, thank you very much for your support in the lab, but especially for your help and nice chats during our cruise to and around Iceland. I will always remember that trip! Sebastiaan, thank you very much for technical and human support on the cruise and during my time at NIOZ, I’ve never met a gentler person with such a dry and clever sense of humor!

Sabine, thanks a lot for the nice collaboration during the Iceland cruise. I have learnt a lot from you scientifically, but also about life! It was my pleasure to meet you and get to know you a bit closer. You are an amazing person! Don’t ever change for others!

Marcel and Sandra, I would like to thank you for the nice collaboration during the Grevelingen campaign. It was funny and exciting! Additionally, a big thanks goes to the people from NIOZ Yerseke. Eric and Filip, thanks for inviting us to join the Grevelingen campaign. Diana, a special thanks to you for the excellent collaboration and your support. Dorina, Fatimah and Mathilde, thank you very much for the good collaboration and your support with data. Alexandra Vasquez Cardenas, thanks a lot for providing Figure 2 of the introduction.

The crazy thing about this small island in the North Sea and the NIOZ is, that I got to know the nicest people from all over the world. Some of them stay friends forever! Jenni, Martina, Marta, Catia, Claudia, Santi and Mireia, you will be always on my mind! Alex, Tristan, Maram, Johann, Santi and Sergio, thank you very much for being the nicest neighbors and friends in the special Potvis situation, which welds together! Thank you for sharing so many things with me, good and bad times! I am looking forward to see you again!

Page 232: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

224

A big thanks to all the people in the MMB department (and others) making work and life more enjoyable; Marc, Daniela, Yuki, Andreas, Darci, Sabine, Petra, Raquel, Elisabeth, Lisa, Eli, Sandra, Dave, John, Marcel, Sebastian, Cindy, Rob, Luke, Frederike, Rodrigo, and Kevin thanks for your company!

My office mates over the past years had to stand me and my moody phases, but they did well and cheered me up a lot! Lisa, Cecile, Sandra, Mireia, Eva and Michel, thanks for the nice time!

Down at NIOZ tropical, I have met some amazing people; Christine, Francesc, Francesco and Michele, it was very nice to talk about other things than science and enjoy your culinary skills.

Ashes over my head, but I have spent some time in the smoking room, and as usual, smokers make friends quite quickly. Kees, Ronald, Geert-Jan, Anneke and Edward, thanks a lot for nice chats and a bit of distraction.

For the “new” PhD students, Julie, Marijke, Gabriella, Saara, Laura, Caitlyn and Sophie; Success for your future and good luck with your thesis!

Michael Hagemann, Danke für Deine Hilfe in den letzten Zügen! Ohne Deinen scharfen Verstand und die Bereitschaft zu helfen, hätte ich es nicht gepackt!

Dieser Dank gilt allen meinen Freundinnen, die mich auf dem Weg ermutigt und unterstützt haben (wenn auch manchmal nur am Telefon); Miri, Helena, Moni, Svenja, Anni und Marén, ihr seid großartige Wegbegleiter!

Miri, Du bist die einzige, die wirklich versteht, was es heißt, auf Texel zu wohnen! Schön, daß wir die Erfahrung teilen!

Jetzt aber das Wichtigste, die Familie! Mama, Reimund und Papa, euch verdanke ich wer und wo ich bin! Ohne Eure langjährige Unterstützung und eure Geduld, hätte ich es nicht geschafft! Danke für Alles! Tobias, mein Bruderherz, Du bist der beste Bruder der Welt! Danke für Deine Unterstützung und Treue! Klausi, mein Herz, Danke, daß Du immer für mich da warst! Es war eine lange Reise, ich bin gespannt wo es als nächstes hingeht!

Page 233: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well

225

About the Author

Yvonne Antonia Lipsewers was born on the 18th of July 1977 in Salzkotten, Germany. During her Abitur she gained interest in aquatic biology and was motivated by her biology teacher with a seminar about hydrology. After leaving secondary school, she first started studying food technol-ogy before finally, studying Biology at the University of Bielefeld from 2003 until 2007. After-wards, she was specializing in Marine Microbiology during her Master’s at the Institute of Chem-istry and Biology of the Marine Environment associated to the Carl-von-Ossietzky Universität in Oldenburg from 2008 until 2010. She completed a Master’s research project investigating the shift of the heterotrophic bacterial community composition during the aggregation of a diatom bloom in the North Sea. After graduating in 2010, Yvonne started a PhD project at NIOZ in 2011, focused on tracing chemolithoautotrophic microorganisms in coastal sediments, under the supervision of Dr. Laura Villanueva, which finally led to the composition of this thesis.

Page 234: Role of chemolithoautotrophic microorganisms involved in ... · of chemolithoautotrophic microorganisms involved in nitrogen and sulfur cycling of coastal marine sediments, as well