162
Structural Characterization of the Flavonoid Enzyme Complex Christopher David Dana Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biology B.S.J. Winkel, Chair D.R. Bevan J.M. McDowell J.C. Sible R.A. Walker July 8, 2004 Blacksburg, VA Keywords: Flavonoids, Arabidopsis, chalcone synthase, chalcone isomerase Copyright 2004, Christopher Dana

Structural Characterization of the Flavonoid Enzyme Complex

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Structural Characterization of the Flavonoid Enzyme Complex

Structural Characterization of the Flavonoid Enzyme Complex

Christopher David Dana

Dissertation submitted to the faculty of the

Virginia Polytechnic Institute and State University

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in

Biology

B.S.J. Winkel, Chair

D.R. Bevan

J.M. McDowell

J.C. Sible

R.A. Walker

July 8, 2004

Blacksburg, VA

Keywords: Flavonoids, Arabidopsis, chalcone synthase, chalcone isomerase

Copyright 2004, Christopher Dana

Page 2: Structural Characterization of the Flavonoid Enzyme Complex

Structural Characterization of the Flavonoid Enzyme Complex

Christopher David Dana

Abstract

Flavonoid biosynthesis is an important secondary metabolic pathway in higher

plants with a range of vital functions in plants and animals. This pathway has been

developed as a model system for the study of multi-enzyme complexes. The goal of the

work presented here was to structurally characterize a series of loss-of-function chalcone

synthase (CHS) alleles and to define the molecular basis of the interaction between CHS

and the second enzyme of flavonoid biosynthesis, chalcone isomerase (CHI).

CHS proteins encoded by five previously characterized alleles were characterized

by homology modeling in an effort to explain the alterations in function, stability, and

dimerization exhibited by these variants. Four of the encoded proteins have a single

amino acid substitution and the fifth is a truncated protein resulting from a frameshift.

Models for each of these proteins were generated in silico and analyzed after molecular

dynamics simulations. This analysis suggested reasons for changes in catalytic ability

and stability for three of the five CHS variants.

To characterize the molecular basis of the CHS-CHI interaction, a model was

developed using X-ray crystallography, small-angle neutron scattering (SANS), in silico

docking, molecular dynamics simulations, and yeast 2-hybrid analyses. These enzymes

appear to be interacting in a manner that could facilitate the flow of intermediates from

one active site to another. These experiments also identified a series of amino acids that

appear to be involved in the interaction, which are currently undergoing alteration and

Page 3: Structural Characterization of the Flavonoid Enzyme Complex

iii

analysis using a yeast 2-hybrid assay to verify the authenticity of the model. The data

presented herein could be used in future engineering experiments to alter pathway flux to

control the levels or types of flavonoid endproducts, resulting in more nutritious plants or

flowers with novel pigments.

These experiments advance the study of the structure of multi-enzyme complexes,

an area that currently contains little information. As well, this is the first known use of

SANS for the investigation of the architecture of metabolons. The techniques described

herein could easily be applied to other systems in an effort to better understand the

organization of multi-enzyme complexes and the implications of these assemblies on

metabolic regulation.

Page 4: Structural Characterization of the Flavonoid Enzyme Complex

iv

Dedication

This work is dedicated to my family, especially my parents and brother, David,

Donna, and William Dana, who have been my biggest cheerleaders throughout this

endeavor. They have always urged me to pursue my dreams, and without them and the

rest of my family, I surely would not be who or where I am today.

Page 5: Structural Characterization of the Flavonoid Enzyme Complex

v

Acknowledgements

There are, indeed, many people to thank for their help and support during my time

at Virginia Tech. First and foremost, I would like to thank my advisor, Dr. Brenda

Winkel, for allowing me to add my two cents to this amazing story. She patiently

listened to my sometimes hair-brained schemes on how to approach an experiment, and

offered invaluable support and guidance when something didn’t go quite as expected. I

also would like to express my gratitude to Dr. David Bevan, Dr. John McDowell, Dr. Jill

Sible, and Dr. Rich Walker for serving on my advisory committee and offering nothing

but support throughout my time here. Finally, I would like to thank all of the laboratories

in both Derring Hall and Fralin Biotechnology Center that have provided advice,

assistance, and facilities over the years.

This work could have not been done without the support of my collaborators. Dr.

David Bevan and Jory Zmuda provided invaluable assistance and advice on the molecular

dynamics simulations. Dr. Joe Noel, Dr. Joe Jez, Dr. Steph Richards, Mike Austin, and

Marianne Bowman at the Salk Institute introduced me to protein crystallography. Dr.

Susan Krueger at the NIST center for neutron research, introduced me to small angle

neutron scattering, and with her unwavering help and support, this project has come to

fruition.

I would also like to thank the members of the Winkel lab, both past and present

for their support and general joviality, which kept the environment of the lab good. I

thank Dr. David Saslowsky for guiding me along and putting up with my incessant

questions when I first got here. Many thanks go to Dr. Mike Santos, who handled my

teasing quite well, and was a wonderful mentor for the molecular biology lab. My

Page 6: Structural Characterization of the Flavonoid Enzyme Complex

vi

gratitude also goes to Daniel “The Notorious HPLC” Owens for the many conversations

about science, life, and music over the years, for they have brightened my day on more

than one occasion. I would also like to thank Dr. Anne Walton, Dr. Ujwala Warek,

Melissa Ramirez, Ana Frazzon, and Michelle Barthet for their help and advice over the

years. In the cast of Winkel lab members, the group of undergrads cannot be overlooked.

Over the years, there have been many wonderful undergrads that have come through the

lab. My thanks go out to all of them and everything they contributed to the project and to

me, but most especially, Ashley Duda, Sandra Jabre, Anna Leung, Eric McFadden, Anna

Watson, Lori Hill, and Liz Tucker.

I have been extremely fortunate to have been surrounded by an amazing group of

people in my time here. First and foremost, I would like to thank James Glazebrook,

director of the New River Valley Symphony at Virginia Tech for allowing me the honor

of playing with his orchestra during my time here, it provided many hours of enjoyment,

learning, and stress relief. I would also like to thank Dr. Robert Goebel at James

Madison University for many hours of laughter from the puns, the reading suggestions

and helping me to look at things beyond science. Special thanks go to Becky Abler, Tom

Bunt, Mike Crandall, and Laura Moffett for their amazing friendship and support, which

has meant more to the completion of this dissertation than any of you will know. Many

thanks also go out to Tracey and Douglas Slotta for all of the support and kindness they

have shown me over the years. Friends in “The Other Group” have provided hours of

laughter and fun, without which I would have gone mad. For those friends who weren’t

mentioned here, your support and friendship has not gone unnoticed and has been a

source of strength for me.

Page 7: Structural Characterization of the Flavonoid Enzyme Complex

vii

Table of Contents

Chapter 1: Literature Review……………………………………………………………...1

Introduction………………………………………………………………………..2

Metabolons………………………………………………………………………...3

Static vs. Dynamic Metabolons…………………………………………………...4

Metabolic Complexes and Whole Cell Visualization……………………………..4

Phenylpropanoid and Flavonoid Biosynthesis…………………………………….5

Regulation of Flavonoid Biosynthesis…………………………………………….8

Evidence for a Flavonoid Metabolon………………………………………...........9

Arabidopsis Flavonoid Biosynthesis as a Model System for Studying Metabolic

Pathways…………………………………………………………………………11

Structural Biology of Flavonoid Metabolism………………………………........12

Chalcone Synthase………………………………………………….........13

Chalcone Isomerase…………………………..……………………….....14

Methods to Determine Interactions Between Proteins…………………………..16

Small-Angle Neutron Scattering………………………………………....16

Surface Plasmon Resonance……………………………...……………...17

Project Goals/Rational……………………………………………...……………20

References Cited……………………………………………………...……….....22

Page 8: Structural Characterization of the Flavonoid Enzyme Complex

viii

Chapter 2: Interpretation of Functional Effects of Mutations in Arabidopsis Chalcone

Synthase Based on Molecular Modeling………………………………………….……..44

Abstract……………………………………………………………………….…45

Introduction……………………………………………………………………...46

Methods ………………………………………………………………………....48

Results…………………………………………………………………………...50

tt4(85)…………………………………………………………………….51

tt4(UV113)……………………………………………………………….53

tt4(38G1R)……………………………………………………………….54

tt4(UV01)………………………………………………………………...55

tt4(UV25)………………………………………………………………...56

Discussion………………………………………………………………………..57

Acknowledgements…………………………………………………………........59

References Cited…………………………………………………………………60

Chapter 3: Structural Characterization of the Chalcone Synthase-Chalcone Isomerase

Interaction………………………………………………………………………………..75

Abstract………………………………………………………………………….76

Introduction…………..………………………………………………………….77

Materials and Methods…………………………………………………………..80

Site Directed Mutagenesis of pET32a(+)………………………………..80

Creation of Expression Constructs…………………………………........81

Expression and Purification of pCD1-CHS and CHI………………........82

Page 9: Structural Characterization of the Flavonoid Enzyme Complex

ix

Determination of the CHS Crystal Structure…………………………….84

Multiple Sequence Alignment of CHS Proteins…………………............85

Site-Directed Mutagenesis……………………………………………….85

Yeast 2-Hybrid Analysis…………………………………………………86

Small-Angle Neutron Scattering…………………………………............87

In Silico Docking………………………………………………………...88

Molecular Dynamics Simulations……………………………………….88

Multiple Sequence Alignment of CHS and CHI………………………...89

Results…………………………………………………………………………...91

CHS Structure……………………………………………………………91

Experimental Evidence for a Role of Specific Residues in CHS in

Assembly of the Flavonoid Enzyme Complex ………………………….91

Small-Angle Neutron Scattering…………………………………............93

In Silico Docking and Molecular Dynamics Analysis of the CHS-CHI

Complex…………………………………………………………………94

Conservation of Residues Involved in the CHS-CHI Interaction……….96

Discussion……………………………………………………………………….97

Acknowledgements…..…………………………………………………...........100

References Cited…………………………………………………………..........101

Chapter 4: Conclusions…………………………………………………………………129

Appendix A: Homology Modeling of Flavonoid Enzymes……………………………135

Page 10: Structural Characterization of the Flavonoid Enzyme Complex

x

Appendix B: SPR Analysis of Flavonoid Enzymes……………………………………139

Page 11: Structural Characterization of the Flavonoid Enzyme Complex

xi

List of Tables and Figures

Chapter 1

Figure 1.1: Schematic of the flavonoid pathway in Arabidopsis

thaliana………………………………………………………………….36

Figure 1.2: Reaction catalyzed by CHS..………………………………..37

Figure 1.3: Structure of Medicago sativa CHS2..……………………….38

Figure 1.4: Close up of CHS2 active site………………………………..39

Figure 1.5: The reaction catalyzed by CHI…………….………………..40

Figure 1.6: The structure of CHI from Medicago sativa………………...41

Figure 1.7: Schematic representation of SANS………………………….42

Figure 1.8: Schematic representation of the principle of SPR…………...43

Chapter 2

Table 2.1: Biochemical and genetic characteristics of Arabidopsis CHS

alleles……………………………………………………………………63

Table 2.2: Dimensions of the CHS proteins.……………………………64

Table 2.3: Phi-psi torsion angles within the CHS active site……………65

Table 2.4: Distances between CoA binding residues.…………………...66

Figure 2.1: RMSD of the backbone atoms for wild-type CHS…………67

Figure 2.2: Multiple sequence alignment of selected regions of a

representative group of plant type III polyketide synthases……………..69

Figure 2.3: Structure of the Arabidopsis wild-type CHS enzyme….……71

Page 12: Structural Characterization of the Flavonoid Enzyme Complex

xii

Figure 2.4: Models of the Arabidopsis wild-type and TT4 CHS

enzymes…………………………………………………………………..73

Chapter 3

Table 3.1: Crystallographic data…………………….………………….110

Table 3.2: Primers used for PCR and site-directed mutagenesis for yeast 2-

hybrid constructions………………………………………………….....111

Table 3.3: Yeast 2-hybrid analysis of amino acid substitutions…….….112

Table 3.4: Neutron Scattering Parameters for CHS and CHI in

solution………………………………………………………………….113

Table 3.5: RMSD and distance measurements of secondary structure and

residues proposed to be involved in the CHS-CHI interaction..………..114

Figure 3.1: Schematic of the flavonoid pathway in Arabidopsis

thaliana…………………………………………………………………115

Figure 3.2: CHS crystal structure………………………………………116

Figure 3.3: Residues on CHS tested for effects on interactions with other

flavonoid enzymes by yeast 2-hybrid…………………………………..117

Figure 3.4: Normalized SANS data…….………………………………118

Figure 3.5: Data fit of the docked CHS-CHI complex…………………120

Figure 3.6: Proposed structure of the CHS-CHI enzyme complex……..121

Figure 3.7: Movements of CHI beta sheets…….………………………122

Figure 3.8: Residues proposed to participate in the CHS-CHI

interaction………………………………………………………………123

Page 13: Structural Characterization of the Flavonoid Enzyme Complex

xiii

Figure 3.9: Close up of the interaction interface between CHS and

CHI……………………………………………………………………...124

Figure 3.10: Multiple sequence alignments of Arabidopsis CHS and CHI

protein sequences……………………………………………………….125

Appendix A

Figure A.1: Homology models of Arabidopsis Flavonoid Enzymes…...137

Appendix B

Figure B.1: Representative SPR output………………………………...141

Curriculum vitae………………………………………………………………………..143

Page 14: Structural Characterization of the Flavonoid Enzyme Complex

xiv

List of Abbreviations

ANR – anthocyanidin reductase

ANS – anthocyanidin synthase

BD – binding domain

C4H – cinnamate 4-hydroxylase

CHI – chalcone isomerase

CHS – chalcone synthase

DFR – dihydroflavonol 4-reductase

DTT - dithiothreitol

ER – endoplasmic reticulum

F3H – flavanone 3-hydroxylase

F3’H – flavonoid 3’-hydroxylase

FLS – flavonol synthase

HIV – human immunodeficiency virus

I(0) – forward scattering intensity

IGF – insulin growth factor

OAS-TL – O-acetylserine (thiol) lyase

PAL – phenylalanine ammonia lyase

PKS – polyketide synthase

PS – pyrone synthase

Rg – radius of gyration

rER – rough endoplasmic reticulum

Page 15: Structural Characterization of the Flavonoid Enzyme Complex

xv

RMSD – root mean square deviation

SANS – small-angle neutron scattering

SAT – serine acetyltransferase

SPR – surface plasmon resonance

STS – stilbene synthase

TA – transactiviation domain

TCA – tricarboxylic acid

tt – transparent testa

TRX – thioredoxin

Page 16: Structural Characterization of the Flavonoid Enzyme Complex

Chapter 1

Literature Review

Page 17: Structural Characterization of the Flavonoid Enzyme Complex

2

Introduction

A driving question in the study of biochemistry is how metabolic pathways are organized

within the cell. Evidence for organized metabolism dates back to the 1950s when David Green

first isolated the enzymes of the tricarboxylic acid (TCA) cycle as a complex (Green, 1958).

Subsequent ultracentrifugation experiments with Neurospora (Zalokar, 1960) and Euglena

(Kempner and Miller, 1968) indicated that the aqueous cytoplasm contains little or no free

floating proteins or enzymes and that the interior of the cell is highly organized. These seminal

experiments revolutionized the concept of cellular organization and laid the groundwork for

future experiments aimed at defining metabolic organization within the cell. The advent of

interactome analyses (Ito et al., 2001; Li et al., 2004; Sanchez et al., 1999) provided new insights

into the extent to which cellular proteins are able to interact. As more examples of multienzyme

complexes, or metabolons, were identified, the notion that cellular metabolism is arranged as a

highly organized network of metabolic complexes gained widespread acceptance.

In our laboratory, we are investigating multi-enzyme complex formation in flavonoid

biosynthesis. Flavonoids are secondary metabolites that are important both for the plant and for

organisms that use plants for nutrition. Helen Stafford first suggested that flavonoid metabolism

is organized as one or more enzyme complexes (Stafford, 1974a; Stafford, 1974b), and later

experiments provided evidence for this (for example, Burbulis and Winkel-Shirley, 1999; Fritsch

and Grisebach, 1975; Hrazdina and Jensen, 1992). One of our goals is to determine the

molecular basis of the interaction between the first two enzymes in this pathway, chalcone

synthase (CHS) and chalcone isomerase (CHI), in the hope that this information will lead to a

better understanding of the architecture of enzyme complexes in this and other systems.

Page 18: Structural Characterization of the Flavonoid Enzyme Complex

3

Metabolons

An enzyme complex is a group of two or more enzymes that interact to catalyze related

metabolic reactions, allowing for higher catalytic efficiency than if the enzymes were not

associated. Metabolic pathways can obtain such an efficiency by controlling the flux of substrate

and product thru channels between proteins (Ovádi and Srere, 2000). Metabolic channeling also

protects intermediates that are unstable or rare and maintains concentration gradients within the

complex (Mathews, 1993). The mechanics behind metabolic channeling include the covalent

binding of intermediates to the active site, a physical barrier acting as a tunnel, direct site-to-site

transfer, transfer in an unstirred layer, and electrostatic channels (reviewed in Ovádi and Srere,

2000).

The TCA cycle represents one of the best-characterized multienzyme complexes. Since

David Green isolated the TCA cycle complex, the functional and structural properties of the

TCA cycle complex have been well characterized in the laboratory of the late Paul Srere

(Robinson et al., 1987; Srere, 1985; Srere, 1987; Sumegi et al., 1993; Vélot et al., 1997; Vélot

and Srere, 2000). Srere’s work eventually led to the first three dimensional model of the citrate

synthase-malate dehydrogenase enzyme complex though in silico docking of the crystal

structures (Ovádi and Srere, 2000). There are now many more examples of this level of

organization, and it has been suggested that all metabolic pathways organize in this way

(Beeckmans et al., 1994; Beeckmans et al., 1989; Gontero et al., 1988; Robinson et al., 1987).

Page 19: Structural Characterization of the Flavonoid Enzyme Complex

4

Static vs. Dynamic Metabolons

Enzyme complexes have been classified as being either static or dynamic (Welch, 1977).

Generally, static complexes are very stable over time and the interactions between the enzymes

are so strong that upon purification, the enzymes remain associated because little energy is

needed to maintain the integrity of the complex (Ovádi and Srere, 2000). In contrast, a dynamic

enzyme complex is held together by relatively weak forces and is dependent upon constant

energy dissipation (Welch, 1977) and is composed of multiple branch points where many

enzymes could compete for one substrate (Ovádi and Srere, 2000). It is generally thought that

static enzyme systems possess no branches, whereas dynamic enzyme systems may contain

many branches (Ovádi and Srere, 2000). Examples of static enzyme complexes are the α-keto

acid dehydrogenase complex, the proteasome, and proteins involved in DNA, RNA, and protein

syntesis (reviewed in Ovádi and Srere, 2000; Winkel, 2004). Two examples of dynamic enzyme

systems are the TCA cycle complex (Robinson et al., 1987) and enzymes involved in carbamoyl

phosphate metabolism (Massant and Glansdorff, 2004).

Metabolic Complexes and Whole Cell Visualization

There is a plethora of in vitro and in vivo evidence for protein-protein interactions and

macromolecular crowding within the cell. It has only recently been possible to visualize this

directly; however, there have been few examples of whole cell visualization showing enzyme

complexes and, ultimately, macromolecular crowding. An initial representation of

Page 20: Structural Characterization of the Flavonoid Enzyme Complex

5

macromolecular crowding came from David Goodsell’s drawings of E.coli ribosomes, DNA, and

RNA to shape and scale and illustrated that the interior of the cell is crowded (Goodsell, 1991).

Electron tomography, an established technique, has recently gained new popularity for

visualization of macromolecular organization within cells. By using this technique, Hoppe et al.

(1974) were able to describe a complex of enzymes involved in fatty acid biosynthesis. As

computational power increased, electron tomography could be used to visualize organelles and

later whole cells. Marsh, et. al (2001) recently used this technology to show that the

environment around the Golgi apparatus in pancreatic ß cells is extremely congested. Elegant

work performed in Wolfgang Baumeister’s lab using electron tomographic imaging of whole

Dictyostelium cells showed that the cell is crowded with actin networks, membranes, and

macromolecular assemblies (Medalia et al., 2002).

Phenylpropanoid and Flavonoid Biosynthesis

The phenylpropanoid pathway converts phenylalanine into a variety of secondary

metabolites including lignins, sinapate esters, and stilbenes (Figure 1.1) that have a multitude of

in planta functions, such as reproduction, defense against fungal infections, and signaling in

pathogenesis and symbiosis (reviewed in Winkel-Shirley, 2001). Phenylpropanoid biosynthesis

also feeds into the flavonoid pathway, which converts coumaroyl-CoA and malonyl-CoA into

flavonoids, isoflavonoids, flavonol glycosides, anthocyanidins, and condensed tannins (Figure

1.1). In plants, these compounds have been postulated to function in reproduction, defense

against the biotic and abiotic environment, and signaling in pathogenesis and symbiosis. These

metabolites, along with sinapate esters derived from sinapic acid (Landry et al., 1995; Li et al.,

Page 21: Structural Characterization of the Flavonoid Enzyme Complex

6

1993), protect the plant by absorbing light in the damaging UV-B spectrum (between 280 and

320 nm), acting as natural sunscreens for the plant. Flavonoids have been shown to accumulate

at the epidermal layer of leaf tissue, pollen where UV-B exposure is the highest as well as in the

epidermis of fruit (Chapple et al., 1992; Day, 1993; Jordan et al., 1998; Landry et al., 1995;

Solovchenko and Schmitz-Eiberger, 2003; Strack et al., 1985).

Other in planta activities for flavonoids include defense against pathogen attack,

signaling molecules, and regulators of auxin transport. Defense-related compounds, called

phytoalexins, are produced when the plant comes under attack from pathogenic fungi or bacteria

and, in the case of bacteria, are potent anti-microbial agents (Liu and Dixon, 2001). Certain

flavonoids such as naringenin, quercetin, kaempferol, and rutin, are constitutively produced and

thought to have some anti-microbial properties (Paiva, 2000). However, when a plant comes

under attack from fungal pathogens, some species produce glyceollin and medicarpin (both

isoflavonoid pterocarbons, and are also potent anti-microbial agents) in the foliage surrounding

the infection site (Paiva, 2000). In legumes, a specific class of flavonoids, the isoflavonoids, act

as signaling compounds that initiate the symbiosis of the nitrogen-fixing rhizobia (Hirsch et al.,

2001). Flavonoids have also been implicated in the regulation of auxin transport (Winkel-

Shirley, 2002). Recent studies showed that in Arabidopsis plants that are unable to produce

naringenin, auxin distribution resembles that of a plant overproducing auxin (Brown et al., 2001;

Buer and Muday, 2004; Murphy et al., 2000). Normal auxin distribution is restored upon the

application of exogenous naringenin.

Flavonoids are also responsible for pigmentation in flower, fruit, and vegetative tissue

(especially blue, red, and purple coloration), functioning as attractants for pollinators and seed

dispersers. These pigments have been the phenotype observed in some important scientific

Page 22: Structural Characterization of the Flavonoid Enzyme Complex

7

breakthroughs such as Mendel’s elucidiation of genetics through the study of phenotypes in pea

(Pisum sativa), one being flower coloration. Altering the pigmentation of flowers is of intense

interest to the horticultural industry (Forkmann and Martens, 2001). The first example of

metabolic engineering of flower pigmentation came involved introduction of the maize gene

encoding dihydroflavonol 4-reductase (DFR) into Petunia, yielding flowers with a novel orange

color (Meyer et al., 1987). Recently, the color of Phalenopsis flowers were altered by a transient

transfection method using a putative flavonoid-3’,5’-hydroxylase from the same organism (Su

and Hsu, 2003).

Flavonoids exhibit a host of nutritional and medicinal properties including anti-oxidant

activities, which are useful in reducing the risk of cardiovascular disease and some cancers

(Dixon and Steele, 1999). Other beneficial properties of flavonoids include prevention of

bacterial infections, enzyme inhibition, and free radical scavenging (reviewed in Havsteen,

2002). Daidzin, an isoflavonoid, is a potent anti-diphsotropic agent that has been used as a

Chinese herbal remedy for alcohol dependence for over a millennium (Keung, 2003). In Asian

societies, the low incidence of cardiovascular disease and cancer has been, in part, linked to the

isoflavones produced in soybean, which is regularly consumed (Sarkar and Li, 2003). Dark

chocolate, interestingly, has the highest concentration of anti-oxidant flavonoids among

commonly-consumed foods (Lotito and Fraga, 2000; Serafini et al., 2003).

Creating plants with elevated levels of antioxidant flavonoids is a focus of recent

metabolic engineering projects. Work done at Unilever Research in the United Kingdom and

Plant Research International in the Netherlands demonstrated that expression of petunia CHI in

tomato plants induces a 78-fold increase of fruit peel flavonols without any change in visible

phenotype (Muir et al., 2001). Interestingly, when the fruit was processed, a majority of the

Page 23: Structural Characterization of the Flavonoid Enzyme Complex

8

flavonols were retained. Yu et al. (2003) were able to increase the levels of isoflavones in

soybean seed by suppressing flavanone 3-hydroxylase (F3H), the entry point to flavonol,

condensed tannin, and anthocyanin biosynthesis, and activating expression of two transcription

factors required for isoflavonoid biosynthesis. These metabolic engineering efforts could create

edible plants that supplement diets with anti-oxidants to help protect against cancer and other

diseases.

Regulation of Flavonoid Biosynthesis

Flavonoid biosynthesis is tightly controlled by a number of internal and external factors.

It can be modulated by development, exposure to UV-B radiation, high intensity and blue light,

and both low and high temperatures (reviewed in Winkel-Shirley, 2002). The molecular basis of

this regulation has been intensely studied in parsley, petunia, snapdragon, maize, and

Arabidopsis and numerous transcriptional regulators have been identified and characterized (

reviewed in Marles et al., 2003; Sainz et al., 1997; Weisshaar and Jenkins, 1998).

In maize, two sets of transcription factors, the R family of regulatory proteins and the C1

myb-domain-containing transcription factors, appear to control the transcription of the entire

flavonoid pathway (Taylor and Briggs, 1990). However, other plants that contain homologues of

the maize transcription factors do not exhibit the same breadth of control over the pathway. For

example, an1, 2, 4, and 11 from petunia and delila from snapdragon all appear to be

homologues of the R family of genes in maize, but control the transcription of a more defined

subset of genes (Martin and Gerats, 1993; Spelt et al., 2000).

Page 24: Structural Characterization of the Flavonoid Enzyme Complex

9

Much is also known about the regulation of flavonoid biosynthesis in Arabidopsis. Some

of the transcription factors that control flavonoid biosynthesis in Arabidopsis have been

identified by characterization of the transparent testa (tt) mutants. For example, TTG1, TTG2,

TT2, TT8, and TT16 encode transcription factors that control DFR and anthocyanidin reductase

(ANR) expression in the seed coat, and have been proposed to cooperate in controlling the

accumulation of proanthocyanidins in developing seed coats (Johnson et al., 2002; Nesi et al.,

2000; Nesi et al., 2002; Nesi et al., 2001; Walker et al., 1999). However, there are transcription

factors that are ubiquitous in their ability to regulate the expression of genes. ICX1 is such an

example which has been implicated in the negative regulation of both light and non- light

responses (Wade et al., 2003).

Evidence for a Flavonoid Metabolon

Helen Stafford’s original hypothesis that phenylpropanoid metabolism is organized as

one or more enzyme complexes (Stafford, 1974a; Stafford, 1974b) has been supported by

experiments performed in many laboratories, including our own (reviewed in Winkel, 2004;

Winkel-Shirley, 1999). Through microsomal fractionation experiments, Czichi and Kindl (1977)

demonstrated a close association of the first two enzymes of general phenypropanoid metabolism

[phenylalanine ammonia lyase (PAL) and cinnamate 4-hydroxylase (C4H)] with the endoplasmic

reticulum. Later evidence for channeling between these enzymes was obtained in experiments

where 3H-L-phenylalanine was fed to tobacco cell cultures. Little 3H-trans-cinnamic acid (the

product of PAL) was identified in the cells, suggesting that this product was rapidly channeled

Page 25: Structural Characterization of the Flavonoid Enzyme Complex

10

through the general phenylpropanoid pathway (Rasmussen and Dixon, 1999). Evidence for a

multi-enzyme complex and channeling has also been obtained for the first two enzymes of

isoflavonoid biosynthesis through in planta localization and feeding experiments (Liu and

Dixon, 2001).

The first experimental evidence for protein-protein interactions within the flavonoid

pathway came in cell fractionation experiments that demonstrated that the enzymes are located

on microsomal or endoplasmic reticulum (ER) membranes (Fritsch and Grisebach, 1975).

Sucrose density gradients of Hippeastrum petals demonstrated activity of PAL, C4H, CHS, and

UDP-glucose flavonoid glucosyltransferase in fractions corresponding to ER (Wagner and

Hrazdina, 1984). Fractionation experiments in buckwheat showed that CHS activity

corresponded with the rough ER (rER) and C4H, a P450 hydroxylase and purported membrane

anchor for the pathway (Hrazdina and Wagner, 1985). Through immunolocalization, Hrazdina’s

laboratory was able to show that CHS associated with the cytoplasmic face of the rER (Wagner

and Hrazdina, 1984). Building on these results, affinity chromatography, co-

immunoprecipitation, and yeast two-hybrid experiments provided the first direct evidence for

interactions within flavonoid metabolism (Burbulis and Winkel-Shirley, 1999). Further evidence

was obtained from immunolocalization assays in fixed cells (Saslowsky and Winkel-Shirley,

2001). These assays showed that CHS and CHI co- localize and were consistent with a

previously-proposed role for flavonoid 3’-hydroxylase (F3’H) as a membrane anchor for the

complex (Saslowsky and Winkel-Shirley, 2001; Stafford, 1990). It has also been demonstrated

that by expressing a phage-derived antibody to CHI in transgenic plants, the activity of CHI is

altered, as is the overall flavonoid composition of the plant (Santos et al., 2004). This antibody

could prevent CHI from assembling with other flavonoid enzymes. These recent findings have

Page 26: Structural Characterization of the Flavonoid Enzyme Complex

11

changed the overall concept of the flavonoid metabolon from one of a linear arrangement of

enzymes (Hrazdina, 1992; Stafford, 1974a; Stafford, 1974b) to a multi-enzyme complex with

contacts between multiple proteins (Burbulis and Winkel-Shirley, 1999). The next challenge is

to determine the molecular basis of these interactions. A greater understanding of how enzyme

complexes form and how this assembly is regulated may facilitate the engineering of plants with

enhanced agronomic, nutritional or medicinal characteristics.

Arabidopsis Flavonoid Biosynthesis as a Model System for Studying Metabolic Pathways

Arabidopsis has been developed as a model system for the study of plant genetics,

biochemistry and molecular biology because of its ease of growth, short generation time, utility

as a genetic system, and the availability of an efficient and simple transformation system (Page

and Grossniklaus, 2002). In our laboratory, Arabidopsis is being used to provide a better insight

into the flavonoid biosynthetic machinery and enzyme complexes as a whole. The Arabidopsis

flavonoid biosynthetic pathway is very well characterized and mutations in flavonoid genes can

be easily identified based on the seed coat color (Shirley et al., 1995). The use of the tt mutants

has provided important new knowledge on the function of flavonoids for the plant. For example,

the tt4(2YY6) mutant (which carries a null allele for CHS) was used to show that, unlike the

situation in maize and petunia, flavonoids do not play a role in male fertility in Arabidopsis

(Burbulis et al., 1996b). Another interesting result is the determination that sinapate esters and

flavonoids play a critical function as sunscreens (Landry et al., 1995; Li et al., 1993). These

mutants have also played an important part in characterizing the organization of the pathway as a

Page 27: Structural Characterization of the Flavonoid Enzyme Complex

12

multi-enzyme complex (Burbulis et al., 1996a; Burbulis and Winkel-Shirley, 1999; Saslowsky

and Winkel-Shirley, 2001).

Genes encoding eight flavonoid enzymes have been isolated in Arabidopsis. These

enzymes are CHS, CHI, F3H, flavonol synthase (FLS1-6), F3’H, DFR, ANS (also known as

leucoanthocyanidin dioxygenase (LDOX)), and ANR (also known as BANYULS (BAN))

(Burbulis et al., 1996a; Pelletier et al., 1999; Pelletier et al., 1997; Pelletier and Shirley, 1996;

Saslowsky and Winkel-Shirley, 2001; Shirley and Goodman, 1993; Xie de et al., 2004). All of

these enzymes are encoded by single copy genes with the exception of FLS, which is present in

six copies, at most three of which are thought to be catalytically active (Owens, Walton, and

Winkel, unpublished data). Five of these enzymes, CHS, CHI, F3H, FLS, ANS, and ANR, have

been overexpressed in E. coli and used to raise polyclonal antibodies (Cain et al., 1997; Pelletier

et al., 1997; Xie de et al., 2004).

Structural Biology of Flavonoid Metabolism

Understanding the structure and function of a protein on an atomic level can lead to a

better understanding the catalytic properties of a protein. Structures have been solved for three

flavonoid enzymes over the past several years, CHS2 and CHI from Medicago sativa, and ANS

from Arabidopsis (Ferrer et al., 1999; Jez et al., 2000b; Turnbull et al., 2001; Wilmouth et al.,

2002). As detailed below, the structures of these enzymes have allowed researchers to better

understand how these enzymes function catalytically and also investigate ways to alter their

substrate specificity and understand their evolution.

Page 28: Structural Characterization of the Flavonoid Enzyme Complex

13

Chalcone Synthase: CHS catalyzes the first committed reaction in flavonoid biosynthesis in

which one molecule of 4-coumaroyl-CoA and 3 molecules of malonyl-CoA are used to create

naringenin chalcone, the starter molecule for the ensuing steps in flavonoid metabolism (Figure

1.2). CHS is a member of the plant type III polyketide synthase family, which encompasses all

CHS enzymes as well as a variety of ‘CHS-like’ enzymes that catalyze similar polyketide

reactions in plants (Austin and Noel, 2003). The structure (Figure 1.3) and the reaction

mechanism of CHS2 from Medicago sativa was solved in the laboratory of Joe Noel at the Salk

Institute (Ferrer et al., 1999; Jez et al., 2000a; Jez et al., 2001; Jez et al., 2000d; Jez and Noel,

2000). The CHS structure is made up of an aßaßa core motif that is conserved among all type

III polyketide synthases as well as enzymes containing a thiolase fold, such as ketoacyl thiolase

(Austin and Noel, 2003). CHS bears a striking resemblance, both in sequence and structure, to

thiolase enzymes such as ketoacyl synthases and thiolases, which provides evidence for the

hypothesis that CHS evolved from an enzyme of fatty acid biosynthesis as first suggested by

Verwoert, et al. (1992).

Catalytically-active CHS is a bi- lobed homodimer containing two distinct active sites.

The CHS homodimer is held together by identical six-residue loops in each opposing monomer

that contribute a methionine to the wall of the active site of the neighboring subunit. The CHS

active site cavity is buried within the structure of the

protein and is accessible only through a 16 ? long channel. However, the CHS crystal structure

has shown that this channel is actually too small for the product to leave CHS, suggesting that it

is a dynamic structure that can change dimensions to accommodate the entry or exit of different

substrates and products (Ferrer et al., 1999).

Page 29: Structural Characterization of the Flavonoid Enzyme Complex

14

The CHS reaction starts by the loading of coumaroyl-CoA onto the active site cysteine

(Fig. 4), which is conserved among all thiolase-fold enzymes (Austin and Noel, 2003). This

allows the CoA moiety to be released through a decarboxylation reaction. Malonyl-CoA is then

freed of its CoA moiety by a second decarboxylation reaction and attached to the coumarate

bound in the active site. Two more malonyl-CoA units are added until a linear tetraketide chain

is created, which is circularized to form the chalcone product through an aromatase- like

mechanism that is not yet completely understood. The overall reaction is carried out by a

‘catalytic triad’ of conserved residues that play critical roles in the decarboxylation and

condensation reactions required for catalysis (Austin and Noel, 2003). The first of these

residues, the active site cysteine, was initially identified

by site directed mutagenesis studies performed in Gerhard Schröder’s group (Lanz et al., 1991).

This cysteine is conserved across all thiolase-fold enzymes and situated in a loop between two

buried amphipathic a-helices within the central core of the enzyme (Austin and Noel, 2003).

Chalcone Isomerase: CHI catalyzes the second step in flavonoid biosynthesis, closing the

center ring (Figure 1.5) and creating the basic flavonoid backbone. The CHI reaction can occur

spontaneously, however there is a 107 increase in reaction rate when the reaction is performed by

CHI (Bednar and Hadcock, 1988). There is also some evidence that CHI is post-translationally

modified, possibly through a thiol linkage, which could anchor CHI to the ER or aid in the

association of CHI with other enzymes in the complex (Burbulis and Winkel-Shirley, 1999).

The structure (Figure 1.6) and catalytic mechanism of CHI from Medicago sativa were

also solved in Joe Noel’s laboratory (Jez et al., 2000c; Jez et al., 2002). The structure resembles

an upside down flower bouquet with an open ß sandwich fold (Jez et al., 2000c). The overall

Page 30: Structural Characterization of the Flavonoid Enzyme Complex

15

CHI fold is maintained in this enzyme across higher plant species. The fold is also predicted to

be present in some bacterial (gamma proteobacteria) and fungal proteins (Gensheimer and

Mushegian, 2004), dispelling the initial notion that this fold is unique to CHI. The functions of

these ‘CHI- like’ proteins are unknown, however.

The cyclization reaction catalyzed by CHI is facilitated by an initial binding of the

tetrahydroxychalcone through a hydrogen bond between the hydroxyl group at the 7 position of

the substrate to the hydroxyl on threonine 190, which has also been suggested to function in

determining substrate preference (Jez et al., 2002). A further hydrogen bonding network is

required in order to catalyze the reaction. One water molecule, in particular, is hydrogen bonded

to tyrosine 106, which is thought to activate the water molecule and is hydrogen bonded to the

substrate. Upon activation, the water molecule attacks the carbon-carbon double bond, thus

reducing it and allowing for the formation of a bond to close the center ring on the substrate (Jez

et al., 2002). CHI appears to have evolved to be stereospecific in the reaction it catalyzes.

Medicago CHI prefers the S-isomer of tetrahydroxychalcone over the R- isomer by an apparent

100,000:1 ratio (Bednar and Hadcock, 1988). In order for CHI to efficiently catalyze the

conversion of the R- isomer of tetrahydroxychalcone, there would have to be significant changes

in the active site structure in order to accommodate the additional stereochemistry (Jez et al.,

2000c). This stereospecificity raises the questions, does CHS always produce the S- isomer and,

if not, is there a use for the R- isomer in plant secondary metabolism? Unfortunately, the answers

are not yet known.

Page 31: Structural Characterization of the Flavonoid Enzyme Complex

16

Methods to Determine Interactions Between Proteins

When investigating interactions between proteins, it is helpful to have multiple

techniques available. Many of the original studies that demonstrated protein-protein interactions

relied on techniques such as precipitation of protein aggregates, chromatography, and

ultracentrifugation (Ovádi and Srere, 2000). More recent techniques such as fluorescence

resonance energy transfer (Day et al., 2001), affinity gel electrophoresis (Beeckmans et al.,

1989), and yeast two-hybrid analysis (Ito et al., 2001) allow for the identification of two or more

interacting proteins. Techniques such as surface plasmon resonance refractometry (Rich and

Myszka, 2000), analytical ultracentrifugation (Lebowitz et al., 2002), and isothermal titration

calorimetry (Weber and Salemme, 2003) are capable of giving more detailed information such as

kinetic binding constants, reaction kinetics, and the thermodynamics of interactions. However,

these techniques cannot approach the atomic level of resolution that is needed for determining

the structure of a complex. The best ways to gain an appreciation of the structure of an enzyme-

complex on an atomic level are through co-crystallography of the proteins, small angle neutron

scattering (SANS) (Koch et al., 2003), and in silico protein-protein docking algorithms (Halperin

et al., 2002). For the purpose of this introduction, two of the above techniques, SANS and

surface plasmon resonance refractometry, will be discussed.

Small-Angle Neutron Scattering: SANS is a technique that uses neutrons to probe three-

dimensional structure of biological macromolecules in a non-destructive manner. Whereas X-

ray diffraction is typically used to study atomic- level structures (under 10 Å), SANS is generally

used to investigate larger structures (10 to 1000 Å) (Koch et al., 2003). Briefly, a sample is

Page 32: Structural Characterization of the Flavonoid Enzyme Complex

17

exposed to a monochromatic, cold neutron beam generated by either a nuclear reactor or a

spallation source (details are shown in Fig. 7). The neutrons are scattered upon contact with the

sample due to interactions with the individual atoms. Neutron scattering intensities are recorded

by a two-dimensional detector, which can be moved in the both the y- and z- axis with respect to

the sample. Typically, neutrons are scattered through small angles (less than 1 degree) when

compared to traditional diffraction angles (more than 10 degrees). The scattering pattern is then

analyzed computationally to determine the overall size and shape of the molecules in solution.

The data obtained from these measurements do not provide a high resolution structure

like traditional crystallography, but rather a low resolution solution structure into which high

resolution crystal structures can be fit in order to develop models of macromolecular assemblies.

SANS can also be used to determine regions of proteins that are highly mobile in crystal

structures. SANS has been used in a number of biological applications, contributing to

determination of the structure of the 30S ribosomal subunit (Capel et al., 1987), elucidating

better ways to crystallize a protein by determining the amount of protein that is aggregated

(Velev et al., 1998), and characterizing the structure of the GroEL/ES chaperonin complex

(Krueger et al., 2003). Beyond biological applications, SANS has been used in the study of

polymers, complex fluids, materials science and condensed matter (Koch et al., 2003).

Surface Plasmon Resonance: Another powerful technique that can be used to determine

molecular interactions between proteins is SPR refractometry (or spectroscopy). This approach

uses monochromatic light to detect interactions between different molecules via a change in

refractive index that occurs during complex formation or dissociation (Rich and Myszka, 2000).

Briefly, protein is first immobilized onto a gold-coated glass slide that has been coupled with oil

Page 33: Structural Characterization of the Flavonoid Enzyme Complex

18

to a sapphire crystal through which monochromatic light is passed (Fig. 8). The slide (or chip)

has a defined surface chemistry through a thiol linkage onto the gold. Two of the more common

surface chemistries currently being used are carboxy-methyl dextran self assembling monolayers

and metal affinity chemistries (Rich and Myszka, 2000). The former involves covalently linking

the protein (typically, any surface amine groups are utilized) to the surface chemistry whereas

the latter generally involves non-covalent coupling via a hexa-histadine tag. Once protein is

bound onto the slide, unbound protein is washed off, and solution containing a second protein is

passed over the slide. Interaction between the two is detected by a change in refractive angle

which is detected by a photo diode array and reported on the monitor.

SPR can be used not only to detect protein interactions, but also to measure kinetics,

affinity constants, the thermodynamics of an interaction, and, in some cases, to determine the

stoichiometry and mechanism of interaction (Morton and Myszka, 1998; Myszka et al., 1998).

Recent advances have also coupled SPR with mass spectrometry to identify interacting proteins

as well as determining interaction domains (Nedelkov and Nelson, 2003a). Various software

packages exist to aid in analyzing the results of SPR experiments, such as CLAMP99 (Myszka

and Morton, 1998), which allows for the determination of association and dissociation values for

an interaction.

One of many examples of current research using SPR is with the human

immunodeficiency virus (HIV). Areas of the HIV life cycle that have been intensely studied

with SPR are binding, mapping of epitopes for antibody development, and effects of inhibitory

drugs on viral machinery. This work has given a unique insight into how HIV functions

(reviewed in Rich and Myszka, 2003). These studies focused on antibody-antigen interactions

Page 34: Structural Characterization of the Flavonoid Enzyme Complex

19

with the hopes of finding a magic bullet that will affect the ability of HIV to infect and replicate

in the human host.

An example of biosynthetic enzymes being studied by SPR is the cysteine synthase

complex. Previous evidence from yeast two-hybrid analysis indicated that serine

acetyltransferase (SAT) and O-acetylserine (thiol) lyase (OAS-TL) interact to form a complex

that is involved in cysteine biosynthesis (Bogdanova and Hell, 1997). These initial studies also

defined some of the interaction domains between SAT and OAS-TL by means of truncation

series. SPR was subsequently used to determine the binding kinetics of these two enzymes when

they are activated in the presence of sulfur (Berkowitz et al., 2002).

A recent innovation has been the addition of mass spectroscopy at the conclusion of the

SPR assay (Nedelkov and Nelson, 2001). The technique of SPR is carried out in the same

manner as in conventional applications, except that the immobilized protein has to be covalently

attached to the slide to inhibit leaching. Once the SPR experiment is complete, interacting

proteins are eluted, leaving the initial protein bound to the slide, and these proteins are subjected

to matrix-assisted laser desorption- ionization time-of-flight (MALDI-TOF) MS (Nedelkov and

Nelson, 2003b). This technique allows for the effective identification of previously-unknown

interacting proteins. An example of the application of SPR-MS is the identification of the

protein complex makeup of insulin- like growth factors (IGF-1 and -2) from human plasma

(Nedelkov et al., 2003). In separate experiments, antibodies against IGF-1 or -2 were used to

immobilize IGF-1 and -2 along with any other proteins interacting with them from human

plasma, which were later identified by MS. It was determined that both growth factors are

present in human plasma, suggesting that IGF-1, -2 and a previously unknown truncated form of

IGF-2 make up a protein complex in vivo (Nedelkov et al., 2003).

Page 35: Structural Characterization of the Flavonoid Enzyme Complex

20

Project Goals/Rationale

There has been a variety of experimental evidence for interactions between enzymes

involved in flavonoid metabolism, both in vitro and in vivo. These data provide a broad

understanding of interactions within the metabolon, but no details regarding the molecular basis

of these interactions. It would seem that the next logical step would be to investigate this

problem on an atomic scale, i.e., determining how the enzymes physically interact with each

other and identifying the specific residues and atoms involved. Data from such experiments

would provide more information, not only on the organization of the flavonoid metabolon, but

also for metabolic organization in general. These insights could be used to create stronger (or

weaker) interactions to modulate pathway flux and engineer plants with higher levels of

flavonoids important for flower pigments, (e.g., create a blue rose), or as antioxidants.

This project was divided into two major objectives. First, the structural implications of

amino acid substitutions were investigated by molecular dynamics simulations. The second part

of this project studied the functional and structural components of flavonoid biosynthesis,

specifically the interactions between CHS, CHI, and F3H. X-ray crystallography and molecular

modeling were utilized to determine or deduce the structure of proteins involved in flavonoid

biosynthesis. Amino acid substitutions were introduced on the surface of CHS to test for effects

on the interaction of CHS with CHI in a yeast two-hybrid assay. SANS, which had not

previously been used for the study of metabolic complexes, was used to generate a structural

model of the interacting proteins. Next, computational docking studies were performed to

develop a model that fit the SANS data. The results from this project provide new insights into

Page 36: Structural Characterization of the Flavonoid Enzyme Complex

21

the self assembly of enzymes involved in flavonoid biosynthesis, with implications for the

organization of enzyme complexes in general.

Page 37: Structural Characterization of the Flavonoid Enzyme Complex

22

References Cited:

Austin, M. B., and Noel, J. P. (2003). The chalcone synthase superfamily of type III polyketide

synthases. Nat Prod Rep 20, 79-110.

Bednar, R. A., and Hadcock, J. R. (1988). Purification and characterization of chalcone

isomerase from soybeans. J Biol Chem 263, 9582-9588.

Beeckmans, S., Khan, A. S., Van Driessche, E., and Kanarek, L. (1994). A specific association

between the glyoxylic-acid-cycle enzymes isocitrate lyase and malate synthase. Eur J

Biochem 224, 197-201.

Beeckmans, S., Van Driessche, E., and Kanarek, L. (1989). The visualization by affinity

electrophoresis of a specific association between the consecutive citric acid cycle

enzymes fumarase and malate dehydrogenase. Eur J Biochem 183, 449-454.

Berkowitz, O., Wirtz, M., Wolf, A., Kuhlmann, J., and Hell, R. (2002). Use of biomolecular

interaction analysis to elucidate the regulatory mechanism of the cysteine synthase

complex from Arabidopsis thaliana. J Biol Chem 277, 30629-30634.

Bogdanova, N., and Hell, R. (1997). Cysteine synthesis in plants: protein-protein interactions of

serine acetyltransferase from Arabidopsis thaliana. Plant J 11, 251-262.

Brown, D. E., Rashotte, A. M., Murphy, A. S., Normanly, J., Tague, B. W., Peer, W. A., Taiz,

L., and Muday, G. K. (2001). Flavonoids act as negative regulators of auxin transport in

vivo in Arabidopsis thaliana. Plant Physiol 126, 524-535.

Buer, C. S., and Muday, G. K. (2004). The transparent testa4 mutation prevents flavonoid

synthesis and alters auxin transport and the response of Arabidopsis roots to gravity and

light. Plant Cell 16, 1191-1205.

Page 38: Structural Characterization of the Flavonoid Enzyme Complex

23

Burbulis, I. E., Iacobucci, M., and Shirley, B. W. (1996a). A null mutation in the first enzyme of

flavonoid biosynthesis does not affect male fertility in Arabidopsis. Plant Cell 8, 1013-

1025.

Burbulis, I. E., Pelletier, M. K., Cain, C. C., and Shirley, B. W. (1996b). Are flavonoids

synthesized by a multi-enzyme complex? So Assoc Agricult Sci Bull 9, 29-36.

Burbulis, I. E., and Winkel-Shirley, B. (1999). Interactions among enzymes of the Arabidopsis

flavonoid biosynthetic pathway. Proc Natl Acad Sci U S A 96, 12929-12934.

Cain, C., Saslowsky, D., Walker, R., and Shirley, B. (1997). Expression of chalcone synthase

and chalcone isomerase proteins in Arabidopsis seedlings. Plant Mol Biol 35, 377-381.

Capel, M. S., Engelman, D. M., Freeborn, B. R., Kjeldgaard, M., Langer, J. A., Ramakrishnan,

V., Schindler, D. G., Schneider, D. K., Schoenborn, B. P., Sillers, I. Y., and et al. (1987).

A complete mapping of the proteins in the small ribosomal subunit of Escherichia coli.

Science 238, 1403-1406.

Chapple, C. C., Vogt, T., Ellis, B. E., and Somerville, C. R. (1992). An Arabidopsis mutant

defective in the general phenylpropanoid pathway. Plant Cell 4, 1413-1424.

Czichi, U., and Kindl, H. (1977). Phenylalanine ammonia- lyase and cinnamic acid hydroxylase

as assembled consecutive enzymes on microsomal membranes of cucumber cotyledons:

cooperation and subcellular distribution. Planta 134, 133-143.

Day, R. N., Periasamy, A., and Schaufele, F. (2001). Fluorescence resonance energy transfer

microscopy of localized protein interactions in the living cell nucleus. Methods 25, 4-18.

Day, T. A. (1993). Relating UV-B radiation screening effectiveness of foliage to absorbing-

compound concentration and anatomical characteristics in a diverse group of plants.

Oecologia 95, 542-550.

Page 39: Structural Characterization of the Flavonoid Enzyme Complex

24

Dixon, R. A., and Steele, C. L. (1999). Flavonoids and isoflavonoids - a gold mine for metabolic

engineering. Trends Plant Sci 4, 394-400.

Ferrer, J.-L., Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (1999). Structure of

chalcone synthase and the molecular basis of plant polyketide biosynthesis. Nat Struc

Biol 6, 775-784.

Forkmann, G., and Martens, S. (2001). Metabolic engineering and applications of flavonoids.

Curr Opinion Biotechnol 12, 155-160.

Fritsch, H., and Grisebach, H. (1975). Biosynthesis of cyanidin in cell cultures of Haplopappus

gracilis. Phytochemistry 14, 2437-2442.

Gensheimer, M., and Mushegian, A. (2004). Chalcone isomerase family and fold: no longer

unique to plants. Protein Sci 13, 540-544.

Gontero, B., Cárdenas, M. L., and Ricard, J. (1988). A functional five-enzyme complex of

chloroplasts involved in the Calvin cycle. Eur J Biochem 173, 437-443.

Goodsell, D. S. (1991). Inside a living cell. Trends Biochem Sci 16, 203-206.

Green, D. E. (1958). Studies in organized enzyme systems. Harvey Lect 52, 177-227.

Guex, N., and Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an

environment for comparative protein modeling. Electrophoresis 18, 2714-2723.

Halperin, I., Ma, B., Wolfson, H., and Nussinov, R. (2002). Principles of docking: An overview

of search algorithms and a guide to scoring functions. Proteins 47, 409-443.

Havsteen, B. H. (2002). The biochemistry and medical significance of the flavonoids. Pharmacol

Ther 96, 67-202.

Hirsch, A. M., Lum, M. R., and Downie, J. A. (2001). What makes the rhizobia- legume

symbiosis so special? Plant Physiol 127, 1484-1492.

Page 40: Structural Characterization of the Flavonoid Enzyme Complex

25

Hoppe, W., Gassmann, J., Hunsmann, N., Schramm, H. J., and Sturm, M. (1974). Three-

dimensional reconstruction of individual negatively stained yeast fatty-acid synthetase

molecules from tilt series in the electron microscope. Hoppe Seylers Z Physiol Chem

355, 1483-1487.

Hrazdina, G. (1992). Compartmentation in Aromatic Metabolism. In Phenolic Metabolism in

Plants, H. A. Stafford, ed.

Hrazdina, G., and Jensen, R. A. (1992). Spatial organization of enzymes in plant metabolic

pathways. Annu Rev Plant Physiol Plant Mol Biol 43, 241-267.

Hrazdina, G., and Wagner, G. J. (1985). Metabolic pathways as enzyme complexes: evidence for

the synthesis of phenylpropanoids and flavonoids on membrane associated enzyme

complexes. Arch Biochem Biophys 237, 88-100.

Ito, T., Chiba, T., Ozawa, R., Yoshida, M., Hattori, M., and Sakaki, Y. (2001). A comprehensive

two-hybrid analysis to explore the yeast protein interactome. Proc Natl Acad Sci U S A

98, 4569-4574.

Jez, J. M., Austin, M. B., Ferrer, J., Bowman, M. E., Schroder, J., and Noel, J. P. (2000a).

Structural control of polyketide formation in plant-specific polyketide synthases. Chem

Biol 7, 919-930.

Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (2000b). Structure and mechanism of

chalcone isomerase: an evolutionarily unique enzyme in plants. Nature Struct Biol 7,

786-791.

Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (2000c). Structure and mechanism of

the evolutionarily unique plant enzyme chalcone isomerase. Nat Struct Biol 7, 786-791.

Page 41: Structural Characterization of the Flavonoid Enzyme Complex

26

Jez, J. M., Bowman, M. E., and Noel, J. P. (2002). Role of hydrogen bonds in the reaction

mechanism of chalcone isomerase. Biochemistry 41, 5168-5176.

Jez, J. M., Ferrer, J. L., Bowman, M. E., Austin, M. B., Schroder, J., Dixon, R. A., and Noel, J.

P. (2001). Structure and mechanism of chalcone synthase-like polyketide synthases. J Ind

Microbiol Biotechnol 27, 393-398.

Jez, J. M., Ferrer, J. L., Bowman, M. E., Dixon, R. A., and Noel, J. P. (2000d). Dissection of

malonyl-coenzyme A decarboxylation from polyketide formation in the reaction

mechanism of a plant polyketide synthase. Biochemistry 39, 890-902.

Jez, J. M., and Noel, J. P. (2000). Mechanism of chalcone synthase. pKa of the catalytic cysteine

and the role of the conserved histidine in a plant polyketide synthase. J Biol Chem 275,

39640-39646.

Johnson, C. S., Kolevski, B., and Smyth, D. R. (2002). TRANSPARENT TESTA GLABRA2, a

trichome and seed coat development gene of Arabidopsis, encodes a WRKY transcription

factor. Plant Cell 14, 1359-1375.

Jordan, B. R., James, P. E., and S, A. H.-M. (1998). Factors affecting UV-B-induced changes in

Arabidopsis thaliana L. gene expression: the role of development, protective pigments

and the chloroplast signal. Plant Cell Physiol 39, 769-778.

Kempner, E. S., and Miller, J. H. (1968). The molecular biology of Euglena gracilis. V. Enzyme

localization. Exp Cell Res 51, 150-156.

Keung, W. M. (2003). Anti-dipsotropic isoflavones: the potential therapeutic agents for alcohol

dependence. Med Res Rev 23, 669-696.

Page 42: Structural Characterization of the Flavonoid Enzyme Complex

27

Koch, M. H., Vachette, P., and Svergun, D. I. (2003). Small-angle scattering: a view on the

properties, structures and structural changes of biological macromolecules in solution. Q

Rev Biophys 36, 147-227.

Krueger, S., Gregurick, S. K., Zondlo, J., and Eisenstein, E. (2003). Interaction of GroEL and

GroEL/GroES complexes with a nonnative subtilisin variant: a small-angle neutron

scattering study. J Struct Biol 141, 240-258.

Lahiri, J., Isaacs, L., Tien, J., and Whitesides, G. M. (1999). A strategy for the generation of

surfaces presenting ligands for studies of binding based on an active ester as a common

reactive intermediate: a surface plasmon resonance study. Anal Chem 71, 777-790.

Landry, L. G., Chapple, C. C. S., and Last, R. L. (1995). Arabidopsis mutants lacking phenolic

sunscreens exhibit enhanced ultraviolet-B injury and oxidative damage. Plant Physiol

109, 1159-1166.

Lanz, T., Tropf, S., Marner, F. J., Schroder, J., and Schroder, G. (1991). The role of cysteines in

polyketide synthases. Site-directed mutagenesis of resveratrol and chalcone synthases,

two key enzymes in different plant-specific pathways. J Biol Chem 266, 9971-9976.

Lebowitz, J., Lewis, M. S., and Schuck, P. (2002). Modern analytical ultracentrifugation in

protein science: a tutorial review. Protein Sci 11, 2067-2079.

Li, J., Ou-Lee, T.-M., Raba, R., Amundson, R. G., and Last, R. L. (1993). Arabidopsis flavonoid

mutants are hypersensitive to UV-B irradiation. Plant Cell 5, 171-179.

Li, S., Armstrong, C. M., Bertin, N., Ge, H., Milstein, S., Boxem, M., Vidalain, P. O., Han, J. D.,

Chesneau, A., Hao, T., et al. (2004). A map of the interactome network of the metazoan

C. elegans. Science 303, 540-543.

Page 43: Structural Characterization of the Flavonoid Enzyme Complex

28

Liu, C.-J., and Dixon, R. A. (2001). Elicitor-induced association of isoflavone O-

methyltransferase with endomembranes prevents the formation and 7-O-methylation of

daizein during isoflavoniod phytoalexin biosynthesis. Plant Cell 13, 2643-2658.

Lotito, S. B., and Fraga, C. G. (2000). Catechins delay lipid oxidation and alpha-tocopherol and

beta-carotene depletion following ascorbate depletion in human plasma. Proc Soc Exp

Biol Med 225, 32-38.

Marles, M. A., Ray, H., and Gruber, M. Y. (2003). New perspectives on proanthocyanidin

biochemistry and molecular regulation. Phytochemistry 64, 367-383.

Marsh, B. J., Mastronarde, D. N., Buttle, K. F., Howell, K. E., and McIntosh, J. R. (2001).

Organellar relationships in the Golgi region of the pancreatic beta cell line, HIT-T15,

visualized by high resolution electron tomography. Proc Natl Acad Sci U S A 98, 2399-

2406.

Martin, C., and Gerats, T. (1993). Control of pigment biosynthesis genes during petal

development. Plant Cell 5, 1253-1264.

Massant, J., and Glansdorff, N. (2004). Metabolic channelling of carbamoyl phosphate in the

hyperthermophilic archaeon Pyrococcus furiosus: dynamic enzyme-enzyme interactions

involved in the formation of the channelling complex. Biochem Soc Trans 32, 306-309.

Mathews, C. K. (1993). The cell-bag of enzymes or network of channels? J Bacteriol 175, 6377-

6381.

Medalia, O., Weber, I., Frangakis, A. S., Nicastro, D., Gerisch, G., and Baumeister, W. (2002).

Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography.

Science 298, 1209-1213.

Page 44: Structural Characterization of the Flavonoid Enzyme Complex

29

Meyer, P., Heidmann, I., Forkmann, G., and Saedler, H. (1987). A new petunia flower color

generated by transformation of a mutant with a maize gene. Nature 330, 677-678.

Morton, T. A., and Myszka, D. G. (1998). Kinetic analysis of macromolecular interactions using

surface plasmon resonance biosensors. Methods Enzymol 295, 268-294.

Muir, S. R., Collins, G. J., Robinson, S., Hughes, S., Bovy, A., Ric De Vos, C. H., van Tunen, A.

J., and Verhoeyen, M. E. (2001). Overexpression of petunia chalcone isomerase in

tomato results in fruit containing increased levels of flavonols. Nat Biotechnol 19, 470-

474.

Murphy, A., Peer, W. A., and Taiz, L. (2000). Regulation of auxin transport by aminopeptidases

and endogenous flavonoids. Planta 211, 315-324.

Myszka, D. G., Jonsen, M. D., and Graves, B. J. (1998). Equilibrium analysis of high affinity

interactions using BIACORE. Anal Biochem 265, 326-330.

Myszka, D. G., and Morton, T. A. (1998). CLAMP: a biosensor kinetic data analysis program.

Trends Biochem Sci 23, 149-150.

Nedelkov, D., and Nelson, R. W. (2001). Delineation of in vivo assembled multiprotein

complexes via biomolecular interaction analysis mass spectrometry. Proteomics 1, 1441-

1446.

Nedelkov, D., and Nelson, R. W. (2003a). Delineating protein-protein interactions via

biomolecular interaction analysis-mass spectrometry. J Mol Recognit 16, 9-14.

Nedelkov, D., and Nelson, R. W. (2003b). Surface plasmon resonance mass spectrometry: recent

progress and outlooks. Trends Biotechnol 21, 301-305.

Page 45: Structural Characterization of the Flavonoid Enzyme Complex

30

Nedelkov, D., Nelson, R. W., Kiernan, U. A., Niederkofler, E. E., and Tubbs, K. A. (2003).

Detection of bound and free IGF-1 and IGF-2 in human plasma via biomolecular

interaction analysis mass spectrometry. FEBS Lett 536, 130-134.

Nesi, N., Debeaujon, I., Jond, C., Pelletier, G., Caboche, M., and Lepiniec, L. (2000). The TT8

gene encodes a basic helix- loop-helix domain protein required for expression of DFR and

BAN genes in Arabidopsis siliques. Plant Cell 12, 1863-1878.

Nesi, N., Debeaujon, I., Jond, C., Stewart, A. J., Jenkins, G. I., Caboche, M., and Lepiniec, L.

(2002). The TRANSPARENT TESTA16 locus encodes the ARABIDOPSIS BSISTER

MADS domain protein and is required for proper development and pigmentation of the

seed coat. Plant Cell 14, 2463-2479.

Nesi, N., Jond, C., Debeaujon, I., Caboche, M., and Lepiniec, L. (2001). The Arabidopsis TT2

gene encodes an R2R3 MYB domain protein that acts as a key determinant for

proanthocyanidin accumulation in developing seed. Plant Cell 13, 2099-2114.

Ovádi, J., and Srere, P. A. (2000). Macromolecular compartmentation and channeling. Int Rev

Cytol 192, 255-280.

Page, D. R., and Grossniklaus, U. (2002). The art and design of genetic screens: Arabidopsis

thaliana. Nat Rev Genet 3, 124-136.

Paiva, N. L. (2000). An Introduction to the Biosynthesis of Chemicals Used in Plant-Microbe

Communication. Journal of Plant Growth Regulation 19, 131-143.

Pelletier, M. K., Burbulis, I. E., and Winkel-Shirley, B. (1999). Disruption of specific flavonoid

genes enhances the accumulation of flavonoid enzymes and end-products in Arabidopsis

seedlings. Plant Mol Biol 40, 45-54.

Page 46: Structural Characterization of the Flavonoid Enzyme Complex

31

Pelletier, M. K., Murrell, J. R., and Shirley, B. W. (1997). Characterization of flavonol synthase

and leucoanthocyanidin dioxygenase genes in Arabidopsis. Plant Physiol 113, 1437-

1445.

Pelletier, M. K., and Shirley, B. W. (1996). Analysis of flavanone 3-hydroxylase in Arabidopsis

seedlings. Coordinate regulation with chalcone synthase and chalcone isomerase. Plant

Physiol 111, 339-345.

Rasmussen, S., and Dixon, R. A. (1999). Transgene-mediated and elicitor- induced perturbation

of metabolic channeling at the entry point into the phenylpropanoid pathway. Plant Cell

11, 1537-1551.

Rich, R. L., and Myszka, D. G. (2000). Advances in surface plasmon resonance biosensor

analysis. Curr Opin Biotechnol 11, 54-61.

Rich, R. L., and Myszka, D. G. (2003). Spying on HIV with SPR. Trends Microbiol 11, 124-133.

Robinson, J. B. J., Inman, L., Semegi, B., and Srere, P. A. (1987). Further characterization of the

Krebs tricarboxylic acid cycle metabolon. J Biol Chem 262, 1786-1790.

Sainz, M. B., Grotewold, E., and Chandler, V. L. (1997). Evidence for direct activation of an

anthocyanin promoter by the maize C1 protein and comparison of DNA binding by

related Myb domain proteins. Plant Cell 9, 611-625.

Sanchez, C., Lachaize, C., Janody, F., Bellon, B., Roder, L., Euzenat, J., Rechenmann, F., and

Jacq, B. (1999). Grasping at molecular interactions and genetic networks in Drosophila

melanogaster using FlyNets, an Internet database. Nucleic Acids Res 27, 89-94.

Santos, M. O., Crosby, W. L., and Winkel-Shirley, B. (2004). Modulation of flavonoid

metabolism in Arabidopsis using a phage-derived antibody. Molecular Breeding, in

press.

Page 47: Structural Characterization of the Flavonoid Enzyme Complex

32

Sarkar, F. H., and Li, Y. (2003). Soy isoflavones and cancer prevention. Cancer Invest 21, 744-

757.

Saslowsky, D., and Winkel-Shirley, B. (2001). Localization of flavonoid enzymes in Arabidopsis

roots. Plant J 27, 37-48.

Serafini, M., Bugianesi, R., Maiani, G., Valtuena, S., De Santis, S., and Crozier, A. (2003).

Plasma antioxidants from chocolate. Nature 424, 1013.

Shirley, B. W., and Goodman, H. M. (1993). An Arabidopsis gene homologous to mammalian

and insect genes encoding the largest proteasome subunit. Mol Gen Genet 241, 586-594.

Shirley, B. W., Kubasek, W. L., Storz, G., Bruggemann, E., Koornneef, M., Ausubel, F. M., and

Goodman, H. M. (1995). Analysis of Arabidopsis mutants deficient in flavonoid

biosynthesis. Plant J 8, 659-671.

Solovchenko, A., and Schmitz-Eiberger, M. (2003). Significance of skin flavonoids for UV-B-

protection in apple fruits. J Exp Bot 54, 1977-1984.

Spelt, C., Quattrocchio, F., Mol, J. N. M., and Koes, R. (2000). anthocyanin1 of petunia encodes

a basic helix- loop-helix protein that directly activates transcription of structural

anthocyanin genes. Plant Cell 12, 1619–1631.

Srere, P. A. (1985). The metabolon. Trends Biochem Sci 10, 109-110.

Srere, P. A. (1987). Complexes of sequential metabolic enzymes. Annu Rev Biochem 56, 89-

124.

Stafford, H. A. (1974a). The metabolism of aromatic compounds. Ann Rev Plant Physiol 25,

459-486.

Stafford, H. A. (1974b). Possible multi-enzyme complexes regulating the formation of C6-C3

phenolic compounds and lignins in higher plants. Rec Adv Phytochem 8, 53-79.

Page 48: Structural Characterization of the Flavonoid Enzyme Complex

33

Stafford, H. A. (1990). Flavonoid Metabolism (Boca Raton, CRC Press).

Strack, D., Pieroth, M., Scharf, H., and Sharma, V. (1985). Tissue distribution of

phenylpropanoid metabolism in cotyledons of Raphanus sativus. Planta 164, 507-511.

Su, V., and Hsu, B. D. (2003). Cloning and expression of a putative cytochrome P450 gene that

influences the colour of Phalaenopsis flowers. Biotechnol Lett 25, 1933-1939.

Sumegi, B., Sherry, A. D., Malloy, C. R., and Srere, P. A. (1993). Evidence for orientation-

conserved transfer in the TCA cycle in Saccharomyces cerevisiae: 13C NMR studies.

Biochem 32, 12725-12729.

Taylor, L. P., and Briggs, W. R. (1990). Genetic regulation and photocontrol of anthocyanin

accumulation in maize seedlings. Plant Cell 2, 115-127.

Turnbull, J. J., Prescott, A. G., Schofield, C. J., and Wilmouth, R. C. (2001). Purification,

crystallization and preliminary X-ray diffraction of anthocyanidin synthase from

Arabidopsis thaliana. Acta Crystallogr D 57, 425-427.

Velev, O. D., Kaler, E. W., and Lenhoff, A. M. (1998). Protein interactions in solution

characterized by light and neutron scattering: comparison of lysozyme and

chymotrypsinogen. Biophys J 75, 2682-2697.

Vélot, C., Mixon, M. B., Teige, M., and Srere, P. A. (1997). Model of a quinary structure

between Krebs TCA cycle enzymes: a model for the metabolon. Biochemistry 36, 14271-

14276.

Vélot, C., and Srere, P. A. (2000). Reversible transdominant inhibition of a metabolic pathway. J

Biol Chem 275, 12926-12933.

Verwoert, II, Verbree, E. C., van der Linden, K. H., Nijkamp, H. J., and Stuitje, A. R. (1992).

Cloning, nucleotide sequence, and expression of the Escherichia coli fabD gene,

Page 49: Structural Characterization of the Flavonoid Enzyme Complex

34

encoding malonyl coenzyme A-acyl carrier protein transacylase. J Bacteriol 174, 2851-

2857.

Wade, H. K., Sohal, A. K., and Jenkins, G. I. (2003). Arabidopsis ICX1 is a negative regulator of

several pathways regulating flavonoid biosynthesis genes. Plant Physiol 131, 707-715.

Wagner, G. J., and Hrazdina, G. (1984). Endoplasmic reticulum as a site of phenylpropanoid and

flavonoid metabolism in Hippeastrum. Plant Physiol 74, 901-906.

Walker, A. R., Davison, P. A., Bolognesi-Winfield, A. C., James, C. M., Srinivasan, N.,

Blundell, T. L., Esch, J. J., Marks, M. D., and Gray, J. C. (1999). The TRANSPARENT

TESTA GLABRA1 locus, which regulates trichome differentiation and anthocyanin

biosynthesis in Arabidopsis, encodes a WD40 repeat protein. Plant Cell 11, 1337-1350.

Weber, P. C., and Salemme, F. R. (2003). Applications of calorimetric methods to drug

discovery and the study of protein interactions. Curr Opin Struct Biol 13, 115-121.

Weisshaar, B., and Jenkins, G. I. (1998). Phenylpropanoid biosynthesis and its regulation. Curr

Opin Plant Biol 1, 251-257.

Welch, G. R. (1977). On the role of organized multienzyme systems in cellular metabolism: a

general synthesis. Prog Biophys Mol Biol 32, 103-191.

Wilmouth, R. C., Turnbull, J. J., Welford, R. W., Clifton, I. J., Prescott, A. G., and Schofield, C.

J. (2002). Structure and mechanism of anthocyanidin synthase from Arabidopsis

thaliana. Structure (Camb) 10, 93-103.

Winkel, B. S. J. (2004). Metabolic Channeling in Plants. Annu Rev Plant Biol 55, 85-107.

Winkel-Shirley, B. (1999). Evidence for enzyme complexes in the phenylpropanoid and

flavonoid pathways. Physiol Plant 107, 142-149.

Page 50: Structural Characterization of the Flavonoid Enzyme Complex

35

Winkel-Shirley, B. (2001). It takes a garden. How work on diverse plant species has contributed

to an understanding of flavonoid metabolism. Plant Physiol 127, 1399-1404.

Winkel-Shirley, B. (2002). Biosynthesis of flavonoids and effects of stress. Curr Opin Plant Biol

5, 218-223.

Xie de, Y., Sharma, S. B., and Dixon, R. A. (2004). Anthocyanidin reductases from Medicago

truncatula and Arabidopsis thaliana. Arch Biochem Biophys 422, 91-102.

Yu, O., Shi, J., Hession, A. O., Maxwell, C. A., McGonigle, B., and Odell, J. T. (2003).

Metabolic engineering to increase isoflavone biosynthesis in soybean seed.

Phytochemistry 63, 753-763.

Zalokar, M. (1960). Cytochemistry of centrifuged hyphae of Neurospora. Exp Cell Res 19, 114-

132.

Page 51: Structural Characterization of the Flavonoid Enzyme Complex

36

COSCoA

OHOH

COOH

NH2

Fatty Acid Metabolism

phenylalanine 4-coumaroyl-CoA

General Phenylpropanoid Pathway

3

Malonyl Co-A

+

CHS

OH

O

OHOH

OH

tetrahydroxychalcone

OH

O

OOH

OH

naringenin

CHI

OH

O

OOH

OH

OH

F3’H

F3H F3H OH

O

OOH

OH

OH

OH

O

OOH

OH

OH

OH

eriodictyol

DFR

F3’H

OH

OH

OOH

OH

OH

OH

OH

O+OH

OH

OH

OH

ANS

Anthocyanins

3GT, 5GT

ANR (BAN)

Condensed Tannins (proanthocyanidins)

OH

OOH

OH

OH

OH

LAR

OH

OOH

OH

OH

R

O

FLS

FLS

Flavonol glycosides

dihydrokaempferol dihydroquercetin

leucopelargonidin

leucocyanidin

cyanidin

Figure 1.1: Schematic of the flavonoid pathway in Arabidopsis thaliana . The pathway creates a diverse set of compounds used in UV protection, signaling, defense, and pigmentation. CHS represents that first committed step in the pathway, where 4-coumaroyl-CoA (from phenylpropanoid biosynthesis) and 3 units of malonyl-CoA (from fatty acid metabolism) are used as substrates. Structures in the figure represent the three enzymes in the pathway that crystal structures have been solved for, CHS (pdb id:1BI5) and CHI (pdb id:1EYQ) (both from Medicago sativa), and anthocyanidin synthase (ANS) (pdb id:1GP6) (from Arabidopsis thaliana). Structures in orange are homology models of dihydroflavonol 4-reductase (DFR), flavanone 3-hydroxylase (F3H), flavonoid 3’ hydroxylase (F3’H), and flavonol synthase (FLS). Other enzyme names are abbreviated as follows: anthocyanidin reductase (ANR), cinnamate 4-hydroxylase (C4H), p-coumarate:CoA ligase (4CL), leucoanthocyanidin reductase (LAR), phenylalanine ammonia-lyase (PAL). Arrows labeled in grey indicate branches of the pathway not present in Arabidopsis.

HOOC

COSCoA

Isoflavonoids

CHS/CHR

PAL C4H 4CL

3GT, RT

Page 52: Structural Characterization of the Flavonoid Enzyme Complex

37

Figure 1.2. Reaction catalyzed by CHS. CHS utilizes 4-coumaroyl-CoA and 3 units of malonyl-CoA in a Claisen condensation reaction to create tetrahydroxychalcone. The atoms on tetrahydroxychalcone which are labeled in black come from 4-coumaroyl-CoA and the atoms in red come from malonyl-CoA.

Page 53: Structural Characterization of the Flavonoid Enzyme Complex

38

Figure 1.3. Structure of Medicago sativa CHS2 (Ferrer et al., 1999). The CHS protein is composed of two identical monomers (illustrated in purple and yellow). The red atoms indicate the methionine that is utilized as part of the active site wall by the partner monomer. The arrow indicates the substrate/product channel leading to the buried active site. The structure was generated using SwissPDB Viewer (Guex and Peitsch, 1997) and rendered using POV-Ray 3.1 (www.povray.org).

Page 54: Structural Characterization of the Flavonoid Enzyme Complex

39

H303

C164

N336

F215

F265

Figure 1.4. Close up of CHS2 active site (Ferrer et al., 1999). F215 and F265 are the ‘gatekeeper’ residues that are proposed to control the entry of substrate. H303 and N336 are involved with the binding of substrate and removal of the CoA moieties while C164 holds the substrate during the reaction.

Page 55: Structural Characterization of the Flavonoid Enzyme Complex

40

Figure 1.5. The reaction catalyzed by CHI. The C ring of tetrahydroxychalcone is closed in a cyclization reaction to create (2S)-naringenin, which serves as the basic backbone for the downstream flavonoid metabolic reactions.

Page 56: Structural Characterization of the Flavonoid Enzyme Complex

41

Figure 1.6. The structure of CHI from Medicago sativa crystallized with its product (Jez et al., 2000c). The structure maintains an overall open ß-sandwich fold. The arrow indicates the location of the product, (2S)-Naringenin.

(2S)-Naringenin

Page 57: Structural Characterization of the Flavonoid Enzyme Complex

42

Neutron Source (Reactor/ Spallation) with LH2 moderator

Velocity Selector

Lens/focusing system

sample

detector

beamstop

Figure 1.7. Schematic representation of SANS. Cold neutrons are emitted from the neutron reactor or spallation source and are then selected for the appropriate wavelength by the velocity selector. A series of lenses then focus the monochromatic neutron beam before passage through the sample. Upon hitting the sample, scattered neutrons are recorded by the 2-D detector and turned into electrical impulses, which are then transmitted to a data collection computer. Non-scattered neutrons are blocked by a beamstop made of neutron-dense material. (adapted from Koch et al., 2003).

Page 58: Structural Characterization of the Flavonoid Enzyme Complex

43

?

Polarized Light Diode Array Detector

Coupling Prism

Solution of Protein

Flow Cell

Glass (0.2 mm) Ti (1.5 nm) Au (38 nm) m SAM (~ 3 nm)

Figure 1.8: Schematic representation of the principle of SPR. Polarized light at 760 nm is directed through the sapphire prism onto a glass slide with evaporated Ti and Au and a surface chemistry has been coupled. The change in observed resonance angle (?) is recorded by a photodetector when protein binds to the surface chemistry (adapted from Lahiri et al., 1999).

Page 59: Structural Characterization of the Flavonoid Enzyme Complex

44

Chapter 2

Interpretation of functional effects of mutations in Arabidopsis chalcone synthase based on

molecular modeling

Christopher D. Dana1, David R. Bevan2 and Brenda S.J. Winkel1

1Department of Biology and Fralin Biotechnology Center and 2Department of Biochemistry,

Virginia Tech

Manuscript to be submitted to Journal of Molecular Modeling

Key Words: molecular dynamics, allelic series, chalcone synthase (CHS), transparent testa (tt)

Page 60: Structural Characterization of the Flavonoid Enzyme Complex

45

Abstract

Chalcone synthase (CHS) catalyzes the first committed step in flavonoid biosynthesis, a

major pathway of plant secondary metabolism. An allelic series for the Arabidopsis CHS locus,

tt4, was previously characterized at the gene, protein, and end product levels (Saslowsky et al.,

2000). The substitutions identified in each of these alleles were used to generate models for five

of the tt4 proteins based on the Arabidopsis CHS crystal structure (Chapter 3) in order to

determine how mutations may affect protein structure and function. tt4(85) contains a

substitution that apparently disrupts catalytic activity due a change in the CoA binding face

geometry. The tt4(UV113) allele disrupts enzymatic activity while maintaining a stable protein,

which could be caused by a stiffening of the loop containing the critical active site residue, C169.

In contrast, the tt4(38G1R) mutation results in a truncated protein that is both catalytically

inactive, does not dimerize, and accumulates at reduced levels, presumably due to the loss of

three beta sheets. However, no structural basis could be discerned for the temperature-dependent

loss-of-function of the enzymes encoded by tt4(UV01) and tt4(UV25).

Page 61: Structural Characterization of the Flavonoid Enzyme Complex

46

Introduction

The type III polyketide synthases (PKSs) comprise a structurally-related class of enzymes

that are key to the production of an array of biologically-active natural products in plants and

some bacteria (Austin and Noel, 2003; Moore et al., 2002). These enzymes, also known as

homodimeric iterative PKSs, include the extensively-studied chalcone synthase (CHS), an

enzyme that is ubiquitous among plants and that catalyzes the first committed step in the

synthesis of flavonoids. In the CHS reaction, one molecule of p-coumaroyl CoA and three

molecules of malonyl CoA are used to generate a tetraketide that is then cyclized into 4,2’,4’,6’-

tetrahydroxychalcone. This molecule serves as a backbone for the production of a wide array of

flavonoid compounds that function in such diverse roles as flower pigmentation, UV protection,

signaling, male fertility, and defense against microbial pathogens (Winkel-Shirley, 2001).

Alfalfa CHS was the first type III PKS, and the first flavonoid enzyme, for which a

crystal structure was solved (Ferrer et al., 1999). Together with the subsequent determination of

structures for mutagenized forms of this enzyme and for 2-pyrone synthase (PS) from daisy (Jez

et al., 2002; Jez et al., 2000b), significant insights have been gained into the molecular basis of

the reaction series catalyzed by the type III PKSs. For example, it was known that, although

each of the two active sites in CHS and the closely related enzyme, stilbene synthase (STS),

function independently, dimerization is required for activity (Tropf et al., 1995). It is now clear

that this is because a methionine in each monomer helps shape the active site cavity of the

adjoining monomer. In addition, it has been shown that each active site is buried within the

enzyme and that substrates enter via a long (16 Å in alfalfa CHS) CoA binding tunnel [Ferrer,

1999 #79]. The three key catalytic residues are located within an initiation/elongation cavity,

one lobe of which forms a coumaroyl binding pocket, while the other accommodates and

Page 62: Structural Characterization of the Flavonoid Enzyme Complex

47

determines the length of the growing polyketide chain. Although CHS and PS use different

substrates, produce different products, and exhibit 74% amino acid sequence identity, the overall

architecture of the two enzymes is nearly identical. Remarkably, it has been possible to engineer

CHS for PS activity by altering just three residues that affect the overall dimensions of the

initiation/elongation cavity (Jez et al., 2002).

Several recent studies have utilized molecular dynamics simulations in efforts to

elucidate the functional properties of a protein containing amino acid substitutions. For example,

Ceruso et al. (2004) were able to suggest a nucleotide exchange mechanism in the Ga subunit of

transducin by modeling a series of substitutions in residues that had been previously

characterized biochemically. Another study, on the molecular switch present in the

glucocorticoid receptor, used molecular dynamics simulations to show that distinct substitutions

that result in a gain-of- function are caused by two different structural mechanisms (Stockner et

al., 2003).

We previously characterized an allelic series of mutations for Arabidopsis CHS, which is

defined by the TT4 locus (Saslowsky et al., 2000). The seven tt4 alleles included in this analysis

were found to have diverse effects on the accumulation of CHS mRNA, the dimerization and

stability of the enzyme, and the accumulation of flavonoid end products. Two of the alleles also

exhibited temperature sensitivity. A homology model was initially generated for the wild-type

Arabidopsis enzyme based on the Medicago CHS2 crystal structure, and the locations of the five

tt4 mutations resulting in single amino acid changes or a small deletion were mapped on this

model (Ferrer et al., 1999). This preliminary analysis suggested that several of the mutations

could have significant effects on the structure of the CHS enzyme. To further explore this

possibility, individual homology models have been generated for these five altered proteins

Page 63: Structural Characterization of the Flavonoid Enzyme Complex

48

based on the recently-solved crystal structure for Arabidopsis CHS (Chapter 3). These models

provide insights into the phenotypic and biochemical consequences that arise from either point

substitutions or truncation in the enzyme in three of these alleles.

Methods

A multiple sequence alignment of 15 representative members of the plant type III PKS

superfamily was generated using Clustal W (Thompson et al., 1994) using the Megalign program

in DNAStar (Madison, WI). Included in this analysis were CHS from Arabidopsis (accession

number P13114), alfalfa (P30074), maize (SYZMCC), parsley (S42523), raspberry

(AAK15174), snapdragon (P06515), and tea (P48387); acridone synthase 2 from rue (Q9FSC0);

bibenzyl synthase from orchid (S71619); coumaroyl triacetic acid synthase from hydrangea

(BAA32733); dihydropinosylvin-forming stilbene synthase from pine (Q02323); PS from daisy

(P48391); trihydroxystilbene synthase sequences from grape (P51070) and peanut (S00334); and

valerophenone synthase (O80400) from hops.

Homology models were generated for five of the Arabidopsis tt4 alleles [tt4(85),

tt4(UV113), tt4(38G1R), tt4(UV01), and tt4(UV25)] based on the crystal structure for

Arabidopsis CHS (Chapter 3). The deduced sequences of the mutant proteins (Saslowsky et al.,

2000; Shirley et al., 1995) were first aligned independently with alfalfa CHS2 using Biology

Workbench (http://workbench.sdsc.edu). Five models were produced for each allele using

Modeller6 of Sali and Blundell (1993) and an average model was generated for each protein by

coordinate averaging. The averaged structures were minimized using the Sander module of

Amber7 for 200 steps using steepest descent (Pearlman et al., 1995). Models were then solvated

with water, and sodium ions were added to give the system a net charge of zero. Minimization

Page 64: Structural Characterization of the Flavonoid Enzyme Complex

49

of the solvent and counterions was performed for 200 steps as described above, followed by a

minimization of the entire system. Molecular dynamics was then performed using the Sander

module of Amber7 for a total of 500 ps with a time step of 2.0 fs and a structure collected at

every 1 ps. All simulations were performed using 4-8 processors on Virginia Tech’s Laboratory

for Advanced Scientific Computing and Applications Linux cluster. The dynamics simulation

was divided into the following components: the first 80 ps was a heating phase to raise the

system temperature from 0 K to either 285 K, 293 K, or 300 K, the next 120 ps was a constant

volume equilibration phase, and the final 300 ps was a constant pressure phase. Final models

were generated by coordinate averaging of the last 100 ps of the dynamics simulation. The

solvent and sodium ions were manually stripped from the average structures and the systems

were minimized again as described above. Assessments of the models were performed using

Procheck on the Whatif server (http://swift.embl-heidelberg.de/servers/). Root mean squared

deviation (RMSD) calculations for the last 100 ps of the dynamics simulations were performed

using the Carnal module of Amber7 (a representative RMSD graph is shown in Figure 2.1).

Distances between alpha carbons, phi-psi torsion angles for residues, and hydrogen bond

networks were determined using the Carnal module of Amber7 using the last 100 ps of the

dynamics simulations. To determine the structural differences between the TT4 proteins and the

wild-type structure, the structures were individually aligned by the least squares fit function

within Swiss-Pdb Viewer, version 3.7 (Guex and Peitsch, 1997) and RMSD was calculated using

the same program. All images of protein structures were generated using Swiss-Pdb Viewer

(Guex and Peitsch, 1997) and rendered using POV-Ray version 3.1a (www.povray.org).

Page 65: Structural Characterization of the Flavonoid Enzyme Complex

50

Results

Seven alleles of the Arabidopsis CHS gene, induced by either EMS or gamma radiation,

were previously characterized at the gene, protein, and end product levels (Table 2.1) (Burbulis

et al., 1996; Saslowsky et al., 2000; Shirley et al., 1995). One of these alleles [tt4(UV118a)]

contains a complex rearrangement, while the other six contain point mutations that either disrupt

splicing to generate a severely- truncated coding region [tt4(2YY6)], introduce a stop codon to

truncate the coding region by 15 codons [tt4(38G1R)], or alter a single amino acid in the gene

product [tt4(85), tt4(UV01), tt4(UV25), and tt4(UV113)]. Most of the mutations completely

abolish the accumulation of flavonoids in affected plants, although one appears to be leaky and

two others display a temperature-sensitive phenotype (unpublished observations and Saslowsky

et al., 2000) (Table 2.1).

To further investigate the effects of small changes in CHS protein sequence on the

properties of the enzyme, the proteins containing single amino acid changes or the C-terminal

truncation were first examined in the context of current information on structure-function

relationships in CHS. A multiple sequence alignment generated using 14 representative

members of the plant type III PKS superfamily, including Arabidopsis CHS, illustrates that each

of the tt4 mutations affect residues that are very highly, if not absolutely, conserved in these

enzymes (Figure 2.2) and that most are also located in highly-conserved stretches of sequence.

Although all five mutations significantly affect flavonoid accumulation, none affect residues

with known functional importance. Mapping of the affected residues on a homology model of

the Arabidopsis wild-type enzyme previously showed that many are located in or near important

functional domains, including the dimerization interface and the catalytic center, suggesting that

Page 66: Structural Characterization of the Flavonoid Enzyme Complex

51

these mutations may have other, indirect effects on various aspects of enzyme function

(Saslowsky et al., 2000) (Figure 2.3). To examine the structural effects of these mutations in

further detail, the crystal structure for the wild-type Arabidopsis enzyme (Chapter 3) was first

subjected to energy minimizations in the presence of solvent and counterions, followed by

molecular dynamics simulations, as described in the Methods. Models were then developed and

similarly refined for the products of the five mutant alleles, tt4(85), tt4(UV01), tt4(UV25),

tt4(UV113), and tt4(38G1R). The effects of the altered residues and the C-terminal truncation

on the overall structure of the enzyme and on the dimensions and architecture of both the active

site and CoA binding site were examined by calculating the RMSD values, measuring the

distances between alpha carbons, and the phi and psi angles.

tt4(85)

The tt4(85) allele was the first Arabidopsis CHS mutation to be reported (Koornneef,

1990). Although plants carrying this allele accumulate wild-type levels of CHS protein that

appears to dimerize normally, flavonoid levels in these plants are reduced to almost undetectable

levels (data not shown, Saslowsky et al., 2000). The tt4(85) allele carries a point mutation that

results in replacement of a glycine at position 268 with serine (G268S) (Shirley et al., 1995).

This glycine residue appears to be absolutely conserved in all type III PKSs and is located in a

beta sheet (β11) in linear and spatial proximity to amino acids involved in CoA binding (L273

and K275) as well as to determinants of substrate specificity (I260 and F271) and product length

(G262) (Jez et al., 2000a) (Figures 2.2 and 2.4), and near an arginine (R265) that has been shown

to be essential for activity of raspberry CHS1 (Zheng et al., 2001). A way of measuring the

difference between two proteins is to fit one onto another structure and calculate the RMSD

Page 67: Structural Characterization of the Flavonoid Enzyme Complex

52

between the two. The RMSD for the tt4(85) when compared to the wild-type structure is 2.48 Å,

also indicative of significant structural differences between the two proteins. To investigate the

effect of the G268S substitution on the geometry of the active site, the distances between the

alpha carbons of the four key active site residues were measured (Table 2.2). The results

indicate that the active site histidine (H309) appears closer to the critical active site cysteine

(C169), but is further from the active site asparagine (N342) and phenylalanine (F220) in tt4(85)

(Table 2.2). In the wild-type enzyme, C169 functions as a nucleophile, removing the CoA

moiety from either coumaroyl-CoA or malonyl-CoA and binding the reaction intermediates

during polyketide formation (Ferrer et al., 1999). Both the phi and psi angles for this residue

appear to be distorted in tt4(85), indicating that the side chain may be in an unfavorable

orientation for catalysis. The phi angle for F271, which is one of two ‘gatekeeper’ residues just

outside the active site, is also much larger than wild-type (Table 2.3), which could affect entry of

substrate into the active site. Moreover, almost all of the residues involved in CoA and substrate

binding appear to be further apart than in the wild-type protein (Table 2.4). The increase in these

distances, some being almost 2 Å larger than that of wild-type, indicates that this protein might

be incapable of binding or correctly positioning the CoA moiety within the active site. The

homology model of the tt4(85) protein also indicates that helix a3, which contains three of the

CoA binding residues is distorted relative to the wild-type, with a kink in the middle of this helix

(Figure 2.4). Although, the tt4(85) protein does retain some activity, as indicated by the leaky

phenotype of mutant plants, it provides a good example of how a single amino acid change can

have substantial structural and mechanistic effects on enzyme function.

Page 68: Structural Characterization of the Flavonoid Enzyme Complex

53

tt4(UV113)

The tt4(UV113) mutation results in the substitution of proline for threonine at position

174 (T174P). This highly-conserved amino acid is located in helix α7 (Figure 2.2; Figure 2.4C)

and is five residues C-terminal to the active site cysteine. As in the case of tt4(85), plants

carrying this mutation produce wild-type levels of CHS protein that appears to dimerize

normally (Saslowsky et al., 2000). However, no detectable flavonoids are produced. T174

appears to be absolutely conserved in CHS enzymes and most other type III PKSs, with the

exception of valerophenone synthase from Hops, where there is a lysine at this position (Figure

2.2). Homology modeling indicates that, although the tt4(UV113) protein maintains its overall

tertiary structure, the spatial relationship of the active site cysteine (C169) to the other active site

residues is altered (Table 2.2). There are also changes in the phi angle for the C169 backbone

relative to the wild-type protein (Table 2.3), which could indicate that the side chain is being

forced into an orientation that is inaccessible to the substrate. Measurement of the distances

between the alpha carbons of the CoA binding residues indicates that, for the most part, the

geometry of these residues is unchanged. However, insertion of a proline residue into the a7

helix is likely to reduce the flexibility of this region, which is indicated by an decrease in the

RMSD for the a7 helix (1.09 ± 0.071 Å for tt4(UV113) and 1.49 ± 0.091 Å for wild-type).

When the same measurement is performed for residues 166-172, the loop containing the active

site cysteine, the RMSD is 0.614 ± 0.106 Å for UV113 and 0.882 ± 0.093 Å for wild-type. Thus,

although the T174P substitution appears to have little effect on the overall structure of the active

site pocket, it could affect the dynamic movement of the active site cysteine during catalysis.

The tt4(UV113) substitution has an even more dramatic effect on enzyme function than the one

in tt4(85), presumably due to a loss in flexibility of the loop containing the active site cysteine.

Page 69: Structural Characterization of the Flavonoid Enzyme Complex

54

tt4(38G1R)

The point mutation in tt4(38G1R) shifts the reading frame starting at codon 381,

introducing 3 altered residues and a premature stop codon so that 15 residues are lost from the C

terminus. The protein accumulates at much lower levels than the wild-type CHS, is catalytically

inactive, and does not form dimers (Table 2.1) (Saslowsky et al., 2000). The first two residues

altered by this mutation are part of the conserved GFGPG loop that provides a scaffold for the

cyclization reaction in type III PKSs (Austin and Noel, 2003; Suh et al., 2000), and are located

on the opposite side of the protein from the dimerization interface (Figure 2.4D).

The model for tt4(38G1R) bears the most striking changes in overall tertiary structure of

all the altered proteins (Figure 2.4). Due to the truncation, beta sheet 15, which appears to be

required for the formation and stabilization of a series of three beta sheets, is lost. The RMSD

between this structure and that of the wild-type enzyme is 2.92 Å, indicating significant

differences in the secondary and tertiary structures of the two proteins. The homology model

also indicates that the residues comprising beta sheets 14 and 7 in the wild-type no longer retain

any beta sheet structure in tt4(38G1R). Interestingly, the loss of this series of beta sheets also

affects one of the active site residues, H309, which functions with N342 as an electron sink to

stabilize the transition state dur ing the transfer of substrate intermediates to C169 (Austin and

Noel, 2003). This residue is located at one end of beta sheet 14 in the wild-type protein and is

moved in tt4(38G1R) (Table 2.2). Moreover, the phi angles for both the gatekeeper amino acids,

F220 and F271, and the psi angles for F271 are significantly different than in the wild-type

protein (Table 2.3), which suggests that these residues could be obstructing the active site

entrance. Also, the CoA binding residues are substantially further apart than in the wild-type

protein (Table 2.4), indicating that these residues may not be able to bind substrate. However,

Page 70: Structural Characterization of the Flavonoid Enzyme Complex

55

from the analyses performed in this study, no reason for the loss of dimerization was determined.

These results suggest the apparent lack of catalytic activity may result from the loss of three beta

sheets required for protein stability and catalytic function.

tt4(UV01)

tt4(UV01) is a temperature-sensitive allele that results in the replacement of a threonine

with an isoleucine at position 49 (T49I). As is the case for G268 in the tt4(85) allele, this residue

appears to be invariant in all type III PKSs (Figure 2.2). Position 49 is near the CoA binding

residues in both the primary sequence and the tertiary structure (Figure 2.3). This amino acid

change does not appear to affect the ability of the enzyme to dimerize, consistent with the fact

that it is located on the opposite side of the enzyme from the dimerization interface, but does

appear to decrease protein stability at all temperatures (Saslowsky et al., 2000) (Table 2.1).

These plants exhibit a temperature-dependent reduction in the accumulation of flavonoid end

products, with almost no flavonoids present at 27°C, very low levels at 20°C (the optimum

growth temperature for Arabidopsis), and moderate levels at 12°C (Saslowsky et al., 2000).

After subjecting the tt4(UV01) homology model to molecular dynamics simulations at 27°, 20°,

and 12°C, and superimposing the resulting structures on that of the wild-type protein, no

substantial changes in overall secondary or tertiary structure were apparent. The phi and psi

angles do not deviate substantially from the wild-type for the majority of the residues with the

exception of F271 (Table 2.3), which is involved in substrate specificity. Interestingly, the phi

angle for F271 is significantly different from wild-type at all three temperatures investigated.

The dimensions of the CoA binding tunnel did vary with temperature, however the greatest

changes occurred at lower temperatures, where the enzyme exhibits some catalytic activity

Page 71: Structural Characterization of the Flavonoid Enzyme Complex

56

(Table 2.4). Overall, the homology model did not provide any clues regarding the temperature-

sensitive phenotype of the tt4(UV01) enzyme.

tt4(UV25)

The tt4(UV25) allele is characterized by the replacement of arginine at position 334 with

a cysteine (R334C) (Saslowsky et al., 2000). This mutation is located eight residues N-terminal

to the active site asparagine (N342) and is highly conserved among type III PKSs. The known

exceptions are two confirmed and two suspected isoforms of acridone synthase from rue

(Junghanns et al., 1995) (accession numbers S60241, Q9FSC0, Q9FSC1, and Q9FSC2,

respectively), which have conservative substitutions of lysine at this position (Figure 2.2).

Similar to tt4(UV01), the tt4(UV25) enzyme is present at reduced levels in mutant plants relative

to wild-type and exhibits a temperature-dependent reduction in flavonoid accumulation,

however, it also appears to have lost the ability to form homodimers (Table 2.1). Molecular

dynamics simulations were performed at 12º, 20º, and 27º C. Analysis of the dihedral angles of

the active site residues as well as the distances between the active site residues, CoA binding

residues, and substrate specificity residues, provided no obvious explanation for the phenotypes

of the tt4(UV25) enzyme. However, the RMSD values are indicative of a structure that is not

very stable at 12º C. Also, the RMSD between the wild-type and the UV25 structures at 12º C is

2.50 Å, indicating that the two structures are different from each other. Both of these results are

opposite to what was expected, as the protein should be more stable at lower temperatures.

Interestingly, this substitution occurs in the region of the protein that is proposed to be important

for interaction with CHI (Figure 2.3) (Chapter 3). Perhaps the alteration of this residue affects

this predicted interaction domain in a temperature-dependent manner, thus impacting not only

Page 72: Structural Characterization of the Flavonoid Enzyme Complex

57

CHS protein accumulation, but its ability to function as part of a proposed flavonoid enzyme

complex.

Discussion

This study indicates that homology modeling can be a useful tool for the analysis of

structure-function relationships in enzymes such as CHS. An allelic series of mutations in the

CHS gene had previously been characterized on the molecular and genetic level. For most of the

alleles, molecular dynamics simulations of homology models were able to suggest a structural

basis for the loss of function, usually due to distortion of the active site pocket, an overall change

in the dimensions of the CoA binding tunnel dimensions, the loss of flexibility in loop, or a loss

of substantial tertiary structure. The molecular dynamics simulations indicated that the

substitutions caused changes in the structures of tt4(85), tt4(UV113), and tt4(38G1R) that may

explain the changes in function and accumulation. However, in the case of the temperature-

sensitive alleles, the molecular dynamics were unable to provide an explanation for the

temperature-dependent loss of function and stability. Also, for the tt4(38G1R) and tt4(UV01)

alleles, this study was unable to provide any predictions for the loss of dimerization of the

protein. It is important to note that the dynamics performed in this study were done on the

monomeric form of CHS, thus simulations on the dimeric form of the protein could provide

additional useful information. Future studies could also include in silico docking of substrate

into the CHS active site to determine if the proteins are able to bind the CoA moiety, which is

required for proper positioning of the substrate.

Page 73: Structural Characterization of the Flavonoid Enzyme Complex

58

Modeling and molecular dynamics simulations can also be applied to other systems

where biochemical data exist for amino acid substitutions in order to gain information about

effects on secondary and tertiary structure. These methods could be used as an inexpensive

approach to rationally engineering a gain or loss of function within proteins. It might also be

possible to engineer novel catalytic activities by first modeling amino acid substitutions and then,

using quantum mechanical calculations, simulate the reaction that would be performed.

Moreover, the techniques described in this study could be used to elucidate the architecture of

multi-enzyme complexes. First, through in silico docking techniques and molecular dynamics

simulations, the identity and location of residues that contribute to protein-protein interactions

could be identified and then verified by experimental techniques such as yeast two-hybrid and

surface plasmon resonance refractometry. Once the interaction has been identified, it might then

be possible to engineer stronger or weaker interactions through amino acid substitutions in order

to modify metabolic flux creating new activities or causing the accumulation of specific end

products.

Page 74: Structural Characterization of the Flavonoid Enzyme Complex

59

Acknowledgments

The authors would like to thank Jory Zmuda Ruscio for her assistance with the molecular

dynamics simulations. This work was supported by the National Science Foundation (grants

MCB-9808117 and MCB-0131010) and by grants from the Virginia Tech Graduate Research

Development Project to C.D.D.

Page 75: Structural Characterization of the Flavonoid Enzyme Complex

60

References

Austin, M. B., and Noel, J. P. (2003). The chalcone synthase superfamily of type III polyketide

synthases. Nat Prod Rep 20, 79-110.

Burbulis, I. E., Iacobucci, M., and Shirley, B. W. (1996). A null mutation in the first enzyme of

flavonoid biosynthesis does not affect male fertility in Arabidopsis. Plant Cell 8, 1013-

1025.

Ceruso, M. A., Periole, X., and Weinstein, H. (2004). Molecular dynamics simulations of

transducin: interdomain and front to back communication in activation and nucleotide

exchange. J Mol Biol 338, 469-481.

Ferrer, J.-L., Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (1999). Structure of

chalcone synthase and the molecular basis of plant polyketide biosynthesis. Nat Struc

Biol 6, 775-784.

Guex, N., and Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an

environment for comparative protein modeling. Electrophoresis 18, 2714-2723.

Jez, J. M., Austin, M. B., Ferrer, J., Bowman, M. E., Schroder, J., and Noel, J. P. (2000a).

Structural control of polyketide formation in plant-specific polyketide synthases. Chem

Biol 7, 919-930.

Jez, J. M., Bowman, M. E., and Noel, J. P. (2002). Expanding the biosynthetic repertoire of plant

type III polyketide synthases by altering starter molecule specificity. Proc Natl Acad Sci

U S A 99, 5319-5324.

Jez, J. M., Ferrer, J. L., Bowman, M. E., Dixon, R. A., and Noel, J. P. (2000b). Dissection of

malonyl-coenzyme A decarboxylation from polyketide formation in the reaction

mechanism of a plant polyketide synthase. Biochemistry 39, 890-902.

Page 76: Structural Characterization of the Flavonoid Enzyme Complex

61

Junghanns, K. T., Kneusel, R. E., Baumert, A., Maier, W., Groger, D., and Matern, U. (1995).

Molecular cloning and heterologous expression of acridone synthase from elicited Ruta

graveolens L. cell suspension cultures. Plant Mol Biol 27, 681-692.

Koornneef, M. (1990). Mutations affecting the testa color in Arabidopsis. Arabid Inf Serv 28, 1-

4.

Moore, B. S., Hertweck, C., Hopke, J. N., Izumikawa, M., Kalaitzis, J. A., Nilsen, G., O'Hare,

T., Piel, J., Shipley, P. R., Xiang, L., et al. (2002). Plant- like biosynthetic pathways in

bacteria: from benzoic acid to chalcone. J Nat Prod 65, 1956-1962.

Pearlman, D. A., Case, D. A., Caldwell, J. W., Ross, W. S., Cheatham, T. A., DeBolt, S.,

Ferguson, D. M., Seibel, G., and Kollman, P. (1995). AMBER, a package of of computer

programs for applying molecular mechanics, a normal mode analysis, molecular

dynamics and free energy calculations to simulate the structural and energetic properties

of molecules. Comput Phys Commun 91, 1-41.

Sali, A., and Blundell, T. L. (1993). Comparative protein modelling by satisfactoin of spatial

restraints. J Mol Biol 234, 779-815.

Saslowsky, D. E., Dana, C. D., and Winkel-Shirley, B. (2000). An allelic series for the chalcone

synthase locus in Arabidopsis. Gene 255, 127-138.

Shirley, B. W., Kubasek, W. L., Storz, G., Bruggemann, E., Koornneef, M., Ausubel, F. M., and

Goodman, H. M. (1995). Analysis of Arabidopsis mutants deficient in flavonoid

biosynthesis. Plant J 8, 659-671.

Stockner, T., Sterk, H., Kaptein, R., and Bonvin, A. M. (2003). Molecular dynamics studies of a

molecular switch in the glucocorticoid receptor. J Mol Biol 328, 325-334.

Page 77: Structural Characterization of the Flavonoid Enzyme Complex

62

Suh, D.-Y., Fukuma, K., Kagami, J., Yamazaki, Y., Shibuya, M., Ebizuka, Y., and Sankawa, U.

(2000). Identification of amino acid residues important in the cyclization reactions of

chalcone and stilbene synthases. Biochem J 350, 229-235.

Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994). CLUSTAL W: improving the

sensitivity of progressive multiple sequence alignment through sequence weighting,

position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22, 4673-

4680.

Tropf, S., Kärcher, B., Schröder, G., and Schröder, J. (1995). Reaction mechanisms of

homodimeric plant polyketide synthases (stilbene and chalcone synthase). J Biol Chem

270, 7922-7928.

Winkel-Shirley, B. (2001). Flavonoid biosynthesis. A colorful model for genetics, biochemistry,

cell biology, and biotechnology. Plant Physiol 126, 485-493.

Zheng, D., Schroder, G., Schroder, J., and Hrazdina, G. (2001). Molecular and biochemical

characterization of three aromatic polyketide synthase genes from Rubus idaeus. Plant

Mol Biol 46, 1-15.

Page 78: Structural Characterization of the Flavonoid Enzyme Complex

63

Table 2.1 Biochemical and genetic characteristics of Arabidopsis CHS alleles.

Allele Mutationa CHS protein levelsb

Allele type Flavonoid glycoside levelsc

Homodimer crosslinkingd

wild-type N.A.e ++++ Wild-type +++ +++

tt4(85)e G268S ++++ Leaky + +++ tt4(UV113) T174P ++++ Null - +++ tt4(38G1R) P381S,

G382H, L383C, T384stop

+ Null - +

tt4(UV01) T49I ++ Temperature-sensitive

+ +++

tt4(UV25) R334C ++ Temperature- sensitive

+ +

a Amino acid(s) altered in the predicted protein. b Determined by immunoblot analysis (Saslowsky et al., 2000). c Determined by HPLC ana lysis (data not shown, Saslowsky et al., 2000). d Determined using disuccinimidyl glutarate and N-succinimidyl-[4-vinylsulfonyl] benzoate

(Saslowsky et al., 2000). e N.D., not determined.

Page 79: Structural Characterization of the Flavonoid Enzyme Complex

64

Table 2.2. Dimensionsa of the CHS proteins

27°C 20ºC 12ºC

ArabidopsisCHS

tt4 (85)

tt4 (UV113)

tt4 (38G1R)

tt4 (UV01)

tt4 (UV25)

Arabidopsis CHS

tt4 (UV01)

tt4 (UV25)

Arabidopsis CHS

tt4 (UV01)

tt4 (UV25)

RMSDb

vs. wt N/A 2.48 1.79 2.92 1.74 1.54 N/A 1.17 1.50 N/A 1.23 2.50

RMSDc

over 100 ps 1.81

(0.06) 2.07

(0.06) 1.87

(0.06) 1.77 (0.1)

1.88 (0.08)

2.16 (0.19)

1.59 (0.07)

1.69 (0.1)

1.67 (0.06)

1.76 (0.07)

1.61 (0.06)

2.08 (0.10)

F220 to N342

4.75 (0.22)

4.78 (0.19)

4.74 (0.18)

4.63 (0.17)

4.80 (0.18)

4.84 (0.22)

4.72 (0.19)

4.65 (0.20)

4.88 (0.19)

4.61 (0.17)

4.61 (0.17)

4.58 (0.17)

C169 to H309

7.74 (0.38)

6.76* (0.22)

7.17 (0.29)

8.22 (0.30)

7.50 (0.29)

7.66 (0.31)

7.64 (0.31)

7.92 (0.26)

8.37 (0.97)

7.44 (0.34)

7.82 (0.26)

7.16 (0.31)

C169 to F220

10.78 (0.42)

11.86* (0.33)

10.49 (0.31)

9.71* (0.34)

10.72 (0.26)

10.83 (0.33)

11.34 (0.32)

10.99 (0.39)

10.42 (0.57)

10.60 (0.33)

11.26 (0.31)

10.42 (0.32)

H309 to N342

8.18 (0.22)

9.14* (0.30)

8.48 (0.24)

7.96 (0.27)

8.52 (0.22)

8.37 (0.22)

7.90 (0.23)

8.28 (0.25)

8.16 (0.25)

8.13 (0.25)

8.01 (0.27)

8.49 (0.25)

C169 to N342

9.19 (0.32)

9.34 (0.27)

8.61* (0.29)

9.22 (0.36)

8.85 (0.22)

9.25 (0.26)

9.90 (0.28)

9.36* (0.31)

9.27 (0.31)

9.24 (0.28)

9.91 (0.27)

8.87 (0.29)

F220 to H309

12.21 (0.33)

13.44* (0.33)

12.64 (0.26)

11.43* (0.28)

12.69 (0.25)

12.41 (0.31)

11.84 (0.32)

12.30 (0.31)

12.16 (0.43)

11.94 (0.32)

11.19 (0.27)

12.27 (0.25)

aValues are in angstroms and represent the average of distances between alpha carbons of the indicated residues in the final 100 structures generated by dynamic simulation. Numbers in parentheses are standard deviations. bRMSD values, which indicate overall similarity of structures relative to Arabidopsis wild-type CHS, were calculated using Swiss-PdbViewer v3.7. cRMSD values, which indicate the overall stability of the Ca backbone, was calculated using Amber7 (Pearlman et al., 1995). An asterisk (*) indicates a 1 Å or greater deviation from the wild-type.

Page 80: Structural Characterization of the Flavonoid Enzyme Complex

65

Table 2.3. Phi-psi torsion angles within the CHS active sitea.

27°C 20ºC 12ºC

Wild-type

tt4 (85)

tt4 (UV113)

tt4 (38G1R)

tt4 (UV01)

tt4 (UV25)

Wild-type

tt4 (UV01)

tt4 (UV25)

Wild-type

tt4 (UV01)

tt4 (UV25)

C169 φ -58.63 (10.61)

41.46* (10.3)

-44.50 (12.1)

-62.49 (10.0)

-49.95 (10.3)

-58.62 (10.2)

-54.47 (8.23)

-60.72 (9.2)

-57.40 (10.46)

-56.92 (7.9)

-54.29 (10.0)

-48.26 (12.1)

C169 ψ -19.71 (14.3)

-100.98* (11.8)

-37.84 (12.0)

-7.04 (10.5)

-29.12 (10.3)

-13.00 (8.5)

-28.12 (10.3)

-14.01 (9.6)

-17.93 (10.3)

-25.69 (9.1)

-23.81 (10.0)

-35.71 (11.17)

F220 φ -65.59 (11.0)

-56.17 (8.9)

-67.35 (9.0)

-76.58* (9.6)

-68.8 (8.4)

-61.47 (10.7)

-72.94 (11.6)

-87.83 (13.7)

-63.77 (9.1)

-75.98 (9.8)

-83.21 (10.1)

-80.89 (9.7)

F220 ψ 142.6 (10.2)

133.1 (8.7)

137.1 (7.4)

150.2 (8.1)

136.4 (9.6)

143.5 (10.3)

147.5 (8.1)

153.3 (7.2)

140.7 (8.7)

156.59 (7.0)

152.9 (6.4)

152.17 (7.0)

H309 φ -68.6 (10.3)

-66.6 (9.5)

-72.5 (11.0)

-68.5 (11.1)

-74.0 (11.6)

-73.42 (12.4)

-72.5 (11.0)

-71.7 (13.2)

-71.3 (10.0)

-74.0 (12.0)

-66.7 (10.7)

-66.0 (11.6)

H309 ψ 120.0 (7.5)

108.4 (8.6)

118.5 (7.2)

115.1 (6.7)

118.6 (7.0)

115.2 (7.1)

113.2 (6.8)

113.7 (5.9)

115.1 (7.1)

114.8 (6.5)

115.4 (6.4)

115.5 (5.8)

N342 φ -87.02 (11.7)

-73.5 (8.5)

-89.6 (18.9)

-83.1 (10.9)

-128.7* (23.7)

-81.4 (10.4)

-79.8 (10.2)

-79.9 (12.2)

-81.6 (12.6)

-81.42 (10.0)

-78.3 (11.4)

-79.8 (10.6)

N342 ψ 96.0

(10.7) 97.6

(10.4) 93.8

(14.0) 99.7 (8.0)

135.21* (12.0)

96.5 (9.7)

104.1 (12.2)

112.4 (12.9)

94.9 (10.1)

96.5 (10.7)

112.6* (16.3)

103.3 (10.2)

F271 φ -78.5 (15.3)

-148.0* (9.0)

-126.0* (34.6)

-56.5* (8.7)

-146.1* (10.2)

-155.5* (8.9)

-149.9 (34.2)

-77.3* (19.4)

-88.3* (18.4)

-89.6 (22.2)

-150.8* (8.4)

-146.7* (8.4)

F271 ψ 120.7 (8.7)

138.7* (11.4)

144.3* (8.0)

134.2* (9.8)

136.5* (13.2)

143.0* (11.3)

129.2 (9.2)

149.5* (7.9)

121.2 (12.2)

127.7 (36.7)

133.9* (103.5)

125.5 (9.1)

aValues are in degrees. The phi angle (φ) indicates rotation around the N-Cα bond, while the psi angle (ψ), indicates rotation around the Cα -C bond. Positive and negative values indicate right-handed and left-handed rotation, respectively. Standard deviations are given in parentheses. An asterisk (*) indicates rotation 15º greater than wild-type.

Page 81: Structural Characterization of the Flavonoid Enzyme Complex

66

Table 2.4. Distances between CoA binding residues.

27°C 20ºC 12ºC

Arabidopsis CHS

tt4 (85)

tt4 (UV113)

tt4 (38G1R)

tt4 (UV01)

tt4 (UV25)

Arabidopsis CHS

tt4 (UV01)

tt4 (UV25)

Arabidopsis CHS

tt4 (UV01)

tt4 (UV25)

K67 to K275

17.7 (0.5)

19.2* (0.8)

16.6* (0.4)

18.8 (0.7)

18.3 (0.5)

17.1 (0.7)

17.6 (0.6)

17.8 (0.6)

17.4 (0.5)

16.8 (0.6)

17.1 (0.5)

17.0 (0.4)

K67 to L273

15.7 (0.4)

16.4 (0.8)

14.8* (0.4)

16.9* (0.8)

16.0 (0.5)

15.1 (0.6)

15.3 (0.6)

15.8 (0.6)

15.4 (0.5)

15.3 (0.8)

13.8* (0.5)

15.3 (0.4)

K60 to K275

18.0 (0.4)

20.8* (0.6)

17.4 (0.6)

19.9* (0.5)

18.9* (0.5)

17.8 (0.7)

18.8 (0.5)

19.9 (0.6)

18.0 (0.4)

17.8 (0.5)

18.7 (0.5)

18.0 (0.5)

K60 to L273

15.8 (0.3)

17.9* (0.4)

15.5 (0.4)

17.2* (0.5)

15.8 (0.5)

15.3 (0.5)

15.8 (0.4)

16.9* (0.7)

16.1 (0.4)

15.8 (0.5)

15.0 (0.5)

15.7 (0.4)

R63 to K275

18.8 (0.4)

21.0* (0.6)

18.2 (0.5)

20.6* (0.5)

19.5 (0.6)

18.4 (0.7)

19.1 (0.5)

17.8* (0.6)

19.0 (0.4)

18.2 (0.7)

18.9 (0.5)

18.5 (0.4)

R63 to L273

16.7 (0.4)

18.0* (0.5)

16.3 (0.3)

18.2* (0.7)

16.7 (0.6)

16.0 (0.6)

16.3 (0.4)

15.8* (0.6)

16.9 (0.3)

16.4 (0.8)

15.3* (0.5)

16.4 (0.4)

K67 to A314

8.9 (0.3)

11.0* (0.6)

9.2 (0.4)

9.4 (0.5)

9.6 (0.3)

9.4 (0.5)

9.2 (0.4)

9.6 (0.4)

9.7 (0.4)

9.0 (0.4)

9.5 (0.4)

9.0 (0.3)

F220 to F271

10.5 (0.3)

10.6 (0.3)

10.6 (0.4)

10.1 (0.3)

10.5 (0.4)

10.3 (0.3)

10.7 (0.3)

9.6 (0.3)

9.1 (0.4)

10.9 (0.5)

10.6 (0.3)

11.1 (0.3)

K60 to I260

20.6 (0.4)

22.4* (0.5)

20.9 (0.4)

21.9* (0.4)

20.5 (0.4)

20.4 (0.6)

20.1 (0.4)

20.9 (0.7)

20.8 (0.4)

20.9 (0.7)

20.4 (0.5)

20.2 (0.5)

K60 to G262

22.8 (0.4)

22.8 (0.4)

22.2 (0.4)

23.3 (0.5)

21.8 (0.4)

21.9 (0.5)

22.4 (0.4)

22.6 (0.7)

23.1 (0.5)

22.0 (0.4)

22.8 (0.4)

22.2 (0.5)

R63 to I260

21.3 (0.5)

22.4* (0.5)

21.3 (0.3)

22.3* (0.5)

20.8 (0.5)

20.7 (0.7)

20.3 (0.4)

20.6 (0.7)

21.4 (0.4)

21.7 (0.9)

20.4* (0.5)

20.6 (0.4)

R63 to G262

23.9 (0.4)

23.8 (0.4)

23.4 (0.4)

24.5* (0.5)

23.2 (0.5)

23.0 (0.5)

23.5 (0.4)

22.9 (0.7)

24.1 (0.4)

22.9 (0.5)

23.3 (0.5)

23.8 (0.4)

K67 to I260

19.7 (0.6)

20.3 (0.6)

18.9 (0.4)

20.2 (0.7)

19.1 (0.5)

18.8 (0.9)

18.6 (0.6)

18.3 (0.6)

19.2 (0.6)

20.3 (0.9)

18.1* (0.5)

18.8* (0.4)

K67 to G262

22.9 (0.4)

22.7 (0.6)

22.2 (0.6)

23.5 (0.7)

22.9 (0.6)

22.2 (0.6)

22.9 (0.5)

21.4* (0.6)

22.6 (0.5)

21.7 (0.5)

21.8 (0.5)

23.2* (0.5)

Page 82: Structural Characterization of the Flavonoid Enzyme Complex

67

00.2

0.40.60.8

1

1.21.41.6

1.82

0 100 200 300 400 500

Time (ps)

RM

SD

(A

ngst

rom

s)

Dana, et al., Figure 2.1

Page 83: Structural Characterization of the Flavonoid Enzyme Complex

68

Figure 2.1. The RMSD of the backbone atoms for wild-type CHS from the starting

confirmation throughout the simulation.

Page 84: Structural Characterization of the Flavonoid Enzyme Complex

69

* * * * *************** Arabidopsis CHS …YFRITNSEHMTDLKEKFKRMCDKS…………YQQGCFAGGTVLRIAKD…………AIDGHLREVGLTFHLLKD…………RATRHVLSEYGNMSS…………GFGPGLTVETVVLHSVPL. 100% Tea CHS2 …YFRITNSEHKTELKEKFQRMCDKS…………YQQGCFAGGTVLRLAKD…………AIDGHLREVGLTFHLLKD…………RATRHVLSEYGNMSS…………GFGPGLTVETVVLHSVST. 84% Snapdragon CHS …YFRITNSEHMTELKEKFKRMCDKS…………YQQGCFAGGTVLRMAKD…………AIDGHLREVGLTFHLLKD…………RSTRQVLSEYGNMSS…………GFGPGLTVETVVLHSVPLN. 84% Parsley CHS …YFRITNSEHMTDLKLKFKRMCDKS…………YQQGCFAGGTVLRLAKD…………AIDGHLREVGLTFHLLKD…………RATRQVLSDYGNMSS…………GFGPGLTVETVVLHSVPATFTH. 84% Raspberry CHS1 …YFRITKSEHKTELKEKFQRMCDKS…………YQQGCFAGGTVLRLAKD…………AIDGHLREVGLTFHLLKD…………EATRNILSEYGNMSS…………GFGPGLTVETVVLHSVAAST. 84% Maize CHS C2 …YFRITKSEHLTDLKEKFKRMCDKS…………YQQGCFAGGTVLRVAKD…………AIDGHLREVGLTFHLLKD…………RATRHVLSEYGNMSS…………GFGPGLTVETVVLHSVPITTGAATA. 82% Alfalfa CHS2 …YFKITNSEHKTELKEKFQRMCDKS…………YQQGCFAGGTVLRLAKD…………AIDGHLREAGLTFHLLKD…………NATREVLSEYGNMSS…………GFGPGLTIETVVLRSVAI. 81% Hops VPS …YFRVTKSEHMTDLKKKFQRMCEKS…………YQLGCYAGGKVLRIAKD…………AIAGHVTEAGLTFHLLRD…………KASREMLSEYGNMSC…………GFGPGLTVETVVLHH. 74% Grape THS2 …YFRVTKSEHMTALKKKFNRICDKS…………DQQGCYAGGTVLRTAKD…………AIAGNLREVGLTFHLWPN…………EATRHVLSEYGNMSS…………GFGPGLTIETVVLHSIPMVTN. 73% Rue ACS2 …YFRITNSEHMTDLKEKFKRMCDKS…………YQQGCYAGGTVLRLAKD…………AIEGHIREEGLTVHLKKD…………RASKHVMSEYGNMSS…………GFGPGLTVETVVLRSVPVEA. 72% Hydrangea CTAS …YFRVTKSDHLTDLKSKFKRMCERS…………YQQGCHAGGTGLRLAKD…………AIDGHLRQEGLTFHLRKD…………RVSRHILSEYGNMSS…………GFGPGFTVETIVLHH. 71% Daisy PS …YFRVTKSEHMVDLKEKFKRICEKT…………YQQGCAAGGTVLRLAKD…………AMKLHLREGGLTFQLHRD…………RASRHVLSEYGNLIS…………GFGPGMTVETVVLRSVRVTAAVANGN. 68% Orchid BBS …YFRVTNSEHLVDLKKKFQRICEKT…………YQQGCFAGGTTLRLAKC…………AIGGHVSEGGLLATLHRD…………SVSRHVLAEYGNMSS…………GFGPGLTVETVVLRSVPL. 65% Peanut STS …YFRVTNGEHMTDLKKKFQRICERT…………YHQGCFAGGTVLRLAKD…………AIGGLLREVGLTFYLNKS…………KATRDVLSNYGNMSS…………GFGPGLTIETVVLRSMAI. 64% Pine DPSS …YFRITGNEHNTELKDKFKRICERS…………FQHGCFAGGTVLRMAKD…………AIGGKVREVGLTFQLKGA…………IPTRHVMSEYGNMSS…………GFGPGLTIETVVLKSVPIQ. 64% •• • ••• • •• •• • • • ••• •• •• • • • ••• ••••••• •• ••

Dana, et al., Figure 2.2

165

188

276

259

331

345

378 TT4(UV113)

TT4(85)

TT4(UV25)

TT4(38G1R)

45

68 TT

4(UV01)

α2 α3 α7 α11 β10 β11 β5 β14 β15

Page 85: Structural Characterization of the Flavonoid Enzyme Complex

70

Figure 2.2. Multiple sequence alignment of selected regions of a representative group of plant

type III polyketide synthases. CHS, acridone synthase (ACS), bibenzyl synthase (BBS),

coumaroyl triacetic acid synthase (CTAS), dihydropinosylvin-forming stilbene synthase (DPS),

2-pyrone synthase (PS), trihydroxystilbene synthase (THS), and valerophenone synthase (VPS)

sequences from a variety of plant species were aligned using CLUSTAL W (Thompson et al.,

1994). Highlighting indicates catalytic residues (red), CoA binding residues (purple), and active

site residues that are structural (yellow), control polyketide length (green), or determine substrate

specificity (blue) (Jez et al., 2000b). Residues that are altered in the four alleles are indicated

with asterisks; numbering is relative to the Arabidopsis CHS protein sequence. Percentages to

the right of the sequences refer to amino acid identity relative to Arabidopsis CHS. Pink dots

below the sequences indicate residues that are absolutely conserved among type II PKSs; green

dots indicate those conserved in CHS but not other type II PKSs. Elements of secondary

structure in the crystal structure of Arabidopsis CHS are identified below the sequences.

Page 86: Structural Characterization of the Flavonoid Enzyme Complex

71

Dana, et al., Figure 2.3

A. B.

Page 87: Structural Characterization of the Flavonoid Enzyme Complex

72

Figure 2.3. Structure of the Arabidopsis wild-type CHS enzyme. The structure represented

here is the 100 ps average from the molecular dynamics simulations at 27º C. Labeled on this

structure are the active site (red), determinants of substrate specificity (orange), product length

determinant (green), amino acids responsible for the binding of the CoA moiety (blue), the

locations of the tt4 substitutions or truncation (purple), and the residues proposed to be involved

in the interaction between CHS and CHI (yellow). A, front view; B, side view.

Page 88: Structural Characterization of the Flavonoid Enzyme Complex

73

Dana, et al., Figure 2.4

A. B.

C. D.

ß15

a3 ß10

ß11

a7 ß14

Page 89: Structural Characterization of the Flavonoid Enzyme Complex

74

Figure 2.4. Models of the Arabidopsis wild-type and tt4 CHS enzymes. The structures

were generated based on the crystal structure CHS from Arabidopsis as described in the

Methods. Active site residues (wild-type) and substituted residues (alleles) are labeled in

red. The secondary elements labeled in green indicate structural changes relative to the

wild-type. A, 27º C wild-type CHS structure; B, tt4(85) with alpha helix a3, showing

disorder, and beta sheets ß10 and ß11 labeled to indicate the location of the substitution;

C, tt4(UV113) with alpha helix a7 and the loop containing the active site cysteine

labeled; D, tt4(38G1R) with the areas corresponding to the location of beta sheets ß7,

ß12, ß13, and ß14 in the wild-type labeled to identify the loss of secondary structure

resulting from the truncation.

Page 90: Structural Characterization of the Flavonoid Enzyme Complex

75

Chapter 3

Structural Characterization of the Chalcone Synthase-Chalcone Isomerase

Interaction

Christopher D. Dana1, Stéphane B. Richard2, Susan Krueger3, Joesph P. Noel2, and

Brenda S. J. Winkel1

1Department of Biology and Fralin Biotechnology Center, Virginia Tech, 2Structural

Biology Laboratory, The Salk Institute for Biological Sciences, and 3NIST Center for

Neutron Research

Page 91: Structural Characterization of the Flavonoid Enzyme Complex

76

Abstract

Biosynthetic pathways are often organized within cells as groups of interacting enzymes

known as enzyme complexes or metabolons. Although this organization appears to be

widespread in biological systems, it is still poorly understood. The flavonoid pathway of

Arabidopsis thaliana has been developed in our laboratory as a model for studying

enzyme complexes (Burbulis and Winkel-Shirley, 1999). Several of the enzymes of this

pathway have been shown to interact in vitro, and have also been found to be co- localized

in roots (Burbulis and Winkel-Shirley, 1999; Saslowsky and Winkel-Shirley, 2001). Our

studies are currently focusing on the first enzymes of the flavonoid biosynthetic pathway,

chalcone synthase (CHS) and chalcone isomerase (CHI), which catalyze the condensation

and cyclization of coumaroyl-CoA and 3 malonyl-CoA into the central flavonoid

backbone. We propose a structure of the Arabidopsis CHS-CHI bienzyme complex as

deduced by X-ray crystallography, small-angle neutron scattering (SANS), in silico

docking, and molecular dynamics simulations. Surface-exposed residues thought to be

involved in this interaction have been altered and tested by yeast 2-hybrid analysis.

These experiments provide the first structural insights into the organization of flavonoid

biosynthesis as well as establishing SANS as a technique that can be used in the study of

metabolic organization.

Page 92: Structural Characterization of the Flavonoid Enzyme Complex

77

Introduction

The study of multi-enzyme complexes, or metabolons, dates to the late 1950s with

the isolation of a complex of enzymes involved in the citric acid cycle (Green, 1958). As

time progressed, many more examples of multi-enzyme complexes were discovered,

introducing the idea that biosynthetic pathways are organized as metabolons and leading

to a broader understanding of how the cytoplasm of the cell is organized. Metabolic

complexes can be characterized as either static, which are very stable and require little

energy for association (Ovadi and Srere, 2000), or dynamic, requiring constant energy

input for complex integrity (Welch, 1977). This organization also has the potential to

offer numerous advantages to the cell such as coordinated flow of substrate from one

active site to another, high local substrate concentration, and sequestering of substrates as

protection for the both the cell and the pathway (Ovadi and Srere, 2000). Metabolic

complexes also provide a mechanism by which the types of end products being produced

can be quickly and effectively altered. It has even been suggested that all biosynthetic

pathways are organized as multi-enzyme complexes (Beeckmans et al., 1994; Beeckmans

et al., 1989; Gontero et al., 1988; Robinson et al., 1987).

Flavonoids constitute an important class of secondary metabolites produced in

higher plants. Chalcone synthase (CHS), catalyzes the first enzymatic reaction in

flavonoid biosynthesis, the condensation of three malonyl-CoA (from acetate

metabolism) molecules with coumaroyl-CoA (from the general phenylpropanoid

pathway) to create tetrahydroxychalcone (Figure 3.1). Chalcone isomerase (CHI),

catalyzes the creation of naringenin, which serves as the main flavonoid backbone for

Page 93: Structural Characterization of the Flavonoid Enzyme Complex

78

ensuing enzymatic reactions that create flavonols, anthocyanins, condensed tannins, and

isoflavonoids (Figure 3.1). The products of general flavonoid biosynthesis have diverse

roles in plants, including serving as pigments in seeds, flowers and fruits and as UV

sunscreens (reviewed in Winkel-Shirley, 2002). These plant products also have anti-

oxidant, anti-dipsotrophic, and anti-cancer properties in mammals (reviewed in Havsteen,

2002). Recent metabolic engineering efforts have been directed at altering the amounts

and types of these compounds for nutritional enhancement of plants as well as

engineering plants with novel flower pigmentation (Forkmann and Martens, 2001; Muir

et al., 2001; Yu et al., 2003).

Some 30 years ago, Helen Stafford speculated that the flavonoid pathway

assembles as a multi-enzyme complex (Stafford, 1974a; Stafford, 1974b). The first

indirect evidence for the existence of a flavonoid metabolon was published by Fritsch and

Griseback (1975) who showed that operationally-soluble enzymes involved in flavonoid

biosynthesis localize to the endoplasmic reticulum (ER). Later sucrose density and

immunolocalization experiments verified that enzymes of both the general

phenylpropanoid and flavonoid pathways are localized to the cytoplasmic face of the ER

(Hrazdina and Wagner, 1985; Hrazdina et al., 1987; Wagner and Hrazdina, 1984). More

recent work demonstrated the first direct evidence of interactions between flavonoid

enzymes via affinity chromatography, yeast 2-hybrid analysis, and

coimmunoprecipitation experiments (Burbulis and Winkel-Shirley, 1999).

Immunofluorescence and immunoelectron microscopy experiments have shown that CHS

and CHI co- localize at the ER and that the intracellular distribution of these enzymes is

altered in an flavonoid 3’-hydroxylase (F3’H) mutant, consistent with a previously-

Page 94: Structural Characterization of the Flavonoid Enzyme Complex

79

proposed role for F3’H as a membrane anchor for the complex (Stafford, 1990). This

work also showed an asymmetric distribution in root cortex cells that points to a possible

physiological significance of the organization in the control of auxin transport by

flavonoids. Additional evidence for a flavonoid metabolon comes from the expression of

anti-CHI antibodies in transgenic plants. These antibodies did not appear to change the

catalytic activity of CHI, but did alter the production of flavonoids in seedlings, possibly

by interfering with assembly of CHI into the flavonoid metabolon (Santos et al., 2004).

Deducing the architecture of the flavonoid complex and the mechanisms governing its

assembly and localization could yield critical new insights for effective engineering of

this and other metabolic systems.

Small angle neutron scattering (SANS) is a technique that probes three-

dimensional structure in a non-destructive manner (Koch et al., 2003). Data from SANS

experiments provides a low resolution structure for molecules in an aqueous

environment, independent of the packing inherent in crystal structures. This technique

has many useful applications in polymer science, materials science, and in the life

sciences. Examples of the latter include aiding in the determination of the structure of the

30S ribosomal subunit (Capel et al., 1987), elucidating better ways to crystallize a protein

by determining the amount of protein that is aggregated (Velev et al., 1998), structural

characterization of the GroEL/ES complex (Krueger et al., 2003), and structural changes

in SecA resulting from nucleotide binding (Bu et al., 2003).

SANS, together with in silico docking algorithms and molecular dynamics

simulations were used to develop a model of the CHS-CHI bienzyme complex. To

further define the interactions between CHS and other enzymes in flavonoid biosynthesis,

Page 95: Structural Characterization of the Flavonoid Enzyme Complex

80

four surface-exposed phenylalanine and tryptophan residues on CHS were converted to

alanine and assayed by yeast 2-hybrid analysis. The molecular dynamics simulations

helped identify a series of residues that may be involved in an electrostatic interaction

between CHS and CHI that are being assessed in parallel 2-hybrid experiments. By

understanding the interaction between CHS and CHI, it is hoped that a better perspective

on metabolon architecture can be achieved, which in turn can be used in experiments to

better understand protein complex assembly and regulation as a whole.

Materials and Methods Site Directed Mutagenesis of pET32a (+)

The pET32a (+) vector was modified to allow cleavage of both the thioredoxin

(TRX) and hexa-histadine tags from the protein of interest, leaving just one spurious

amino acid (serine) on the N-terminus. Mutagenesis of pET32a(+) was performed using

the Quik Change site-directed mutagenesis kit (Stratagene La Jolla, CA). Primers were

designed to incorporate an NcoI site next to the thrombin proteolytic cleavage site. The

primers were 5’-GGTGCCACGCGGTTCCATGGTATGAAAGAA ACC-3’, with a

second primer containing the complementary sequence. The underlined portion of the

primer denotes the sequence that was altered. Mutagenesis was performed with Pfu

Turbo (Stratagene) under the following conditions: 95ºC for 30 s as an initial denaturing

step followed by 18 cycles of 95ºC for 30 s, 55ºC for 1 min, and 68ºC for 12 min. The

PCR product was treated with DpnI (Promega) for 1 h at 37ºC to digest the methylated

template DNA, then used to transform E. coli XL1-Blue MRF’ (Novagen) cells using a

heat shock method (Ausubel et al., 1989). The transformed cells were grown overnight at

Page 96: Structural Characterization of the Flavonoid Enzyme Complex

81

37ºC on LBAmp100 plates. Confirmation of successful mutagenesis was via PCR using the

primer, 5’-GGTGCCACGCGGTTCCAT-3’, and the T7 terminator primer and then

sequenced using the BigDye Terminator kit (Stratagene) in order to verify sequence

fidelity. This modified pET32a plasmid was given the designation pCD1.

Creation of Expression Constructs

The CHS and CHI coding regions were subcloned into the NcoI and NotI or SalI

sites, respectively, of the pCD1 expression vector to create pCD1-CHS and pCD1-CHI.

PCR was performed using Pfu turbo (Stratagene) and the following primers: 5’-CATGC

CATGGTGATGGCTGGTGC-3’ and 5’-AGCGGCCGCTTAGAGAGGAACGC TGTG-

3’ for CHS and 5’-CATGCCATGGGAATGTCTTCATCCAACGC-3’ and 5’-

CGGTCGACTCAGTTCTCTTTGGCTAG-3’ for CHI, with pCHS.cr2 or pCHI.cr as the

templates (Pelletier and Shirley, 1996). Reactions were carried out under the following

conditions: 95ºC for 2 min as an initial denaturing step followed by 50 cycles of 94ºC for

30 s, 52ºC for 30 s, and 72ºC for 2 min and then a final elongation step of 72ºC for 10

min. The PCR products were then digested with NcoI and NotI (CHS) or SalI (CHI). T4

DNA Ligase (Promega, Madison, WI) was used to insert these fragments into the pCD1

expression vector. The ligation reaction was then used to transform E. coli DH5a cells

by electroporation (Dower et al., 1988). Transformed colonies were recovered on

LBAmp100 medium and screened by isolating plasmids using an alkaline lysis miniprep

(Ausubel et al., 1989; Birnboim and Doly, 1979) which were then digested with with

NcoI and NotI (CHS) or SalI (CHI). Positive clones were sequenced using the BigDye

Page 97: Structural Characterization of the Flavonoid Enzyme Complex

82

Terminator kit (Stratagene, La Jolla, CA) in order to verify sequence fidelity. Plasmids

were then used to transform E.coli BL21(DE3) pLys-S (Novagen) using a heat shock

method (Ausubel et al., 1989) for expression and purification of recombinant protein.

Expression and Purification of pCD1-CHS and -CHI

A single colony containing either pCD1-CHS or pCD1-CHI was used to inoculate

LBAmp100-Cm34, which was grown overnight, then diluted 1:100 into the same medium.

The cultures were grown to mid- log phase (OD595 = 0.5-0.7) at 37ºC (approximately 4h),

then isopropyl-beta-D-thiogalactopyranoside (Alexis Biochemicals) was added to a final

concentration of 0.5 mM to induce protein expression. The cultures were grown at room

temperature with shaking (250 RPM) for a further 4 h. The cells were harvested by

centrifugation at 2500 x g for 10 min at 4ºC and frozen at -80º C until protein

purification.

Frozen cells were resuspended in lysis buffer (50 mM Tris, pH 8.0; 150 mM

NaCl; 10% glycerol; 1% Tween-20 and 10 mM ß-mercaptoethanol, approximately 33 ml

of lysis buffer per one liter of culture) and resuspended by stirring at 4ºC for 30 min. The

mixture was then sonicated six times for 30 sec each time on ice, with 2 min between

sonication. This lysate was centrifuged at 21,000 xg for 1 h at 4ºC. The supernatant

(crude extract) was passed over a Ni-NTA (Qiagen) column, which was packed and

equilibrated with 10 bed volumes of ddH2O and then 10 bed volumes of lysis buffer. The

column was washed with 10 bed volumes of lysis buffer and then with the same volume

of wash buffer (50 mM Tris, pH 8.0; 500 mM NaCl; 10 mM imidazole, pH 8.0; 10%

Page 98: Structural Characterization of the Flavonoid Enzyme Complex

83

glycerol; and 10 mM ß-mercaptoethanol). Elution buffer (lysis buffer containing 250

mM imidazole, pH 8.0, but no Tween-20) was then passed over the column and four

fractions were collected (0.5, 2, 1 and 1 bed volumes). The fractions were analyzed by

SDS-PAGE in order to determine purity. Next, the TRX and hexa-histadine fusion

partner was removed by the addition of the protease, thrombin (Sigma) (1:1000 dilution

of a 1U/µl solution), into the protein preparation during dialysis into lysis buffer,

performed using SpectraPor 12-14,000 MWCO dialysis tubing (Spectrum), overnight at

4ºC, with stirring.

The dialyzed sample was analyzed by SDS-PAGE to confirm complete cleavage

of the TRX and hexa-histadine fusion partner from the protein of interest. The protein

preparation was then passed over a benzamidine-sepharose 4B (Amersham Biosciences)

to remove the thrombin. The TRX and hexa-histidine partner was removed by passing

the protein solution back over the Ni-NTA column and collecting the flow-through.

After sample purity was assessed by SDS-PAGE, the protein was dialyzed into 25 mM

HEPES, pH 7.5, 100 mM NaCl, 2 mM dithiothreitol (DTT), and 30-50% glycerol (in

order to concentrate the protein). The final purification step was carried out using a

Superdex 200 (Amersham Biosciences) gel filtration column on an Äkta FPLC system at

a flow rate of 0.5 ml/min with a pressure limit of 0.5 MPa. The column was prepared by

first passing over 10 bed volumes of ddH2O and then the same volume of FPLC buffer

(25 mM HEPES, pH 7.5, 100 mM NaCl, and 2-5 mM DTT). Fractions were collected,

assessed by SDS-PAGE, and those containing proteins of the correct size were pooled.

Protein concentration was determined either by the Bio-Rad microassay or based on an

A280 extinction coefficient calculated for each of the proteins. The protein preparation

Page 99: Structural Characterization of the Flavonoid Enzyme Complex

84

was then concentrated to 20-30 mg/ml using Centriprep and Centricon (Millipore)

centrifugal filter devices.

Determination of the CHS Crystal Structure

CHS crystals were grown by vapor diffusion in hanging drops containing a 1:1

mixture of protein and crystallization buffer (16% PEG 8000 and 0.1 M TAPS pH 7.7) at

4º C. The crystals were stabilized in 18% PEG 8000, 0.1M TAPS, pH 7.7, and 15%

glycerol. CHS crystals grew in the space group P21 with 4 monomers per asymmetric

unit cell, with unit cell dimensions of a= 55.06, b=139.27 and c=109.60. Diffraction data

were collected from a single crystal mounted in a cryoloop and flash frozen in a nitrogen

stream at 105K. Data were collected at beamline 7-1 at the Stanford Synchrotron

Radiation Laboratory (SSRL 7-1) on a 30 cm MAR imaging plate detector. All images

were indexed and integrated using DENZO and the reflections merged with

SCALEPACK (Otwinowski and Minor, 1997). Data reduction was completed using

programs in CCP4 (Collaborative Computational Project, 1994) (Table 3.1). Structures

were determined using AMoRE (Navaza, 1994) with an Arabidopsis CHS homology

model (Saslowsky et al., 2000) as the starting structure. Refinement of the protein model

was performed with CNS (Brunger et al., 1998). Model building was carried out with the

program O (Jones et al., 1991) using SIGMAA-weighted |2Fo-Fc|, |Fo-Fc| electron density

maps (Collaborative Computational Project, 1994). Refinement consisted of iterative

cycles of simulated annealing, positional refinement, and B-value refinement. Model

quality was assessed with PROCHECK (Laskowski et al., 1993). All structure images

Page 100: Structural Characterization of the Flavonoid Enzyme Complex

85

were manipulated with SwissPDB Viewer (Guex and Peitsch, 1997) and rendered with

POV-Ray (www.povray.org).

Multiple Sequence Alignment of CHS Proteins

A multiple sequence alignment of 15 representative members of the plant type III

PKS superfamily was generated using Clustal W (Thompson et al., 1994) using the

Megalign program in DNAStar (Madison, WI). Included in this analysis were CHS from

Arabidopsis (accession number P13114), alfalfa (P30074), maize (SYZMCC), parsley

(S42523), raspberry (AAK15174), snapdragon (P06515), and tea (P48387); acridone

synthase 2 from rue (Q9FSC0); bibenzyl synthase from orchid (S71619); coumaroyl

triacetic acid synthase from hydrangea (BAA32733); dihydropinosylvin-forming stilbene

synthase from pine (Q02323); PS from daisy (P48391); trihydroxystilbene synthase

sequences from grape (P51070) and peanut (S00334); and valerophenone synthase

(O80400) from hops.

Site-Directed Mutagenesis

Mutagenesis of CHS was carried out using the pCD1-CHS construct and the

QuikChange® system. The sequences of the primers used for F293A, F305A, W306A,

and W373A are given in Table 3.2. Mutagenesis was performed with Pfu Turbo

(Stratagene, La Jolla, CA) under the following conditions: 95ºC for 30 s as an initial

denaturing step followed by 16 cycles of 95ºC for 30 s, 55ºC for 1 min, and 68ºC for 7

Page 101: Structural Characterization of the Flavonoid Enzyme Complex

86

min. Recovery of modified CHS plasmid constructs was as described as above for

construction of pCD1. Confirmation of the mutated CHS coding region was by PCR

using the appropriate confirmation primers (Table 3.2) and by sequencing.

Yeast Two -Hybrid Analysis

The mutated CHS coding regions were amplified using the primers listed in Table

3.2, digested with XhoI and NotI, and then cloned into the yeast 2-hybrid vectors, pBI-

880 (containing the Gal41-147 DNA-binding domain) and pBI-881 (containing the Gal4768-

881 transactivation domain) (Kohalmi et al., 1998). The ligation ractions were used to

transform E.coli DH10b cells by electroporation (Dower et al., 1988). Clones were

screened by PCR using the same primers used for amplification. The recombinant

plasmids were isolated using an alkaline lysis miniprep method (Ausubel et al., 1989;

Birnboim and Doly, 1979) and then used to transform Saccharomyces cerevisiae HF7c

cells (Feilotter et al., 1994) in pairwise combinations using the method of Gietz and

Schiestl (1995). Double transformants were selected on leu-, trp- synthetic dextrose

minimal medium (Ausubel et al., 1989). Interactions between fusion proteins were

assayed after 4 d at 30º C on leu-, trp-, his- minimial medium containing 5 mM 3-

aminotriazole (Sigma).

Page 102: Structural Characterization of the Flavonoid Enzyme Complex

87

Small-Angle Neutron Scattering

Protein samples for SANS were purified as described above, then diluted to the

desired concentration (0.75, 1, 2, or 4 mg/ml) in FPLC buffer containing 5% glycerol.

For 2H2O studies, the samples were diluted to the appropriate concentration in FPLC

buffer prepared in 2H2O (Cambridge Isotopes) from stock solutions, which were made in

H2O, (resulting in approximately 9.5% H2O in the final solution), and dialyzed overnight

using Slide-A-Lyzer 10,000 MWCO cassettes (Pierce) at 4ºC against 2H2O buffer with

one buffer change at approximately 3 h.

SANS experiments were performed on the CHRNS NG3 30-m SANS (Glinka et

al., 1998) at the NIST Center for Neutron Research (NCNR). The neutron wavelength, ?,

was 5.50 Å with a wavelength spread, ??/?, of 0.150. Sample-to-detector distances of

5.0 and 1.5 m were used to obtain a wavevector transfer range of 0.009 Å-1 = Q = 0.3335

Å-1, where Q = 4psin(?)/? and 2? is the scattering angle. Data were collected at 25ºC

with total protein concentrations ranging from 0.75 to 4 mg/ml.

Neutron intensities were corrected for scattering from the buffers and from

background neutrons. Data were placed on an absolute scale by normalizing the scattered

intensity to the incident beam flux. Finally, the data were radially averaged to produce

scattered intensity, I(Q), vs Q curves. Distance distribution functions, P(r), were

calculated using the GNOM program (Semenyuk and Svergun, 1991), along with the

radius of gyration, Rg, forward scattered intensity, I(0), and the maximum dimension,

Dmax. Rg and I(0) values calculated from Guinier plots (Guinier and Fournet, 1955) were

used as a guide to determine the best P(r) function.

Page 103: Structural Characterization of the Flavonoid Enzyme Complex

88

In Silico Docking

The CHS and CHI structures were docked using the geometric-hashing

methodology within the program Protein-Protein Dock (PPD) (Norel et al., 1994; Norel

et al., 1995; Norel et al., 1999; Norel et al., 2001). PPD performs rigid body docking of

one molecule into another via three steps: generation of a molecule’s critical points by a

molecular surface calculation, docking of the molecules using the critical points through

a matching algorithm, and evaluation of the solution. Intensities for each of the docked

complexes were scored against the SANS 1 mg/ml data set using the program CRYSON

(Svergun et al., 1995).

Molecular Dynamics Simulations

The CHS-CHI complex that best fit the SANS data was refined by molecular

dynamics simulations. The structure was first solvated with water, and sodium ions were

added to give the entire system a net charge of zero. Minimization of the solvent and

counterions was performed for 200 steps using Amber8 using steepest descent (Pearlman

et al., 1995). Molecular dynamics was then performed using the Sander module of

Amber8 for a total of 5 ns with a time step of 2.0 fs and a structure being collected every

1 ps. The dynamics simulations were performed using 16 processors on Virginia Tech’s

Laboratory for Advanced Scientific Computing and Applications Linux cluster. The

resulting structure was again compared against the SANS data as described above.

Interatomic distances, hydrogen bonding patterns, and root mean square deviation

Page 104: Structural Characterization of the Flavonoid Enzyme Complex

89

(RMSD) values were determined using the Carnal program within Amber7 (Pearlman et

al., 1995). An average model was generated by coordinate averaging of the last 100 ps of

the dynamics simulation. The solvent and sodium ions were manually stripped from the

average structure and the system was minimized again as described above. The final,

minimized model was again compared against the SANS H20 and 2H2O data.

Multiple Sequence Alignments of CHS and CHI

Multiple sequence alignments of CHS and CHI from different Arabidopsis

ecotypes, members of the Brassicaceae family, and diverse plant species were generated

using ClustalW (Thompson et al., 1994) using the Megalign program in DNAStar

(Madison, WI). The ecotypes included in the CHS alignment are as follows, with

abbreviations and accession numbers given in parentheses: Arabidopsis thaliana

Landsberg erecta (Ler P13114), Maerkt (Mrk-0 CAF04429), Kashmir (Kas-1

CAF04430), Llagostera (Ll-0 CAF04432), Columbia (Col-0 CAF04433), and Lipowiec

(Lip-0 CAF04434). The CHI alignment included the following ecotypes: Landsberg

erecta (Ler AAA32766), Canary (Can-0 CAB94969), Champex (Cha-0 CAB94970),

Condhara (CAB94972), Granollers (Gran1 CAB94973), Graponne (Grap1 CAB94974),

Ibel Tazekka (Ita-0 CAB94975), Kashmir (Kas-1 CAB94976), Mechtshausen (Me-0

CAB94977), Muehlen (Mh-0 CAB94978), Monte (Mr-0 CAB94979), Wilp (Wlp2

CAB94982), Cape Verdi (Cvi-0 CAB94983), Eden (Ed1 CAB94984), Graz (Gr-5

CAB94985), Lulep (Lu1 CAB94986), Perm (Per-1 CAB94987), Richmond (Ri-0

CAB94988), Turk Lake (Tul-0 CAB94989), Yosemite (Yo-0 CAB94990), Rubezhnoe

Page 105: Structural Characterization of the Flavonoid Enzyme Complex

90

(Rub-1 CAB94991), Ruds Vedby (RV8 CAD42211), and Tvärminne (TV5 CAD42220).

The sequence alignment for CHS from the Brassicaceae were A. thaliana (P13114),

Aethionema grandiflora (AAF23557), A. lyrata (CAF04461), A. griffithiana

(AAF23568), A. halleri (CAF04425), Arabis alpina (Q9SEP4), Cardamine amara

(Q9SEP2), Aubrieta deltoidea (AAF23584), Barbarea vulgaris (AAF23583), Arabis

turrita (AAF23582), Capsella rubella (AAF23581), Arabis procurrens (AAF23580),

Arabis pauciflora (AAF23577), Arabis parishii (AAF23576), Arabis lyallii (AAF23574),

Arabis lignifera (AAF23573), Arabis jacquinii (AAF23572), Arabis hirsuta

(AAF23571), Halimolobos perplexa (AAF23569), Arabis glabra (AAF23566), Arabis

fendleri (AAF23565), Arabis drummondii (AAF23564), and Arabis blepharophylla

(AAF23562). Sequences from Brassicaceae CHI family members were A. thaliana

(AAA32766) and A. lyrata (Q9LKC3). The multi-species aligment included CHS from

A. thaliana (accession number P13114), china aster (CAA91930), orange CHS1

(BAA81664), clove pink (CAA91923), morning glory CHS1 (BAA75310) and CHS2

(BAA36224), kudzu vine (BAA01075), maize (P24825), alfalfa CHS2 (P30074), petunia

(AAF60297), radish (AAB87072), rice (CAA61955), and grape (BAA31259). The

analysis using CHI proteins included sequences from the following species: A. thaliana

(AAA32766), china aster (CAA91921), orange (BAA36552), clove pink (CAA06202),

morning glory (AAB86474), kudzu vine (BAA09795), maize (S41570), alfalfa CHI1

(P28012), petunia CHI1 (ISPJCB) and CHI2 (ISPJA1), radish (AAB87071), rice

(AAO65886), and grape (CAA53577).

Page 106: Structural Characterization of the Flavonoid Enzyme Complex

91

Results

CHS Structure

In order to solve the crystal structure of Arabidopsis CHS, recombinant protein

was expressed, purified, assayed for activity, and crystallized. The structure was solved

at 1.71 Å resolution with molecular replacement using a homology model for

Arabidopsis CHS (Saslowsky et al., 2000). The Arabidopsis CHS crystal structure

(Figure 3.2A) bears substantial similarity to that of Medicago sativa CHS2 (pdb id: 1BI5)

(Ferrer et al., 1999). When overla id, the Arabidopsis and Medicago CHS structures are

nearly identical, with the exception of a four amino acid insertion at the N-terminus of the

Arabidopsis protein (Figure 3.2B). The RMSD between the two proteins is 0.66 Å,

indicating very few structural differences. Arabidopsis CHS maintains the same aßaßa

core motif as other enzymes in the plant type III polyketide synthase superfamily (Ferrer

et al., 1999). The amino acids involved in catalysis, substrate specificity, and CoA

binding, are all conserved and in a similar location as in the Medicago CHS structure.

Experimental Evidence for a Role of Specific Residues in CHS in Assembly of the

Flavonoid Enzyme Complex

Ma et al. (2003) were able to identify phenylalanine and tryptophan residues that

are grouped together on the surface of a protein as important for interactions through

analysis of binding sites using the structure comparison algorithm MUSTA (Leibowitz et

Page 107: Structural Characterization of the Flavonoid Enzyme Complex

92

al., 2001a; Leibowitz et al., 2001b). CHS residues F293, F305, W306, and W373 were

selected for analysis as these residues are surface-exposed on the same area of CHS.

Interestingly, each of these residues is highly conserved among plant type III polyketide

synthases. Both F293 and F305 are strictly conserved across 15 representative members

of this enzyme super-family, while W306 is conserved in all but soybean, and W373 has

substitutions in CHS and two stilbene synthases from pine.

The effects of these amino acid substitutions on the ability of CHS to interact with

other flavonoid enzymes were assayed using a yeast 2-hybrid system after alanine

substitution using site-directed mutagenesis. This system had previously been used by

Burbulis and Winkel-Shirley (1999) to identify interactions among wild-type CHS, CHI,

and dihydroflavonol 4-reductase (DFR). In this experiment, each of the CHS variants

was fused to either the binding domain (BD) or the transcriptional activation domain

(TA) of the yeast GAL4 protein and tested in all pairwise combinations with Arabidopsis

CHI and DFR. The wild-type constructs were tested to reproduce the original

experiments, showing interactions between the Gal4-TA CHS and Gal4-BD CHI

constructs as well as between the Gal4-TA DFR and the Gal4-BD CHS constructs (Table

3.3). Remarkably, each of the four introduced changes appeared to disrupt both of these

interactions. These results suggest that F293, F305, W306, and W373 define a surface

domain that functions in the interactions of CHS with several other flavonoid enzymes.

It is also possible that these substitutions have indirect effects on other surface domains

involved in protein-protein interactions. However, there are no indications of significant

structural changes from homology modeling and molecular dynamics simulations (data

not shown).

Page 108: Structural Characterization of the Flavonoid Enzyme Complex

93

Small-Angle Neutron Scattering

SANS was used to examine the structure of the CHS and CHI bienzyme system in

solution. SANS data were collected for CHS and CHI at 0.75 and 1 mg/ml in either a

H2O or 2H2O buffer. The resulting scattered intensity curves are shown in Figure 3.4.

The data were normalized to a scattered intensity of I(Q = 0) = 1.0 so that the shapes of

the curves can be more easily compared. In this case, data were collected to a Q value of

0.3335 Å-1, which translates into a resolution of approximately 33 Å using the relation D

= 2p/Qmax (Jacrot et al., 1976).

Neutron scattering parameters calculated from the scattered intensity curves using

both Guinier and distance distribution function, P(r), analyses are shown in Table 3.4.

Although there were slight differences, the data from the P(r) analyses generally

concurred with the data from the Guinier approximation. Data from the P(r) analysis are

considered to be more reliable in comparisons to model structures because P(r) analysis

uses the entire Q range to calculate the radius of gyration and forward scattering intensity

[I(0)] values.

Interestingly, there seemed to be a difference in the makeup of the CHS-CHI

complex in H2O and 2H2O. From I(0), it is possible to calculate a rough molecular

weight estimate by the following equation:

2)(

)0(

vc

NIeightMolecularW A

ρ∆= Equation 1

where c is the concentration in g/cm3, ∆? is the contrast [scattering length density of the

protein minus that of the solvent (for protein in 2H2O: -3.40*1010 cm-2 and for protein in

Page 109: Structural Characterization of the Flavonoid Enzyme Complex

94

H2O: 2.36*1010 cm-2)], v bar is the partial specific volume (0.73 cm3/g ) and I(0) is the

intensity at Q=0 (cm-1) from the normalized scattering data. When the molecular weight

was calculated for the 1 mg/ml data sets, there was a significant difference in the

predicted molecular weight of the complex in H2O and 2H2O (Table 3.4). It appears that

two CHI monomers interact with each CHS dimer in H2O, whereas only one CHI

monomer interacts with the dimer in 2H2O. 2H2O easily exchanges from H on solvent-

exposed carbons and nitrogens with 2H. This result suggests that the presence of 2H on

the protein interferes with the interaction, possibly by disrupting either hydrogen bonds

or electrostatic bonds between CHS and CHI.

In Silico Docking and Molecular Dynamics Analysis of the CHS-CHI Complex

The interaction of CHS and CHI was modeled using the crystal structures of

Arabidopsis CHS and CHI (J.P.N. unpublished results) with the program Protein-Protein

Dock (Norel et al., 1994; Norel et al., 1995; Norel et al., 1999; Norel et al., 2001), which

performs a rigid body docking of the two molecules based on the surface shape and

charge complementarily. This process produced 55 potential docking solutions that were

compared to the SANS intensity data using CRYSON (Svergun et al., 1995). The

proposed complex model that best fit the SANS data (Figure 3.5) was then subjected to

molecular dynamics simulations to allow for any structural changes resulting from the

interaction. The interaction of CHI with CHS in this model did not obscure either the

substrate channel or the surface exposed-residues on CHS involved in the binding of

Page 110: Structural Characterization of the Flavonoid Enzyme Complex

95

substrate. Interestingly, the location of the CHS-CHI interface did not include the

residues that were analyzed in the yeast 2-hybrid experiments.

The molecular dynamics simulation did not identify any significant structural

changes at the interaction interface, as indicated by a RMSD value of 2.33 Å for the

entire system. In particular, two alpha helicies on CHI (a3 and a7) and one on CHI (a11)

remained relatively static (Table 3.5). The dynamics simulations identified three salt

bridges that are present between charged residues in the two enzymes: between E327 and

E328 on CHS and K18 on CHI, between R71 on CHS and E219 on CHI, and K361 of

CHS and E87 CHI (Figures 3.9 and 3.10). The distances between the alpha carbon atoms

remained fairly constant over the last 100 ps of the simulation, with standard deviations

no greater than 0.31 Å, and, in most cases, the residues moved closer to each other during

the simulation (Table 3.5). The model also predicted a hydrogen bond between T69 of

CHS and Q222 on CHI (Figures 3.9 and 3.10). To further investigate the function of

these residues in the CHS-CHI interaction, each is currently being altered to alanine for

analysis in the yeast 2-hybrid system.

CHS and CHI each exhibited one relatively mobile region throughout the

dynamics simulations. In CHS, this region corresponded to the first 13 residues, which is

solvent exposed in the model. The most mobile regions in the entire complex were sheets

ß4 and ß5 of CHI (Figure 3.6). Over the last 1 ns of simulation, the alpha carbon of

residue G53, a residue located in the turn between ß4 and ß5, moved 5.61 Å (Figure 3.7)

with a RMSD of 2.43 ± 0.682 Å. Interestingly, ß4 and ß5 help form the naringenin

binding cleft and include residues involved in the active site hydrogen bonding network

Page 111: Structural Characterization of the Flavonoid Enzyme Complex

96

(Jez et al., 2000). It is possible that the presence of bound substrate stabilizes these two

sheets.

Conservation of Residues Involved in the CHS-CHI Interaction

In order to investigate conservation of residues predicted by these experiments to

be involved in the CHS-CHI interaction, multiple sequence alignments were generated

using representative CHS and CHI sequences from different ecotypes of Arabidopsis

thaliana, members of the Brassicaceae family, and more distantly-related plant species

for which sequences of both enzymes are known (Figure 3.10). For both CHS and CHI

there is absolute conservation of these residues as well as the surrounding sequences

among proteins from the various ecotypes of Arabidopsis (Figure 3.10). However, only

two of the residues in CHS, -R71 and -E328, are absolutely conserved among members

of the Brassicaceae family. In other Brassicaceae species, CHS-E327 is an alanine.

Interestingly, CHS-K361 is conserved in 14 of the species surveyed, and the other species

contain a glutamic acid at this position, which has the opposite charge (Figure 3.10).

Unfortunately, the corresponding CHI sequence is not available for these species to

determine whether compensatory changes have occurred in the other enzyme.

Investigation of the sequences from distantly-related plant species indicated many

differences between CHS and CHI at the proposed interaction interface. Across the CHS

species, two conserved residues were observed, R71 and E328, with few non-conserved

substitutions (Figure 3.10). Among the CHI proteins, the only proposed interaction

Page 112: Structural Characterization of the Flavonoid Enzyme Complex

97

residue that appears to be conserved is E87. China aster contains the only significant

difference, with an alanine at this position; in all other species surveyed, position 87

retains a negative charge (Figure 3.10C).

Discussion

This study presents an experimental model for the interaction of CHS and CHI,

which catalyze the first two enzymatic reactions in flavonoid biosynthesis. Previous work

had implicated specific surface-exposed phenylalanine and tryptophan residues as

‘hot spots’ that are critical for protein-protein interactions (Ma et al., 2003). However,

from the data presented here, it appears that the surface-exposed phenylalanines and

tryptophans on CHS that were tested are not directly involved in the interaction between

CHS and CHI, but may play an indirect role by maintaining the structure of adjacent

domains that function in the interactions.

The model presented here encompasses only the first two enzymes of the

flavonoid pathway. However, many other questions relating to the flavonoid metabolon

remain to be answered, including the organization of the entire complex both in vitro and

in vivo as well as the effects of substrates and end products on its stability and

composition. The next step in the characterization of the flavonoid metabolon is to

determine how other flavonoid enzymes interact with the CHS-CHI bienzyme complex.

Previous studies from our laboratory indicate that DFR and F3H also interact with CHS

(Burbulis and Winkel-Shirley, 1999) as part of the flavonoid metabolon. Further

structural and SANS studies could help elucidate how these proteins interact with each

Page 113: Structural Characterization of the Flavonoid Enzyme Complex

98

other. Surface plasmon resonance refractometry is a relatively new technology that could

be used to determine the kinetics of protein association and dissociation. Amide

hydrogen/deuterium exchange-MS experiments could also give information on both the

rates of association and dissociation as well as the solvent-accessible area present on a

protein (Mandell et al., 2001). Once more is known about the complex as a whole, it

might then be possible to alter the flux of the pathway to lead to the production of novel

flower coloration or more nutritious plants.

SANS is strictly an in vitro technique, and in vivo verification of potential models

is required. Validation of a SANS-developed model can be performed by site-directed

mutagenesis and yeast two-hybrid analysis as described here. Another way to validate a

model of such a complex could be through tandem affinity purification (TAP) tagging.

Recently, TAP tagging, coupled with mass spectrometry, has been used to identify

protein complexes present in yeast (Gavin et al., 2002; Graumann et al., 2004), mammals

(Brajenovic et al., 2004; Knuesel et al., 2003), and plants (Rohila et al., 2004) and could

easily be applied to the flavonoid metabolon.

The SANS experiments presented herein investigated two proteins in an

environment that is not crowded, unlike the environment present in the cell. Future

SANS experiments should be performed in the presence of a crowding agent, such as

polyethylene glycol (PEG), to provide a crowded environment, similar to the cell. PEG

has been shown previously to serve as a suitable crowding agent in studying metabolic

flux (Rohwer et al., 1998) and interactions (Kozer and Schreiber, 2004), however, there

are no examples of PEG being used as a crowding agent for specific SANS experiments.

Page 114: Structural Characterization of the Flavonoid Enzyme Complex

99

Currently, little is known about the structure of metabolic enzyme complexes.

One of the few structural models of such organization is of enzymes involved in the citric

acid cycle (Vélot et al., 1997). The work presented here represents the first effo rt to use

SANS in the determination of the structure of a metabolic enzyme complex. SANS

coupled with in silico docking algorithms provided a powerful approach for predicting

the structure of the CHS-CHI bienzyme complex, which is now being tested

experimentally. These techniques are likely to find applications in the study of other

metabolic complexes where at least some structural information is available.

Page 115: Structural Characterization of the Flavonoid Enzyme Complex

100

Acknowledgements

The authors would like to thank Mike Austin and Marianne Bowman of the Noel lab for

assistance with CHS purification and crystallization, the staff at the NIST Center for

Neutron Research for support with the SANS experiments, and Dr. David Bevan and Jory

Zmuda Ruscio for assistance with the molecular dynamics simulations. This work was

supported by the National Science Foundation (grants MCB-9808117 and MCB-

0131010) and by a grant from the Virginia Tech Graduate Research Development Project

to CDD. Portions of this research were carried out at the Stanford Synchrotron Radiation

Laboratory, a national user facility operated by Stanford University on behalf of the U.S.

Department of Energy, Office of Basic Energy Sciences. The SSRL Structural Molecular

Biology Program is supported by the Department of Energy, Office of Biological and

Environmental Research, and by the National Institutes of Health, National Center for

Research Resources, Biomedical Technology Program. We also acknowledge the

National Institute of Standards and Technology, U.S. Department of Commerce, in

providing facilities used in this work which are supported by the National Science

Foundation under Agreement No. DMR-9986442.

Page 116: Structural Characterization of the Flavonoid Enzyme Complex

101

References Cited:

Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A.,

and Struhl, K. (1989). Current Protocols in Molecular Biology (New York, Green

Publishing Associates Wiley Interscience).

Beeckmans, S., Khan, A. S., Van Driessche, E., and Kanarek, L. (1994). A specific

association between the glyoxylic-acid-cycle enzymes isocitrate lyase and malate

synthase. Eur J Biochem 224, 197-201.

Beeckmans, S., Van Driessche, E., and Kanarek, L. (1989). The visualization by affinity

electrophoresis of a specific association between the consecutive citric acid cycle

enzymes fumarase and malate dehydrogenase. Eur J Biochem 183, 449-454.

Birnboim, H. C., and Doly, J. (1979). A rapid alkaline extraction procedure for screening

recombinant plasmid DNA. Nucleic Acids Res 7, 1513-1523.

Brajenovic, M., Joberty, G., Kuster, B., Bouwmeester, T., and Drewes, G. (2004).

Comprehensive proteomic analysis of human Par protein complexes reveals an

interconnected protein network. J Biol Chem 279, 12804-12811.

Brunger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse-Kunstleve,

R. W., Jiang, J. S., Kuszewski, J., Nilges, M., Pannu, N. S., et al. (1998).

Crystallography & NMR system: A new software suite for macromolecular

structure determination. Acta Crystallogr D Biol Crystallogr 54 ( Pt 5), 905-921.

Bu, Z., Wang, L., and Kendall, D. A. (2003). Nucleotide binding induces changes in the

oligomeric state and conformation of Sec A in a lipid environment: a small-angle

neutron-scattering study. J Mol Biol 332, 23-30.

Page 117: Structural Characterization of the Flavonoid Enzyme Complex

102

Burbulis, I. E., and Winkel-Shirley, B. (1999). Interactions among enzymes of the

Arabidopsis flavonoid biosynthetic pathway. Proc Natl Acad Sci U S A 96,

12929-12934.

Capel, M. S., Engelman, D. M., Freeborn, B. R., Kjeldgaard, M., Langer, J. A.,

Ramakrishnan, V., Schindler, D. G., Schneider, D. K., Schoenborn, B. P., Sillers,

I. Y., and et al. (1987). A complete mapping of the proteins in the small ribosomal

subunit of Escherichia coli. Science 238, 1403-1406.

Collaborative Computational Project, N. (1994). The CCP4 suite: Programs for protein

crystallography. Acta Cryst D50, 760-763.

Dower, W. J., Miller, J. F., and Ragsdale, C. W. (1988). High efficiency transformation

of E. coli by high voltage electroporation. Nucleic Acids Res 16, 6127-6145.

Feilotter, H. E., Hannon, G. J., Ruddell, C. J., and Beach, D. (1994). Construction of an

improved host strain for two hybrid screening. Nucleic Acids Res 22, 1502-1503.

Ferrer, J.-L., Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (1999). Structure

of chalcone synthase and the molecular basis of plant polyketide biosynthesis. Nat

Struc Biol 6, 775-784.

Forkmann, G., and Martens, S. (2001). Metabolic engineering and applications of

flavonoids. Curr Opinion Biotechnol 12, 155-160.

Fritsch, H., and Grisebach, H. (1975). Biosynthesis of cyanidin in cell cultures of

Haplopappus gracilis. Phytochemistry 14, 2437-2442.

Gavin, A. C., Bosche, M., Krause, R., Grandi, P., Marzioch, M., Bauer, A., Schultz, J.,

Rick, J. M., Michon, A. M., Cruciat, C. M., et al. (2002). Functional organization

Page 118: Structural Characterization of the Flavonoid Enzyme Complex

103

of the yeast proteome by systematic analysis of protein complexes. Nature 415,

141-147.

Gietz, R. D., and Schiestl, R. H. (1995). Transforming yeast with DNA. Methods Mol

Cell Biol 5, 255-269.

Glinka, C. J., Barker, J. G., Hammouda, B., Krueger, S., Moyer, J. J., and Orts, W. J.

(1998). The 30-meter small angle neutron scattering instruments at the National

Institute of Standards and Technologies. J Appl Cryst 31, 430-445.

Gontero, B., Cárdenas, M. L., and Ricard, J. (1988). A functional five-enzyme complex

of chloroplasts involved in the Calvin cycle. Eur J Biochem 173, 437-443.

Graumann, J., Dunipace, L. A., Seol, J. H., McDonald, W. H., Yates, J. R., 3rd, Wold, B.

J., and Deshaies, R. J. (2004). Applicability of tandem affinity purification

MudPIT to pathway proteomics in yeast. Mol Cell Proteomics 3, 226-237.

Green, D. E. (1958). Studies in organized enzyme systems. Harvey Lect 52, 177-227.

Guex, N., and Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an

environment for comparative protein modeling. Electrophoresis 18, 2714-2723.

Guinier, A., and Fournet, G. (1955). Small Angle Scattering of X-rays (New York, Wiley

Press).

Havsteen, B. H. (2002). The biochemistry and medical significance of the flavonoids.

Pharmacol Ther 96, 67-202.

Hrazdina, G., and Wagner, G. J. (1985). Compartmentation of plant pheno lic compounds;

sites of synthesis and accumulation. Annu Proc Phytochem Soc Europe 25, 120-

133.

Page 119: Structural Characterization of the Flavonoid Enzyme Complex

104

Hrazdina, G., Zobel, A. M., and Hoch, H. C. (1987). Biochemical, immunological, and

immunocytochemical evidence for the association of chalcone synthase with

endoplasmic reticulum membranes. Proc Natl Acad Sci U S A 84, 8966-8970.

Jacrot, B., Pfeiffer, P., and Witz, J. (1976). The structure of a spherical plant virus

(bromegrass mosaic virus) established by neutron diffraction. Philos Trans R Soc

Lond B Biol Sci 276, 109-112.

Jez, J. M., Bowman, M. E., Dixon, R. A., and Noel, J. P. (2000). Structure and

mechanism of the evolutionarily unique plant enzyme chalcone isomerase. Nat

Struct Biol 7, 786-791.

Jones, T. A., Zou, J. Y., Cowan, S. W., and Kjeldgaard, M. (1991). Improved methods for

binding protein models in electron density maps and the location of errors in these

models. Acta Crystallogr A 47, 110-119.

Knuesel, M., Wan, Y., Xiao, Z., Holinger, E., Lowe, N., Wang, W., and Liu, X. (2003).

Identification of novel protein-protein interactions using a versatile Mammalian

tandem affinity purification expression system. Mol Cell Proteomics 2, 1225-

1233.

Koch, M. H., Vachette, P., and Svergun, D. I. (2003). Small-angle scattering: a view on

the properties, structures and structural changes of biological macromolecules in

solution. Q Rev Biophys 36, 147-227.

Kohalmi, S. E., Reader, L. J. V., Samach, A., Nowak, J., Haughn, G. W., and Crosby, W.

L. (1998). Identification and characterization of protein interactions using the

yeast 2-hybrid system. Plant Mol Biol Manual M1, 1-30.

Page 120: Structural Characterization of the Flavonoid Enzyme Complex

105

Kozer, N., and Schreiber, G. (2004). Effect of crowding on protein-protein association

rates: fundamental differences between low and high mass crowding agents. J

Mol Biol 336, 763-774.

Krueger, S., Gregurick, S. K., Zondlo, J., and Eisenstein, E. (2003). Interaction of GroEL

and GroEL/GroES complexes with a nonnative subtilisin variant: a small-angle

neutron scattering study. J Struct Biol 141, 240-258.

Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. (1993).

PROCHECK: a program to check the stereochemical quality of protein structures.

J Appl Cryst 26, 283-291.

Leibowitz, N., Fligelman, Z. Y., Nussinov, R., and Wolfson, H. J. (2001a). Automated

multiple structure alignment and detection of a common substructural motif.

Proteins 43, 235-245.

Leibowitz, N., Nussinov, R., and Wolfson, H. J. (2001b). MUSTA--a general, efficient,

automated method for multiple structure alignment and detection of common

motifs: application to proteins. J Comput Biol 8, 93-121.

Ma, B., Elkayam, T., Wolfson, H., and Nussinov, R. (2003). Protein-protein interactions:

structurally conserved residues distinguish between binding sites and exposed

protein surfaces. Proc Natl Acad Sci USA 100, 5772-5777.

Mandell, J. G., Baerga-Ortiz, A., Akashi, S., Takio, K., and Komives, E. A. (2001).

Solvent accessibility of the thrombin-thrombomodulin interface. J Mol Biol 306,

575-589.

Muir, S. R., Collins, G. J., Robinson, S., Hughes, S., Bovy, A., De Vos, C. H. R., van

Tunen, A. J., and Verhoeyen, M. E. (2001). Overexpression of petunia chalcone

Page 121: Structural Characterization of the Flavonoid Enzyme Complex

106

isomerase in tomato results in fruit containing increased levels of flavonols.

Nature Biotechnol 19, 470-474.

Navaza, J. (1994). AMoRE: Automated Program for Molecular Replacement. Acta

Crystallogr A50, 157-163.

Norel, R., Lin, S. L., Wolfson, H. J., and Nussinov, R. (1994). Shape complementarity at

protein-protein interfaces. Biopolymers 34, 933-940.

Norel, R., Lin, S. L., Wolfson, H. J., and Nussinov, R. (1995). Molecular surface

complementarity at protein-protein interfaces: the critical role played by surface

normals at well placed, sparse, points in docking. J Mol Biol 252, 263-273.

Norel, R., Petrey, D., Wolfson, H. J., and Nussinov, R. (1999). Examination of shape

complementarity in docking of unbound proteins. Proteins 36, 307-317.

Norel, R., Sheinerman, F., Petrey, D., and Honig, B. (2001). Electrostatic contributions to

protein-protein interactions: fast energetic filters for docking and their physical

basis. Protein Sci 10, 2147-2161.

Otwinowski, Z., and Minor, W. (1997). Processing of X-Ray Diffraction Data Collected

in Oscillation Mode. Meth Enzymology 276, 307-326.

Ovadi, J., and Srere, P. A. (2000). Macromolecular compartmentation and channeling. Int

Rev Cytol 192, 255-280.

Pearlman, D. A., Case, D. A., Caldwell, J. W., Ross, W. S., Cheatham, T. A., DeBolt, S.,

Ferguson, D. M., Seibel, G., and Kollman, P. (1995). AMBER, a package of of

computer programs for applying molecular mechanics, a normal mode analysis,

molecular dynamics and free energy calculations to simulate the structural and

energetic properties of molecules. Comput Phys Commun 91, 1-41.

Page 122: Structural Characterization of the Flavonoid Enzyme Complex

107

Pelletier, M. K., and Shirley, B. W. (1996). Analysis of flavanone 3-hydroxylase in

Arabidopsis seedlings. Coordinate regulation with chalcone synthase and

chalcone isomerase. Plant Physiol 111, 339-345.

Robinson, J. B. J., Inman, L., Semegi, B., and Srere, P. A. (1987). Further

characterization of the Krebs tricarboxylic acid cycle metabolon. J Biol Chem

262, 1786-1790.

Rohila, J. S., Chen, M., Cerny, R., and Fromm, M. E. (2004). Improved tandem affinity

purification tag and methods for isolation of protein heterocomplexes from plants.

Plant J 38, 172-181.

Rohwer, J. M., Postma, P. W., Kholodenko, B. N., and Westerhoff, H. V. (1998).

Implications of macromolecular crowding for signal transduction and metabolite

channeling. Proc Natl Acad Sci USA 95, 10547-10552.

Santos, M. O., Crosby, W. L., and Winkel-Shirley, B. (2004). Modulation of flavonoid

metabolism in Arabidopsis using a phage-derived antibody. Molecular Breeding,

in press.

Saslowsky, D., and Winkel-Shirley, B. (2001). Localization of flavonoid enzymes in

Arabidopsis roots. Plant J 27, 37-48.

Saslowsky, D. E., Dana, C. D., and Winkel-Shirley, B. (2000). An allelic series for the

chalcone synthase locus in Arabidopsis. Gene 255, 127-138.

Semenyuk, A. V., and Svergun, D. I. (1991). J Appl Cryst 24, 537-540.

Stafford, H. A. (1974a). The metabolism of aromatic compounds. Ann Rev Plant Physiol

25, 459-486.

Page 123: Structural Characterization of the Flavonoid Enzyme Complex

108

Stafford, H. A. (1974b). Possible multi-enzyme complexes regulating the formation of

C6-C3 phenolic compounds and lignins in higher plants. Rec Adv Phytochem 8,

53-79.

Stafford, H. A. (1990). Flavonoid Metabolism (Boca Raton, CRC Press).

Svergun, D. I., Baberato, C., and Koch, M. H. (1995). CRYSOL - a program to evaluate

X-ray solution scattering of biological macromolecules from atomic coordinates. J

Appl Cryst 28, 768-773.

Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994). CLUSTAL W: improving the

sensitivity of progressive multiple sequence alignment through sequence

weighting, position-specific gap penalties and weight matrix choice. Nucleic

Acids Res 22, 4673-4680.

Velev, O. D., Kaler, E. W., and Lenhoff, A. M. (1998). Protein interactions in solution

characterized by light and neutron scattering: comparison of lysozyme and

chymotrypsinogen. Biophys J 75, 2682-2697.

Vélot, C., Mixon, M. B., Teige, M., and Srere, P. A. (1997). Model of a quinary structure

between Krebs TCA cycle enzymes: a model for the metabolon. Biochemistry 36,

14271-14276.

Wagner, G. J., and Hrazdina, G. (1984). Endoplasmic reticulum as a site of

phenylpropanoid and flavonoid metabolism in Hippeastrum. Plant Physiol 74,

901-906.

Welch, G. R. (1977). On the role of organized multienzyme systems in cellular

metabolism: a general synthesis. Prog Biophys Mol Biol 32, 103-191.

Page 124: Structural Characterization of the Flavonoid Enzyme Complex

109

Winkel-Shirley, B. (2002). Biosynthesis of flavonoids and effects of stress. Curr Opin

Plant Biol 5, 218-223.

Yu, O., Shi, J., Hession, A. O., Maxwell, C. A., McGonigle, B., and Odell, J. T. (2003).

Metabolic engineering to increase isoflavone biosynthesis in soybean seed.

Phytochemistry 63, 753-763.

Page 125: Structural Characterization of the Flavonoid Enzyme Complex

110

Table 3.1: Crystallographic data Space Group P21 Unit Cell dimensions (Å) a = 55.06, b = 139.27, c = 109.60 Wavelength (Å) 1.08 Resolution (Å) 29.387 - 1.710 Unique reflections 152185

Page 126: Structural Characterization of the Flavonoid Enzyme Complex

111

Table 3.2: Primers used for PCR and site directed mutagenesis for yeast 2-hybrid constructions A: Mutagenesis Primers (sections underlined indicate the location of the mutation): F293A (sense): 5'-GAAGAGTCTAGACGAAGCGGCAAAACCTTTGGGGATAAGTGAC-3' F293A (antisense): 5'-GTCACTTATCCCCAAAGGTTTTGCCGCTTCGTCTAGACTCTTC-3' F305A (sense): 5'-TGACTGGAACTCCCTCGCATGGATAGCCCACCCTG-3' F305A (antisense): 5'-CAGGGTGGGCTATCCATGCGAGGGAGTTCCAGTCA-3' W306A (sense): 5'-CTGGAACTCCCTCTTCGCAATAGCCCACCCTGGAG-3' W306A (antisense): 5'-CTCCAGGGTGGGCTATTGCGAAGAGGGAGTTCCAG-3' W373A (sense): 5'-CAGGAGAAGGGTTGGAGGCAGGTGTCTTGTTTGGTTTCG-3' W373A (antisense): 5'-CGAAACCAAACAAGACACCTGCCTCCAACCCTTCTCCTG-3' B: CHS amplification primers for yeast 2-hybrid constructs: CHS forward primer: 5’-CCGCTCGAGCATGGTGATGGCTGGTGC-3’ CHS reverse primer: 5’-AGCGGCCGCTTAGAGAGGAACGCTGTG-3’ C: Confirmation primers: F293A confirmation: 5’-CCCCAAAGGTTTAGC-3’ F305A confirmation: 5’-GGGTGGGCTATCCATGC-3’ W306A confirmation: 5’-GTGGGTGGGCTATCGC-3’ W373A confirmation: 5’-CCAAACAAGACACCCGC-3’

Page 127: Structural Characterization of the Flavonoid Enzyme Complex

112

Table 3.3: Yeast 2-hybrid analysis of amino acid substitutions. Gal4-DB construct Gal4-TA construct

CHS CHI DFR CHSF239A CHSF305A CHSW306A CHSW373A

CHS +/- +/- +++ N/T N/T N/T N/T CHI +++ +/- - - - - - DFR +/- +++ +/- - - - - CHSF293A N/T - - N/T N/T N/T N/T CHSF305A N/T - - N/T N/T N/T N/T CHSW306A N/T - - N/T N/T N/T N/T CHSW373A N/T - - N/T N/T N/T N/T

+++ : Growth of yeast cells +/-, - : No growth of yeast cells N/T: Not tested

Page 128: Structural Characterization of the Flavonoid Enzyme Complex

113

Table 3.4 Neutron scattering parameters for CHS and CHI in solution Gunier analysis P(r) Analysis Rg (Å) I(0) Rg (Å) I(0) CHS/CHI, H2O (1 mg/ml) 28.57 ± 0.827 0.0838 27.37 ± 0.393 0.0846 ± 0.0097 CHS/CHI, H2O (0.75 mg/ml) 30.99 ± 1.188 0.0380 27.70 ± 0.427 0.0336 ± 0.0084 CHS/CHI, D2O (1 mg/ml) 28.97 ± 0.242 0.120 26.67 ± 0.103 0.105 ± 0.0044 CHS/CHI, D2O (0.75 mg/ml) 28.47 ± 0.330 0.0615 26.80 ± 0.139 0.0599 ± 0.0045 Calculated molecular weigh of the CHS/CHI complexa

CHS/CHI, H2O (1 mg/ml) 174 kDa CHS/CHI, 2H2O (1 mg/ml) 117 kDa Predicted weight for complex composed of 1 CHS homodimer 140 kDa and 2 CHI monomers Predicted weight for complex composed of 1 CHS homodimer 113 kDa and 1 CHI monomer aMolecular weights from SANS data were calculated using equation 1.

Page 129: Structural Characterization of the Flavonoid Enzyme Complex

114

Table 3.5: RMSD and distance measurements of secondary structure and residues proposed to be involved in the CHS-CHI interaction. (A): RMSD measurements for secondary structural elements involved in the CHS-CHI interaction. CHS a11 CHS ß13 CHS ß3 CHI a7 CHI a3 CHI ß1 CHS a1 Average of last 100 ps

1.72 (0.08)

0.911 (0.07)

1.22 (0.06)

1.04 (0.04)

1.75 (0.09)

0.927 (0.07)

3.32 (0.13)

Average of last 1 ns

1.56 (0.13)

0.852 (0.10)

1.34 (0.14)

0.991 (0.06)

1.56 (0.18)

0.934 (0.11)

3.43 (0.23)

(B): Distance averages of proposed residues involved in the CHS-CHI interaction. CHS residue # and atom type

E327 CD

E327 CA

E328 CD

E328 CA

T69 OG1

T69 CA

R71 NH1

R71 CA

K361 NZ

K361 CA

K361 NZ

K361 CA

CHI residue # and atom type

K18 NZ

K18 CA

K18 NZ

K18 CA

Q222 CD

Q222 CA

E218 CD

E218 CA

T86 OG1

T86 CA

E87 OE2

E87 CA

Average over last 100 ps of simulation

3.63 (0.31)

9.00 (0.29)

3.31 (0.13)

9.77 (0.29)

5.12 (0.42)

8.19 (0.26)

3.98 (0.37)

10.6 (0.31)

2.88 (0.11)

9.90 (0.20)

2.78 (0.12)

11.3 (0.19)

Page 130: Structural Characterization of the Flavonoid Enzyme Complex

115

COSCoA

OHOH

COOH

NH2

Fatty Acid Metabolism

phenylalanine 4-coumaroyl-CoA

General Phenylpropanoid Pathway

3

Malonyl Co-A

+

CHS

OH

O

OHOH

OH

tetrahydroxychalcone

OH

O

OOH

OH

naringenin

CHI

OH

O

OOH

OH

OH

F3’H

F3H F3H OH

O

OOH

OH

OH

OH

O

OOH

OH

OH

OH

eriodictyol

DFR

F3’H

OH

OH

OOH

OH

OH

OH

OH

O+OH

OH

OH

OH

ANS

Anthocyanins

3GT, 5GT

ANR (BAN)

Condensed Tannins (proanthocyanidins)

OH

OOH

OH

OH

OH

LAR

OH

OOH

OH

OH

R

O

FLS

FLS

Flavonol glycosides

dihydrokaempferol dihydroquercetin

leucopelargonidin

leucocyanidin

cyanidin

Figure 3.1: Schematic of the flavonoid pathway in Arabidopsis thaliana . The pathway creates a diverse set of compounds used in UV protection, signaling, defense, and pigmentation. CHS represents that first committed step in the pathway, where 4-coumaroyl-CoA (from phenylpropanoid biosynthesis) and 3 units of malonyl-CoA (from fatty acid metabolism) are used as substrates. Structures in the figure represent the three enzymes in the pathway that crystal structures have been solved for, CHS (pdb id:1BI5) and CHI (pdb id:1EYQ) (both from Medicago sativa), and anthocyanidin synthase (ANS) (pdb id:1GP6) (from Arabidopsis thaliana). Structures in orange are homology models of dihydroflavonol 4-reductase (DFR), flavanone 3-hydroxylase (F3H), flavonoid 3’ hydroxylase (F3’H), and flavonol synthase (FLS). Other enzyme names are abbreviated as follows: anthocyanidin reductase (ANR), cinnamate 4-hydroxylase (C4H), p-coumarate:CoA ligase (4CL), leucoanthocyanidin reductase (LAR), phenylalanine ammonia-lyase (PAL). Arrows labeled in grey indicate branches of the pathway not present in Arabidopsis.

HOOC

COSCoA

Isoflavonoids

CHS/CHR

PAL C4H 4CL

3GT, RT

Page 131: Structural Characterization of the Flavonoid Enzyme Complex

116

Figure 3.2: CHS crystal structure. A. Structure of the CHS dimer. Each monomer is labeled in a different color. B. Overlay of Medicago sativa CHS2 (gray) with Arabidopsis thaliana CHS (purple).

A. B.

Page 132: Structural Characterization of the Flavonoid Enzyme Complex

117

Figure 3.3: Residues on CHS tested for effects on interactions with other flavonoid enzymes by yeast 2-hybrid analysis. The front (a) and side (b) views. The surface exposed phenylalanines and tryptophans selected for alanine substitutions are labeled in blue and the active site residues are labeled in red.

A. B.

Page 133: Structural Characterization of the Flavonoid Enzyme Complex

118

-0.03

-0.02

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0 0.05 0.1 0.15 0.2 0.25 0.3

-0.02

0

0.02

0.04

0.06

0.08

0.1

0.12

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35

0

0.00002

0.00004

0.00006

0.00008

0.0001

0.00012

0.00014

0.00016

0.00018

0 10 20 30 40 50 60 700

0.00002

0.00004

0.00006

0.00008

0.0001

0.00012

0.00014

0.00016

0.00018

0 10 20 30 40 50 60 70

Dana, et al. Figure 3.4

A. B.

C. D.

Q (Å-1) Q (Å-1)

r (Å) r (Å)

I(Q

)

I(Q

)

P(r)

P(r)

Page 134: Structural Characterization of the Flavonoid Enzyme Complex

119

Figure 3.4. Normalized SANS data. The CHS/CHI data in either H2O (A) or 2H2O (B) solvent. Normalized distance distribution function, P(r), for the CHS-CHI complex in H2O (C) or 2H2O (D) solvent. The CHS-CHI data at 0.75 mg/ml is labeled as x and the CHS-CHI data at 1.0 mg/ml is labeled as ?.

Page 135: Structural Characterization of the Flavonoid Enzyme Complex

120

-0.03

-0.02

-0.01

0

0.01

0.02

0.03

0.04

0.05

0 0.05 0.1 0.15 0.2 0.25 0.3

1mg/ml H2O data

Docked Structure

I(Q)

Figure 3.5: Data fit of the Docked CHS-CHI Complex. The optimized structure was fit to the 1 mg/ml data set from the SANS analysis. Data fit was performed using CRYSON (Svergun, et al., 1995).

Q (Å-1)

Page 136: Structural Characterization of the Flavonoid Enzyme Complex

121

Figure 3.6: Proposed structure of the CHS-CHI enzyme complex. The CHS homodimer is represented in purple, CHI is in yellow, active site residues are labeled in red and the residues involoved in CoA binding are labeled in blue. (A) front view, (B) view from 45º on the left, (C) side view, and (D) top view.

A. B.

C. D.

Page 137: Structural Characterization of the Flavonoid Enzyme Complex

122

Figure 3.7: Movement of beta sheets ß3a and ß3b of CHI at 100 ps intervals over the last 500 ps of dynamics simulation. Color labels are as follows, red: 4500 ps, orange: 4600 ps, yellow: 4700 ps, green: 4800 ps, blue: 4900 ps, and violet: 5000 ps. Over the course of the dynamics simulations the residue on the loop connecting the two beta sheets, glutamine 53 moves 5.61Å.

Page 138: Structural Characterization of the Flavonoid Enzyme Complex

123

Figure 3.8: Residues proposed to participate in the CHS-CHI interaction. (a) front view and (b) top view. Residues contributed to the interaction by CHS are labeled in blue and the residues contributed by CHI are labeled in orange. In all of the structures, the active site residues are labeled in red.

A. B.

Page 139: Structural Characterization of the Flavonoid Enzyme Complex

124

Figure 3.9: Close up of the interaction interface between CHS and CHI from the (a) front and (b) top. CHS is in purple, CHI is in yellow. The residues proposed to be important in this interaction are labeled with: basic residues in blue, acidic residues in red, and polar residues are in green.

A. B.

K18

K18

E327

E328

E218

E327

E328 Q222 T69

R71

Q222

R71

T69

E218

K361

E87

T86

Page 140: Structural Characterization of the Flavonoid Enzyme Complex

125

Ler CHS …CDKSTIRKRHM…………LGLKEEKMRA…………RKSAKDGVA… 100% Mrk-0 CHS …...........…………..........………….........… 100% Kas-1 CHS …...........…………..........………….........… 100% Ll-0 CHS …...........…………..........………….........… 99.7% Col-0 CHS …...........…………..........………….........… 100% Lip-0 CHS …...........…………..........………….........… 100% Ler CHI …PAVTKLHVD…………KGKTTEELTE…………LSVAERLSQLMMK… 100% Can-0 CHI ….........…………..........………….............… 100% Cha-0 CHI ….........…………..........………….............… 99.6% Condhara CHI….........…………..........………….............… 100% Gran1 CHI ….........…………..........………….............… 100% Grap1 CHI ….........…………..........………….............… 100% Ita-0 CHI ….........…………..........………….............… 99.6% Kas-1 CHI ….........…………..........………….............… 100% Me-0 CHI ….........…………..........………….............… 100% Mh-0 CHI ….........…………..........………….............… 99.6% Mr-0 CHI ….........…………..........………….............… 100% Wlp2 CHI ….........…………..........………….............… 100% Cvi-0 CHI ….........…………..........………….............… 99.6% Ed1 CHI ….........…………..........………….............… 100% Gr-5 CHI ….........…………..........………….............… 99.6% Lu1 CHI ….........…………..........………….............… 100% Per-1 CHI ….........…………..........………….............… 99.6% Ri-0 CHI ….........…………..........………….............… 100% Tul-0 CHI ….........…………..........………….............… 100% Yo-0 CHI ….........…………..........………….............… 100% Rub-1 CHI ….........…………..........………….............… 100% RV8 CHI ….........…………..........………….............… 99.6% TV5 CHI ….........…………..........………….............… 100%

14

22

82

91

214

226

65

75

323

332

357

365

Dana, et al. Figure 3.10A

Page 141: Structural Characterization of the Flavonoid Enzyme Complex

126

Arabidopsis thaliana CHS …CDKSTIRKRHM…………LGLKEEKMRA…………RKSAKDGVA… 100% Aethionema grandiflora CHS ….V.........…………....P.....…………K..RE..AV… 89.0% Arabidopsis lyrata CHS …-...M......………….........M………….........… 98.0% Arabidopsis griffithiana CHS …....M......…………..........………….........… 99.0% Arabidopsis halleri CHS …-...M......…………..........………….........… 98.5% Arabis alpine CHS …....M......…………....A.....…………K......A.… 96.7% Cardamine amara CHS …...........…………..........…………K........… 95.7% Aubrieta deltoidea CHS …....M......…………....A...K.…………K..EE..A.… 93.2% Barbarea vulgaris CHS …....M......…………....A.....…………K........… 94.9% Arabis turrita CHS …....M......…………....A.....…………K..VE..A.… 93.4% Capsella rubella CHS …....M......…………..........………….........… 98.5% Arabis procurrens CHS …....M......…………....A.....…………K..VE..A.… 93.7% Arabis pauciflora CHS …....M......…………....A.....…………..A.E..A.… 89.4% Arabis parishii CHS …....M......…………....A.....………….........… 98.0% Arabis lyallii CHS …....M......…………....A.....………….........… 98.0% Arabis lignifera CHS …....M......…………....A.....………….........… 97.7% Arabis jacquinii CHS …....M......…………....A.....…………K..VE..A.… 93.4% Arabis hirsute CHS …....M......…………....A.....…………K..VE..A.… 93.7% Halimolobos perplexa CHS …....M......…………..........………….........… 98.7% Arabis glabra CHS …....M......…………..........………….........… 98.2% Arabis fendleri CHS …....M......…………....A.....………….........… 98.0% Arabis drummondii CHS …....M......…………....A.....………….........… 97.5% Arabis blepharophylla CHS …....M......…………....A.....…………K..VE..A.… 93.2% Arabidopsis thaliana CHI …PAVTKLHVD…………KGKTTEELTE…………LSVAERLSQLMMK… 100% Arabidopsis lyrata CHI ….S....Q..…………........S.…………..I....AK..T.… 89.4%

65

75

323

332

357

365

14

22

82

91

214

226

Dana, et al. Figure 3.10B

Page 142: Structural Characterization of the Flavonoid Enzyme Complex

127

Arabidopsis CHS …CDKSTIRKRHM…………LGLKEEKMRA…………RKSAKDGVA… 100% China Aster CHS …....M....Y.…………..........…………K...E..A.… 83.9% Orange CHS1 …....M.K..Y.…………....G..L..…………K..VEEAK.… 85.2% Clove pink CHS …....M.K..Y.…………........A.…………K..LR..AT… 79.8% Morning glory CHS …....M.K..Y.…………....P..L..…………KA.SNA.LG… 80.9% Morning glory CHS2…....M.T..Y.…………....P..L..…………KA.SN..LG… 83.5% Kudzu vine CHS …....M.T..Y.…………....P...K.…………R...EN.LK… 83.8% Maize CHS C2 …....M....Y.…………V..DKAR...…………KR..E..Q.… 82.5% Alfalfa CHS2 …....M.KR.Y.………….A..P...N.…………K..TQN.LK… 82.0% Petunia CHS ….E..M.K..y.…………....P..LK.…………KA...E.LG… 84.8% Radish CHS …....M......…………....A.....…………R..LD....… 94.9% Rice CHS …....Q....Y.…………V..DK.RM..…………KR..E..H.… 81.7% Grape CHS ….E.......Y.………….......L..…………K..IEE.K.… 85.8% Arabidopsis CHI …PAVTKLHVD…………KGKTTEELTE…………LSVAERLSQLMMK… 100% China aster CHI ….LT.G.Q.E…………....VA..LD…………Q.L.S...D..KQ… 58.6% Orange CHI ….S..E.Q.E…………....A.....…………K.L.....A.LNV… 68.0% Clove pink CHI …AEI.QIQ.E…………....AD...D…………K.LSA...K.INQ… 60.2% Morning glory CHI….C.AEVK.E…………N..SV....D…………Q.L.V...EMFIN… 55.2% Kudzu vine CHI …ATISAVQ.E…………...PS...V.…………R.L.S..PAVLSH… 47.8% Maize CHI …M.CRRWWST…………G...AD..AS…………..L.A.V.E.LA.… 56.3% Alfalfa CHI1 …ASI.AIT.E…………...SS...LE…………RCL.A..PA.LNE… 46.4% Petunia CHI …VS...VQ.E…………...SA...IH…………C.I.A.M.E.FKN… 61.4% Petunia CHI2 …VS..RM..E…………....P....D…………C.....VAE.LKK… 60.2% Radish CHI ….TAP..Q..…………E.........………….R.....A...KE… 81.0% Rice CHI …A..SEVE..…………A..SAD..AA…………R.I.A.V...LKA… 53.2% Grape CHI ….S..AVQ.E…………....V...AD…………K.L.A...E.FCK… 65.8%

14

22

82

91

214

226

65

75

323

332

357

365

Dana, et al. Figure 3.10C

Page 143: Structural Characterization of the Flavonoid Enzyme Complex

128

Figure 3.10: Multiple sequence alignments of Arabidopsis CHS and CHI protein sequences against other ecotypes of Arabidopsis thaliana (A), members of the Brassicaceae family (B), and different plant species. A dot represents the same amino acids as the Landsberg erecta ecotype of Arabidopsis thaliana. Residues proposed to be important for the interaction are shaded and dashed lines indicate the interacting pairs of residues. Boxes indicate conservative changes. Percentages to the right of the sequences refer to total amino acid identity relative to the Arabidopsis proteins.

Page 144: Structural Characterization of the Flavonoid Enzyme Complex

129

Chapter 4

Conclusions

Page 145: Structural Characterization of the Flavonoid Enzyme Complex

130

The organization of metabolism has become a focus of study over the past 40

years. In our laboratory, the Arabidopsis flavonoid biosynthetic pathway has been

developed as a model system for the study of metabolic organization. More than 30 years

ago, Helen Stafford suggested that enzymes in flavonoid biosynthesis are organized as a

metabolon (Stafford, 1974a; Stafford, 1974b), and subsequent experiments investigating

both the subcellular localization, channeling and co-fractionation (Hrazdina, 1992;

Winkel-Shirley, 1999) and the direct interactions of the enzymes (Burbulis and Winkel-

Shirley, 1999) have supported her hypothesis. This work laid the foundation for the

structural studies of the flavonoid complex described herein, the goals of which were

two-fold: 1) to structurally characterize an allelic series of the gene encoding chalcone

synthase (CHS), the first enzyme of the flavonoid pathway, using homology modeling

and molecular dynamic simulations, and 2) to determine the molecular basis of the

complex formed by interaction between CHS and the second flavonoid enzyme, chalcone

isomerase (CHI).

Computational analyses contributed significantly to this project. In silico

techniques were used to develop structural models for a series of CHS variants that had

been characterized on the gene, protein and endproduct levels (Saslowsky et al., 2000).

The single amino acid substitutions or truncation characterizing these variants result in

proteins with altered activity, stability, and dimerization ability. These computational

techniques provide an efficient way to predict the effect of these changes on local and

global protein structure without solving the crystal structure. However, this methodology

yielded few insights into the molecular basis of the temperature-sensitive variants.

Page 146: Structural Characterization of the Flavonoid Enzyme Complex

131

Coupling these in silico techniques with quantum mechanical algorithms should aid in the

determination of catalytic mechanisms leading to the engineering of proteins with altered

activities. Previous molecular dynamics and quantum mechanics studies into the reaction

mechanism of Medicago CHI demonstrated that the substrate, tetrahydroxychalcone,

undergoes significant conformational changes during catalysis to naringenin (Hur et al.,

2004). These same types of computational approaches could be applied in preliminary

engineering efforts to determine the catalytic or structural roles of specific residues.

For the first time, a three-dimensional model for the interactions between

enzymes involved in flavonoid metabolism has been developed. Using a series of both

experimental and computational techniques, we were able to determine the structure of

CHS, suggest a model for the interaction between CHS and CHI, and test the residues

predicted to be important for the interaction using site-directed mutagenesis and yeast

two-hybrid analysis. Surface plasmon resonance refractometry (SPR) showed promise as

a way to determine the binding affinities of the CHS-CHI complex (C. Dana and B.

Winkel, unpublished data) and could be expanded to other interacting enzymes in the

pathway.

The next important step will be to expand these analyses to other enzymes in the

pathway in order to obtain a complete picture of the flavonoid metabolon, both on a

biochemical and structural level. Structures for most of the other flavonoid enzymes

have been determined either through X-ray crystallography [CHS, CHI, and

anthocyanidin synthase (Wilmouth et al., 2002)] or by homology modeling [flavanone 3-

hydroxylase, flavonol synthase, and dihydroflavonol 4-reductase (C. Dana, D. Owens,

and B. Winkel, unpublished data)] and could be used in SANS studies to piece together

Page 147: Structural Characterization of the Flavonoid Enzyme Complex

132

the overall structure of the complex. However, flavanone 3’-hydroxylase (F3’H), the

proposed membrane anchor, would be more difficult to include in this type of analysis

due to its N-terminal transmembrane domain. An alternative approach that may be useful

in this situation is neutron reflectivity, a technique similar to SANS that uses neutrons to

probe membrane-associated molecules. As with SANS, a structural picture could be

gained by docking the protein structures using in silico algorithms followed by molecular

dynamics simulations. Briefly, F3’H would be expressed in mammalian cells and

purified with the transmembrane domain incorporated into a membrane bilayer. The

purified protein could then be subjected to neutron reflectivity to confirm initial structure

of F3’H to that of a previously generated homology model (C. Dana and B. Winkel,

unpublished data). Adding the other flavonoid enzymes sequentially to the immobilized

F3’H could then be used to help build a general model of the metabolon. There are, as

yet, no examples of neutron reflectivity being used in this manner. However, a new

instrument at the NIST center for neutron research, the advanced neutron

diffractometer/reflectometer, has been designed to be used for investigations into the

structure of membrane-bound proteins as well as biological membrane structure. This

type of analysis could lead to the definition of the structure of the complex as well as the

channels available for substrate movement between active sites and determining the

molecular basis of the interaction between proteins in the metabolon.

The experiments presented herein represent a significant leap forward in the

understanding of the structure of the flavonoid metabolon. The structure of the CHS-CHI

provides, for the first time, details about the interfaces required for the interaction

between these two cooperating proteins. Once the residues required for the interaction

Page 148: Structural Characterization of the Flavonoid Enzyme Complex

133

are verified by in vitro assays, it might then be possible to start metabolic engineering to

lead to novel flavonoid production in Arabidopsis, such as the production of specific

pigments. To rationally design a change in pathway flux or endproduct levels, it is

imperative that the structure and biochemistry of the remaining parts of the pathway be

determined. This information could lead to eventual engineering efforts of the flavonoid

pathway in other plants to produce more nutritious plants or flowers with novel

pigmentation.

The CHS-CHI bienzyme model helped to establish SANS as a technology

through which the structure of a metabolic complex can be determined. To our

knowledge, this investigation was the first application of SANS in the study of metabolic

complex structure and assembly, for which little is known. Our findings suggest that

SANS offers a powerful approach in characterizing multi-enzyme systems and should

have applications in other systems where little is known about the organization of the

complex, adding to the paltry knowledge of metabolon organization. The data gained

from these experiments can also be used to gain more information on the structure of

metabolic complexes and then later, in metabolic engineering projects to alter flux within

a pathway in such a way to alter substrate flux or create novel endproducts beneficial to

animals.

Page 149: Structural Characterization of the Flavonoid Enzyme Complex

134

References Cited:

Burbulis, I. E., and Winkel-Shirley, B. (1999). Interactions among enzymes of the

Arabidopsis flavonoid biosynthetic pathway. Proc Natl Acad Sci U S A 96,

12929-12934.

Hrazdina, G. (1992). Compartmentation in Aromatic Metabolism. In Phenolic

Metabolism in Plants, H. A. Stafford, ed.

Hur, S., Newby, Z. E., and Bruice, T. C. (2004). Transition state stabilization by general

acid catalysis, water expulsion, and enzyme reorganization in Medicago savita

chalcone isomerase. Proc Natl Acad Sci U S A 101, 2730-2735.

Saslowsky, D. E., Dana, C. D., and Winkel-Shirley, B. (2000). An allelic series for the

chalcone synthase locus in Arabidopsis. Gene 255, 127-138.

Stafford, H. A. (1974a). The metabolism of aromatic compounds. Ann Rev Plant Physiol

25, 459-486.

Stafford, H. A. (1974b). Possible multi-enzyme complexes regulating the formation of

C6-C3 phenolic compounds and lignins in higher plants. Rec Adv Phytochem 8,

53-79.

Wilmouth, R. C., Turnbull, J. J., Welford, R. W., Clifton, I. J., Prescott, A. G., and

Schofield, C. J. (2002). Structure and mechanism of anthocyanidin synthase from

Arabidopsis thaliana. Structure (Camb) 10, 93-103.

Winkel-Shirley, B. (1999). Evidence for enzyme complexes in the phenylpropanoid and

flavonoid pathways. Physiol Plant 107, 142-149.

Page 150: Structural Characterization of the Flavonoid Enzyme Complex

135

Appendix A

Homology Modeling of Arabidopsis flavonoid enzymes.

Page 151: Structural Characterization of the Flavonoid Enzyme Complex

136

Homology models were generated for flavonone 3-hyrdoxylase, flavonoid 3’-

hydroxylase, and dihydroflavonol 4-reductase based on the crystal structures for

Arabidopsis anthocyanidin reductase (Turnbull et al., 2004; Wilmouth et al., 2002),

human cytochrome P450 2C9 (Williams et al., 2003) and E. coli UDP-galactose-4-

epimerase (Liu et al., 1997), respectively. Alignment of the amino acid sequences was

performed using Modeller6 of Sali and Blundell (1993). Five models were produced for

each protein and an average model was generated for each protein by coordinate

averaging. The averaged structures were minimized using the Sander module of Amber7

for 200 steps using steepest descent (Pearlman et al., 1995). All structure images were

manipulated with SwissPDB Viewer (Guex and Peitsch, 1997) and rendered with POV-

Ray (www.povray.org).

Page 152: Structural Characterization of the Flavonoid Enzyme Complex

137

Figure A.1: Homology models of flavanone 3-hydroxylase (a), flavonoid 3’ hydroxylase (b), and dihydroflavonol 4-reductase (c).

A. B.

C.

Page 153: Structural Characterization of the Flavonoid Enzyme Complex

138

References Cited: Guex, N., and Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an

environment for comparative protein modeling. Electrophoresis 18, 2714-2723.

Liu, Y., Thoden, J. B., Kim, J., Berger, E., Gulick, A. M., Ruzicka, F. J., Holden, H. M.,

and Frey, P. A. (1997). Mechanistic roles of tyrosine 149 and serine 124 in UDP-

galactose 4-epimerase from Escherichia coli. Biochemistry 36, 10675-10684.

Pearlman, D. A., Case, D. A., Caldwell, J. W., Ross, W. S., Cheatham, T. A., DeBolt, S.,

Ferguson, D. M., Seibel, G., and Kollman, P. (1995). AMBER, a package of of

computer programs for applying molecular mechanics, a normal mode analysis,

molecular dynamics and free energy calculations to simulate the structural and

energetic properties of molecules. Comput Phys Commun 91, 1-41.

Sali, A., and Blundell, T. L. (1993). Comparative protein modelling by satisfactoin of

spatial restraints. J Mol Biol 234, 779-815.

Turnbull, J. J., Nakajima, J. I., Welford, R. W., Yamazaki, M., Saito, K., and Schofield,

C. J. (2004). Mechanistic studies on three 2-oxoglutarate-dependent oxygenases

of flavonoid biosynthesis. J Biol Chem 279, 1206-1216.

Williams, P. A., Cosme, J., Ward, A., Angove, H. C., Matak Vinkovic, D., and Jhoti, H.

(2003). Crystal structure of human cytochrome P450 2C9 with bound warfarin.

Nature 424, 464-468.

Wilmouth, R. C., Turnbull, J. J., Welford, R. W., Clifton, I. J., Prescott, A. G., and

Schofield, C. J. (2002). Structure and mechanism of anthocyanidin synthase from

Arabidopsis thaliana. Structure (Camb) 10, 93-103.

Page 154: Structural Characterization of the Flavonoid Enzyme Complex

139

Appendix B

SPR Analysis of Flavonoid Enzymes

Page 155: Structural Characterization of the Flavonoid Enzyme Complex

140

Real-time surface plasmon resonance refractometry was performed on a Leica

AR600 automatic refractometer equipped with a 10% planar carboxymethyl PEG/PEG

mixed self assembled monolayer (CM PEG/PEG m SAM) gold chip (Reichert, Inc).

Onto the CM PEG/PEG mSAM chip a NTA moiety was attached by NHS/EDC coupling

and was then charged with NiCl2 (Lahiri et al., 1999) All experiments were carried out at

25ºC with a constant flow rate of 0.3 ml/min in PBS-T buffer. The chip was allowed to

equilibrate in PBS-T for 30 minutes before the start of the experiment. Upon the start of

the experiment, each following solution was passed over the flow cell for 10 minutes

each starting first with PBS-T, then 10µM BSA, a PBS-T wash step, next the his-tagged

protein to be immobilized is passed over the chip (anywhere from 1-3 µM), another PBS-

T wash is performed to remove any unbound protein, the unbound protein to be studied is

next passed over the chip (again, anywhere from 1-3 µM), followed by a final wash with

PBS-T.

The his-tagged protein to be immobilized onto the slide consisted of thioredoxin-

his fusions of chalcone synthase (CHS) or chalcone isomerase (CHI) that were expressed

and purified as in Chapter 3 except that the fusion partner was not cleaved. The unbound

protein consisted of CHS or CHI which was expressed and purified as in Chapter 3 with

cleavage of the fusion partner.

Page 156: Structural Characterization of the Flavonoid Enzyme Complex

141

Figure B.1. Representative SPR output. In these experiments either his::TRX::CHI (a)

or his::TRX::CHS (b) was immobilized and CHS or CHI, respectively, was passed over

the immobilized protein to detect for interactions between the two.

-5

0

5

10

15

2 0

25

3 0

35

4 0

45

0 10 20 30 40 50 60 70 80 90 100

time (minutes)

PBS-T

PBS-T

PBS-T

PBS-T

10 uM BSA his::TRX::CHI

CHS

SPR Experiment, December 10, 2001. Flow rate: 0.3 ml/min. His::TRX::CHI at 3 uM. Free CHS 1 uM.

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0 50 100 150 200 250 300 350 400 450

time - 10s

chan

ge

in r

efra

ctiv

e in

dex

PBS-T

PBS-T PBS-T

PBS-T

BSA

TRX::CHS

CHI

SPR experiment, August 21, 2001. His-tagged TRX::CHS (3 uM) immobilized on Ni-NTA gold slide with free CHI (3 uM) passed over. Flow rate: 0.3 ml/min for

10 min ea

A.

B.

Page 157: Structural Characterization of the Flavonoid Enzyme Complex

142

References Cited:

Lahiri, J., Isaacs, L., Tien, J., and Whitesides, G. M. (1999). A strategy for the generation

of surfaces presenting ligands for studies of binding based on an active ester as a

common reactive intermediate: a surface plasmon resonance study. Anal Chem

71, 777-790.

Page 158: Structural Characterization of the Flavonoid Enzyme Complex

143

Christopher David Dana http://main.biol.vt.edu/department/faculty/winkel/chris.htm

709 Earl of Chesterfield Lane 107 Fralin Biotechnology Center Virginia Beach, VA 23454 Virginia Tech (540)449-9770 Blacksburg, VA 24061 e-mail: [email protected] (540)231-9375 Education Virginia Polytechnic Institute and State University, Blacksburg, VA Ph.D. Student (expected graduation date: July, 2004) Major: Biology QCA: 3.4 James Madison University, Harrisonburg, VA Majors: Integrated Science and Technology; Biotechnology & Engineering/Manufacturing Concentrations German Minor: Music Graduation Date: May 1999 Cumulative GPA: 3.4 Work Experience and Professional Development Virginia Tech Dates: 8/99-12/03 Graduate Teaching Assistant • Teach upper level molecular biology laboratory class (Fall 2000-present) • Assisted with development of current molecular biology laboratory manual • Taught freshman level general and principles of biology laboratory class (Fall 1999-

Spring 2000)

PPD Pharmaco (Contract Research Organization) Dates: 5/98-8-98 2244 Dabney Road Richmond, VA 23230 • Received GLP training • Developed database-tracking system for audits • Developed database for all archived records on site James Madison University Dates: 8/96-5/99 • Laboratory technician • Assisted with teaching of biotechnology labs • Assisted with development of new laboratory experiments for the junior level

biotechnology class

Page 159: Structural Characterization of the Flavonoid Enzyme Complex

144

Ondal France E.U.R.L. (Subsidiary of Wella AG) Dates: 5/97-7/97 2. Rue Dennis Papin - B.P. 70305 57203 Sarrguemines Cedex, France • Laboratory technician analyzing products for conformance to specifications • Assisted with ISO 9002 certification efforts

Wella Manufacturing of Virginia, Inc. Dates: 5/96-7/96 4650 Oakley’s Lane Richmond, VA 23231 • Assisted with ISO 9002 certification efforts

Extracurricular Activities and Affiliations • Serving on the Structural Biology Search Committee • Principal 2nd Violinist, New River Valley Symphony (9/99-present) • Summer Musical Enterprises, Violinist in performance of “King and I” (5/03-8/03) • Treasurer, Biology Graduate Student Association (1/00-present) • Member: Virginia Academy of Science • Member: Sigma Xi Scientific Honor Society • Alpha Phi Omega: National Co-ed Community Service Fraternity; Spring, 1999:

Chapter parliamentarian • Golden Key National Honor Society Presentations • Structural Analysis of Flavonoid Enzyme Interactions in Arabidopsis. Presentation to

Dean DellaPenna’s Lab at Michigan State University. October 2003. • Heterologous Gene Expression in Bacteria and Yeast. Plant Metabolism Working

Group. April 2003. • Structural Analysis of Flavonoid Enzyme Interactions in Arabidopsis. 4th Annual

Arabidopsis Minisymposium. University of Maryland, College Park. April 2003 • Structural Analysis of Flavonoid Enzyme Interactions in Arabidopsis. Plant

Molecular Biology Discussion Group, April 2003. • Characterizing the Functional and Structural Domains in Flavonoid Enzymes. NIST

Neutron Division. February 2003. (Received small Honorarium for this talk) • The Flavonoid Biosynthetic Pathway: A Model for Protein-Protein Interactions in

Plants? Gordon Research Conference “Stage Invasion.” Queen’s College, Oxford, England. August 2002.

• Characterizing the Functional and Structural Domains in Flavonoid Enzymes. Plant Molecular Biology Discussion Group. May 2002.

• Homology Modeling of the tt4 Mutants in Arabidopsis thaliana. Virginia Academy of Science Meeting. May 2001.

Page 160: Structural Characterization of the Flavonoid Enzyme Complex

145

• Characterizing the Functional and Structural Domains in Arabidopsis Flavonoid Enzymes. Botany Seminar. April 2001

• Characterizing the Functional and Structural Domains in Arabidopsis Flavonoid Enzymes. Plant Molecular Biology Discussion Group. October 2000.

• Characterizing the Functional and Structural Domains in Arabidopsis Chalcone Isomerase. Virginia Academy of Science Meeting. May 2000.

• Determination of Interaction Domains in Arabidopsis Chalcone Isomerase. James Madison University, College of Integrated Science and Technology. March 2000.

Posters Presented • A.S. Duda, C.D. Dana, and B. Winkel. Substrate Specificify of Dihydroflavonol 4-

Reductase. ACS undergraduate research symposium. Poster was awarded “Best Poster Presentation.”

• C.D. Dana, S. Krueger, J.P. Noel, and B. Winkel. Structural Characterization of the Flavonoid Biosynthetic Enzyme Complex. 1st Annual Biology Department Research Symposium. September 2003.

• C.D. Dana, S. Krueger, J.P. Noel, and B. Winkel. Structural Characterization of the Flavonoid Biosynthetic Enzyme Complex. 14th International Conference on Arabidopsis Research. University of Wisconsin. June 2003.

• C.D. Dana, S. Krueger, J.P. Noel, and B. Winkel. Structural Characterization of the Flavonoid Biosynthetic Enzyme Complex. The Gordon Research Conference on Macromolecular Organization and Cell Function. Queen’s College, Oxford, England. August 2002.

• C.D. Dana, D.R. Bevan, and B. Winkel. Homology Modeling of Mutations in the Arabidopsis Chalcone Synthase Gene. Phytochemical Society of North America. Oklahoma City, OK, August 2001.

• C.D. Dana, D.R. Bevan, and B. Winkel. Homology Modeling of the tt4 Mutants in Arabidopsis thaliana. 17th Annual Research Symposium of Virginia Tech, March 2001. Received 2nd place ($150.00 award)

• C.D. Dana, D.R. Bevan, and B. Winkel. Homology Modeling of the tt4 Mutants in Arabidopsis thaliana. Virginia Tech Workshop on Bioinformatics and Computational Biology, March 2001.

• C.D. Dana, D.R. Bevan, and B. Winkel. Homology Modeling of the tt4 Mutants in Arabidopsis thaliana. The Gordon Research Conference on Macromolecular Organization and Cell Function. Queen’s College, Oxford, England. August 2000.

Publications • Dana, C., Bevan, D.R., and Winkel, B. (in preparation) Molecular Modeling of the

Chalcone Synthase Mutant Alleles. • Winkel, B., DeCourcy, K., Dana, C., Grabau, E., and Lederman, M. (2001)

Laboratory Manual: BIOL 4774 Molecular Biology Laboratory. • Saslowsky, D.E., Dana, C.D., and Winkel-Shirley, B. (2000) An allelic series for the

chalcone synthase locus in Arabidopsis. Gene. 255:127-138.

Page 161: Structural Characterization of the Flavonoid Enzyme Complex

146

Meetings Organized • Helped organize the 1st Annual Biology Department Research Symposium.

September 2003. • Gordon Research Conference “Stage Invasion.” A brief symposium before the start

of the GRC for graduate students and post docs to present their research. Queen’s College, Oxford, UK. August 2002. (This is not affiliated with the GRC).

Grants Applied for and Received • GSA Travel Fund award, $200, with matching funds from department, for travel to

the Gordon Research Conference on Macromolecular Organization and Cell Function (August 2002).

• NIST Center for Neutron Research, ~$850, to attend the NIST summer school on Neutron Scattering (June 2002).

• Joint Institute for Neutron Sciences, $1000 ($500 matching from department), to attend workshop on “Using Neutrons to Probe Structure and Dynamics in Biological Systems” at Oak Ridge National Laboratory (April 2002).

• GSA Travel Fund award, $100, Fall 2001, for travel to Phytochemical Society of North America annual meeting in Oklahoma City, OK (August 2001).

• Phytochemical Society Travel Award, $150, for travel to Phytochemical Society of North America annual meeting in Oklahoma City, OK (August 2001).

• Graduate Research Development Project, $500, with matching funds from biology department (Spring 2001).

• GSA Travel Fund award, $300, with matching funds from biology department, Fall 2000, for travel to Gordon Research Conference on Macromolecular Organization and Cell Function (August 2000).

• Graduate Research Development Project, $300, with matching funds from biology department (Spring 2000).

Other • Supervised the training and work of 4 undergraduate researchers: Ashley Duda (5/03-

current), Eric McFadden (5/01-5/03), Elizabeth Tucker (5/00 to 8/00), and Anna Watson (5/01-8/01).

• Redesigned and maintain Winkel Lab web page. Collaboratorators • Dr. Joseph Noel, Salk Institute. Crystallization of Flavonoid Enzymes. • Dr. Susan Krueger, NCNR, NIST. Solution Structure of the Flavonoid Multienzyme

Complex. • Dr. Geza Hrazdina. Cornell. Homology Modeling and Refinement of Polyketide

Synthases from Raspberry.

Page 162: Structural Characterization of the Flavonoid Enzyme Complex

147

• Dr. Craig Nessler, Virginia Tech, PPWS. Homology Modeling of Alkaloid O-Methyltransferases.