194
Effect of Plasmodiophora brassicae resting spore concentration and crop rotation on growth of clubroot-resistant crops By Jill A. Dalton A Thesis Presented to The University of Guelph In partial fulfillment of requirements for the degree of Master of Science in Plant Agriculture Guelph, Ontario © Jill A. Dalton, January 2016

Effect of Plasmodiophora brassicae ... - University of Guelph

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Effect of Plasmodiophora brassicae ... - University of Guelph

Effect of Plasmodiophora brassicae resting spore concentration and

crop rotation on growth of clubroot-resistant crops  

 

By

Jill A. Dalton

A Thesis

Presented to

The University of Guelph

In partial fulfillment of requirements

for the degree of

Master of Science

in

Plant Agriculture

Guelph, Ontario

© Jill A. Dalton, January 2016

Page 2: Effect of Plasmodiophora brassicae ... - University of Guelph

ABSTRACT

Effect of Plasmodiophora brassicae resting spore concentration and

crop rotation on growth of clubroot-resistant crops

Jill Allison Dalton Advisors:

University of Guelph Dr. Mary Ruth McDonald

Dr. Bruce D. Gossen

 

 

Clubroot caused by Plasmodiophora brassicae Woronin is an important threat to production of

canola (Brassica napus L.) and Brassica vegetables in Canada and worldwide. To help

understand the role of crop rotation in integrated clubroot management, this research examined

the pattern of decline in resting spores and the influence of spore concentration on growth and

development of clubroot-resistant crops. It was found that a portion of the resting spore

population is long-lived, but most (>99%) spores survive for only 1–2 years. Higher

concentrations of resting spores resulted in reduced plant growth and delayed development in

resistant canola and napa cabbage. However, the growth response was inconsistent across studies

and repetitions, and may be influenced by other factors such as soil type and crop species. For

canola growers, a ≥ 2-year break from canola, in combination with a clubroot-resistant cultivar, is

recommended as a clubroot management strategy wherever clubroot is found.

Page 3: Effect of Plasmodiophora brassicae ... - University of Guelph

  iii  ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my co-advisors Dr. Mary Ruth McDonald

and Dr. Bruce D. Gossen for providing me with this opportunity and for their guidance, patience,

insight and support throughout this process. Thank you for helping me to strengthen my skills in

scientific analysis, writing and public speaking. I would also like to recognize the advice and

encouragement of committee members Dr. Katerina Serlemitsos Jordan and Dr. Sean Westerveld.

Thank you all for your time and expertise.

I appreciate the financial support provided by the Canola Council of Canada, Agriculture

and Agri-Food Canada through the Canola Science Cluster of Growing Forward 2, and the

University of Guelph Department of Plant Agriculture.

Thank you to Dr. Abhi Deora, Dr. Fadi Al-Daoud, Travis Cranmer and many other lab

colleagues for technical advice, support and encouragement. I am very grateful to Kevin Vander

Kooi, Laura Riches, Shawn Janse, Misko Mitrovic, Dennis Van Dyk and summer staff at the

Muck Crops Research Station for all of your help with the 2014 and 2015 field trials. I would

also like to acknowledge Godfrey Chu, Ken Bassendowski, Denis Pageau and others at AAFC

and University of Guelph for providing materials and support for the crop rotation studies.

I am very grateful to my parents for encouraging me to continue learning, to always

do my best and to believe in myself. I would like to say a special thank you to my sister for

teaching me to think critically. I appreciate the tremendous amount of support I have received

from my friends and family, especially Jack & Wilda Mardlin, Elizabeth Fraser, Gabe Sawhney,

Ashleigh Dalton, Laura Dalton, Martin Dalton and, finally, John Jeffrey Mardlin, for truly

going above and beyond a reasonable amount of support.

 

Page 4: Effect of Plasmodiophora brassicae ... - University of Guelph

  iv  TABLE  OF  CONTENTS  

 

CHAPTER  ONE  –  LITERATURE  REVIEW  .........................................................................................................  1  1.1   Brassica  crops  and  their  relatives  ......................................................................................................  1  1.1.1   Canola  and  other  oilseed  Brassica  crops  ................................................................................  1  1.1.2   Brassica  vegetables  ..........................................................................................................................  2  1.1.3  History  .......................................................................................................................................................  4  1.1.3   Economic  importance  of  Brassica  crops  in  Canada  ...........................................................  4  1.1.4   Pests  and  diseases  in  Brassica  crops  ........................................................................................  5  

1.2   Clubroot  of  Brassica  .................................................................................................................................  7  1.2.1   Host  range  and  history  of  clubroot  ...........................................................................................  7  1.2.2   Causal  agent  and  taxonomy  ..........................................................................................................  8  1.2.3   Disease  cycle  .......................................................................................................................................  9  1.2.4   Pathotypes  ........................................................................................................................................  12  1.2.5   Incompatible  Interactions  ..........................................................................................................  13  

1.3   Factors  influencing  infection  and  development  ........................................................................  14  1.3.1   Soil  moisture  ....................................................................................................................................  15  1.3.2   Temperature  ....................................................................................................................................  15  1.3.3   Soil  pH  .................................................................................................................................................  16  1.3.4   Resting  spore  concentration  .....................................................................................................  17  

1.4   Disease  management  ............................................................................................................................  17  1.4.1   Cultural  control  ...............................................................................................................................  17  1.4.2   Synthetic  fungicides  and  surfactants  .....................................................................................  19  1.4.3   Biological  control  and  biofungicides  .....................................................................................  21  1.4.4   Host  plant  resistance  ....................................................................................................................  23  1.4.5   Metabolic  cost  of  resistance  ......................................................................................................  25  

1.5   Summary  and  objectives  .....................................................................................................................  26  CHAPTER  TWO  -­‐  EFFECT  OF  RESTING  SPORE  CONCENTRATION  ON  GROWTH  OF    CLUBROOT-­‐RESISTANT  BRASSICA  CROPS  .................................................................................................  28  2.1   Introduction  ..............................................................................................................................................  28  2.2   Materials  and  Methods  .........................................................................................................................  33  2.2.1   Controlled  environment  study  –  canola  ...............................................................................  33  2.2.2   Large  pot  studies  –  outdoors  ....................................................................................................  35  2.2.3   Controlled  environment  study  of  canola,  cabbage  and  napa  cabbage  ....................  36  2.2.4   Field  trials  .........................................................................................................................................  37  2.2.5   Controlled  environment  pH  study  ..........................................................................................  39  2.2.6   Statistical  analysis  .........................................................................................................................  40  

2.3   Results  .........................................................................................................................................................  41  2.3.1   Controlled  environment  study  .................................................................................................  41  2.3.2   Large  pot  studies  –  outdoors  ....................................................................................................  44  2.3.3   Controlled  environment  study  –  canola,  napa  cabbage  and  cabbage  .....................  45  2.3.4   Field  trials  .........................................................................................................................................  46  2.3.5   pH  study  .............................................................................................................................................  52  

2.4   Discussion  ..................................................................................................................................................  54  

Page 5: Effect of Plasmodiophora brassicae ... - University of Guelph

  v  CHAPTER  THREE  -­‐  DECLINE  IN  RESTING  SPORES  AND  EFFECT  OF  CROP  ROTATION  FOLLOWING  A  SUSCEPTIBLE  CROP  ...............................................................................................................  62  3.1   Introduction  ..............................................................................................................................................  62  3.2   Materials  and  Methods  .........................................................................................................................  67  3.2.1   Decline  in  resting  spores  over  time  following  susceptible  canola  ............................  67  3.2.2   Effect  of  resistant  canola  on  the  concentration  of  resting  spores  in  soil  ...............  70  3.2.3   Controlled  environment  study  –  crop  rotation  and  spore  concentration  .............  71  3.2.4   Statistical  analysis  .........................................................................................................................  73  

3.3   Results  .........................................................................................................................................................  74  3.3.1   Decline  in  resting  spores  over  time  following  susceptible  canola  ............................  74  3.3.2   Effect  of  resistant  canola  on  the  concentration  of  resting  spores  in  soil  ...............  75  3.3.3   Weather  ..............................................................................................................................................  77  3.3.4   Controlled  environment  –  crop  rotation  and  spore  concentration  ..........................  78  

3.4   Discussion  ..................................................................................................................................................  81  CHAPTER  FOUR  -­‐  GENERAL  DISCUSSION  ...................................................................................................  91  LITERATURE  CITED  ..............................................................................................................................................  97  APPENDIX  1:  SUPPLEMENTARY  TABLES  FOR  CHAPTER  TWO  ......................................................  116  APPENDIX  2:  SUPPLEMENTARY  TABLES  FOR  CHAPTER  THREE  .................................................  172  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 6: Effect of Plasmodiophora brassicae ... - University of Guelph

  vi  LIST  OF  TABLES  

Table 2.1 Soil properties of field trial sites at the Muck Crops Research Station at Holland Marsh, Ontario in 2014 and 2015.  ...........................................................................................................................  38  

Table 2.2 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible canola inoculated with 1 x 106 spores ml-1 under controlled conditions.  ...................................................  42  

Table 2.3 Clubroot incidence (CI) and severity (disease severity index, DSI) in canola breeding line ACS-N39 (susceptible check) and resistant cultivar 45H29 in an outdoor trial using large pots near Bradford, ON, 2014 and 2015.  .....................................................................................  44  

Table  2.4 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible canola, napa cabbage and cabbage inoculated with 1 x 106 spores ml-1 P. brassicae, under controlled conditions.  ....................................................................................................................................  45  

Table 2.5 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible canola and napa cabbage grown in field soil at Muck Crops Research Station, 2014.  .........................  46  

Table 2.6 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible and resistant canola cultivars, grown in field soil at sites with high, lower and no measurable concentration of resting spores at the Muck Crops Research Station, 2015.  ..............................  47  

Table  3.1 Quantification of resting spore concentration using qPCR after crop rotation treatments including fallow (F), resistant (R) and susceptible canola (S), sampled in spring and fall of 2014 and spring of 2015 at Normandin, Quebec.  .............................................................................  777  

Table  3.2 Comparison of soil locations used in crop rotation and inoculum trials, under controlled environment in Guelph, Ontario, 2014.    .............................................................................  71

Table 3.3 Quantification of resting spore concentration using qPCR after crop rotation treatments including fallow (F), resistant (R) and susceptible canola (S), sampled in spring and fall of 2014 and spring of 2015 at Normandin, Quebec..  ...............................................................................  76  

Table 3.4 Clubroot incidence (CI) and severity (disease severity index, DSI) in canola breeding line ACS-N39 (susceptible check) and the resistant cultivar 45H29, inoculated with P. brassicae and grown under controlled conditions  ..........................................................................  78

Table 3.5 Effect of crop rotation and spore concentration on biomass (dry shoot weight, g) of clubroot-resistant canola, under controlled conditions in field soil from Elora, ON, and Scott, SK, 2014.  ...............................................................................................................................................  81  

 

 

 

 

 

 

 

 

 

 

Page 7: Effect of Plasmodiophora brassicae ... - University of Guelph

  vii  LIST  OF  FIGURES  

 

Figure 2.1 Plant height of clubroot-resistant canola at 6 weeks after inoculation with increasing concentrations of Plasmodiophora brassicae resting spores under controlled conditions (two repetitions, n=8, P = 0.005).  ........................................................................................................................  43  

Figure 2.2 Plant maturity at harvest in canola grown in controlled conditions in response to increasing concentration of Plasmodiophora brassicae resting spores (P = 0.007) in the first repetition.  ...........................................................................................................................................................  43  

Figure  2.3 Plant height of three resistant canola at the Muck Crops Research Station, Ontario, at sites with high, lower and undetectable resting spore concentrations, 2014 and 2015.  ..........  49  

Figure  2.4 Plant height at 8 weeks after seeding (WAS), biomass at harvest and maturity at 8 WAS for resistant canola cultivars grown in field soil with 7 x 105 spores g-1 soil (Low) and 7 x 106 spores g-1 soil (High) at Muck Crops Research Station, 2014. Bars with the same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison test.  ................  50  

Figure 2.5 Plant height and maturity at 8 weeks after seeding (WAS), biomass at harvest for resistant canola cultivars grown in field soil with <1000 spores g-1 soil (BDL), 3 x 106 spores g-1 soil (Low) and 1 x 107 spores g-1 soil (High) at Muck Crops Research Station, 2015. Bars with the same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison test.  .......................................................................................................................................................................  50  

Figure 2.6 Leaf length and biomass of napa cabbage at sites with 7 x 105 spores g-1 soil (Low) and 7 x 106 spores g-1 soil (High) in field trials at Muck Crops Research Station, Ontario, 2014. Bars with the same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison test.  ...............................................................................................................................................  51  

Figure 2.7 Leaf length and biomass of napa cabbage at sites with <1000 spores g-1 soil (BDL), 3 x 106 spores g-1 soil (Low) and 1 x 107 spores g-1 soil (High) in field trials at the Muck Crops Research Station, Holland Marsh, Ontario, 2015. Bars with the same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison test.  ..........................................................  52  

Figure 2.8 Clubroot incidence (CI) in susceptible canola grown under controlled conditions at a range of pH (5.5 to 7.5) and concentrations of resting spores of P. brassicae at 22 days after inoculation (DAI).  ..........................................................................................................................................  53  

Figure 3.1 Decline in resting spore concentrations of Plasmodiophora brassicae over time following susceptible canola at Normandin, Quebec, assessed in 2014.  .....................................  75  

Figure 3.2 Monthly rainfall (bars) and mean monthly air temperature (line) at Normandin, Quebec, for May to September, 2007-2013.  ..........................................................................................  78  

Figure 3.3 Effect of crop rotation and inoculation with Plasmodiophora brassicae on plant height (area under growth stairs, AUGS) of clubroot-resistant canola, under controlled conditions in field soil from Elora, ON, and Scott, SK in 2014.  ..........................................................................  80  

 

 

Page 8: Effect of Plasmodiophora brassicae ... - University of Guelph

  1  CHAPTER ONE – LITERATURE REVIEW

 

1.1 Brassica crops and their relatives 1.1.1 Canola and other oilseed Brassica crops

The family Brassicaceae (formerly Cruciferae) is composed of angiosperms in the order

Brassicales, distinguishable by their corolla, made up of four flower petals shaped like a cross

(Cartea et al., 2011). Plants in the Brassicaceae have six stamens, and the four inner stamens are

shorter than the two outer stamens (Franzke et al., 2011). The family contains 338 genera and

3709 species, including many important crops used for a variety of purposes including vegetables

for human consumption, oils for cooking and industrial use, condiments and livestock fodder (Al-

Shehbaz et al., 2006). There are three important diploid Brassica species, B. oleracea L., B. rapa

L. and B. nigra L. (Dixon, 2007). Three amphidiploid species, B. napus L., B. carinata L. and

B. juncea L., have emerged from crosses of diploid species (U, 1935). Arabidopsis thaliana (L.)

Heynh, which is a member of the family Brassicaeae, is an important model organism used in the

study of angiosperms (Arabidopsis Genome Initiative, 2000; Al-Shehbaz et al., 2006).  

Canola is a quality standard for oilseed rape developed from B. napus (rapeseed) in the

1970s by Canadian plant breeders from Agriculture and Agri-Food Canada and the University of

Manitoba. By definition, canola produces seed with oil that contains less than 2% erucic acid and

less than 30 μmol g-1 of glucosinolates (Canola Council of Canada, 2011). Other names for

oilseeds with a comparable nutritional profile include Double Zero (00) Rapeseed and LEAR

(Low Erucic Acid Rapeseed) oil (Health Canada, 2003). The majority of canola cultivars grown

in Canada are B. napus, but canola-quality cultivars of B. rapa and B. juncea have also been

released (Canola Council of Canada, 2011).  

Page 9: Effect of Plasmodiophora brassicae ... - University of Guelph

  2  Canola may be fall-seeded or spring-seeded, but almost all of the canola grown in western

Canada is spring-seeded. Like all Brassica crops, canola plants emerge quickly, with a pair of

cotyledons developing 4-7 days after seeding (Lamb, 1989). After the first pair of true leaves

develop, a rosette of broader leaves is formed. At this point, the fall-seeded winter annual enter a

dormancy state, whereas the spring-seeded plants continue growth. The next stage of growth is

stem elongation, followed by flower bud production (Hayward, 2012). Flowers are produced in

racemes, with one raceme developing at the very top of the plant, followed by axillary racemes.

The corolla consists of four yellow petals, which open beginning at the base of each raceme.

Petals drop after fertilization, and a cylinder-shaped pod is formed, called a silique. Upon

maturation, this pod contains between 15 and 40 seeds. Leaf senescence begins as the pods

mature, starting with the leaves at the base of the plant (Lamb, 1989).

Brassica napus and B. rapa make up a large majority of the global Brassica production

(Hayward, 2012). B. rapa, one of the original crossing parents of B. napus, contains the most

cold-tolerant cultivars of Brassica oilseeds. These are popular in areas where early maturity is

important due to the colder climate, including parts of Canada, Sweden and Finland. In India,

B. juncea makes up 80% of the oilseed production because of its heat and drought tolerance

(Hayward, 2012). In East Africa, B. carinata is cultivated and has recently been developed into a

biofuel (Abukutsa and Onyango, 2005; Hayward, 2012).

1.1.2 Brassica vegetables

Brassica vegetables are also known as cruciferous vegetables, cole crops, cabbage family

vegetables or mustard family vegetables. They are morphologically diverse, and are cultivated for

various edible parts of the plant, including roots, stems, leaves, seeds, flowers and sprouts

Page 10: Effect of Plasmodiophora brassicae ... - University of Guelph

  3  (Dixon, 2007). They are cultivated in many countries around the world, most commonly in

temperate climates of the Northern Hemisphere. Europe is a major center of diversity for

B. oleracea, or European Brassica vegetables, while Asia is the center of diversity for B. rapa, or

Asian Brassica vegetables (Cartea et al. 2011). Brassica oleraceae includes cabbage, broccoli,

cauliflower, kale, kohlrabi and Brussels sprouts. Brassica rapa includes turnip and Asian

vegetables that are leafy (such as pak choy) and headed (napa cabbage, also called Chinese

cabbage). Rutabaga is the only common B. napus vegetable. Brassica juncea crops include

mustard seed for condiment mustard and Asian greens such as red leaf mustard and mizuna.

Culinary vegetables in this family outside of the Brassica genus include Eruca sativa (arugula)

and Raphanus sativus (radish) (Tsunoda, 1980; Appel and Al-Shehbaz, 2003). Plants become

woody and less flavourful during the reproductive stage, and therefore non-bolting varieties are

popular for vegetable production. Most are biennial in nature, but are grown as annuals when

cultivated.

There has been increased attention to the health benefits of eating Brassica vegetables.

Kale and broccoli have been reported to reduce the risk of age-related chronic and degenerative

diseases and several types of cancer (Traka et al., 2010; Soengas et al., 2011). Health-promoting

phytochemicals in Brassica greens include vitamin C, carotenoids and DL-a-tocopherol (Singh et

al., 2007). Glucosinolates, such as glucoraphanin, glucoiberin and glucoraphanin, have received

special attention for their antioxidant and anticarcinogenic properties (Cartea et al., 2011).

Brassica greens are also a source of dietary fiber, which interacts with phytochemicals in the

digestive system (Palafox-Carlos et al., 2011).

 

Page 11: Effect of Plasmodiophora brassicae ... - University of Guelph

  4  1.1.3 History

Wild species of Brassica can be found across Europe and Central Asia, and the origin of

Brassica species is believed to be in the Mediterranean region (Tsunoda, 1980). Brassica rapa is

native to the highlands, rather than the coastal areas of the Mediterranean, which led it to adapt

relatively quickly to Scandinavia, Eastern Europe and China (Tsunoda, 1980). However, Ignatov

et al. (2010) has suggested that China is a centre of origin for the B. rapa Asian subgroup. Wild

types of B. oleracea with distinctive phenotypes are spread throughout Europe and the

Mediterranean, in small, isolated areas (Rakow, 2004).

 

1.1.3 Economic importance of Brassica crops in Canada

Oilseed rape is the second most economically important oilseed crop worldwide

(Snowdon et al., 2007). It is second only to soybean, and has been since the 1980s. As of 2008,

Canada was the largest producer of oilseed rape in the world (FAOSTAT data 2008, cited in

Hayward, 2012). Canola has the highest farm cash receipts of any agricultural crop in Canada

(Statistics Canada, 2013), and total production and farm cash receipts have increased from 12.9

million tonnes ($5.1 billion) in 2009 to 14.7 million tonnes ($8.2 billion) in 2013. In 2011, the

total acreage of canola production increased. However, there was an 8% decrease in canola

production due to a 17.5% decrease in yield per hectare (Statistic Canada, 2013). Production of

Brassica vegetable crops is listed in Table 1.1.

Page 12: Effect of Plasmodiophora brassicae ... - University of Guelph

  5  Table 1.1. Production (metric tonnes) and value of Brassica vegetables in Canada in 2012

(Statistics Canada, 2013).  

  Farm cash receipts (millions $ Canadian)  

Production (thousand metric tonnes)

Cabbage   62.5   165  

Broccoli   40.2   35  

Cauliflower   23.9   31  

Rutabaga and turnip   21.5   50  

Radish   10.8   12  

Brussels sprouts   7.5   6  

1.1.4 Pests and diseases in Brassica crops

There are many diseases that affect Brassica crops in Canada, caused by bacteria, fungi,

viruses and other pathogens. Sclerotinia sclerotiorum (Lib.) Massee, a fungal stem rot, affects

both oilseed and vegetable crops worldwide. It is one of the most devastating diseases of

B. napus (rapeseed) crops in China (Zhao and Meng, 2003). Blackleg, also known as phoma stem

canker or dry rot, is caused by Leptosphaeria maculans (Desmaz.) Ces & De Not. This disease

affects oilseed crops as well as vegetable crops (Fitt et al., 2006). Alternaria leaf spot, caused by

Alternaria brassicae (Berk.) Sacc. and A. brassicicola (Schwein.) Wiltshire, attack oilseed,

vegetable and condiment brassica crops worldwide (Humpherson-Jones and Phelps, 1989).

Pathogens that cause damping off in Brassica crops include Fusarium spp., Rhizoctonia solani

(Jensen et al., 1999), and Pythium ultimum Trow and P. debaryanum Auct. non R. Hesse

(Abdelazher, 2003; Tanina et al., 2004). Downy mildew, caused by Peronospora parasitica, is an

oomycete pathogen of Brassica crops (Jensen et al., 1999). Clubroot, caused by Plasmodiophora

brassicae Woronin,  is an economically important disease all over the world (Garber and Aist,

Page 13: Effect of Plasmodiophora brassicae ... - University of Guelph

  6  1979; Howard et al., 2010; Karling, 1968; Voorrips, 2003), but has the largest impact in

temperate countries (Karling, 1968).

Bacterial diseases are more problematic for vegetable crops than for oilseed crops. The

bacterial disease of greatest economic importance is black rot, caused by Xanthomonas

campestris pv. campestris (Pammel) Dowson (Williams, 1980; Berg et al., 2005). Another

important bacterial disease in vegetables is bacterial soft rot, caused by Erwinia carotovora var.

carotovora (Jones) Bergey et al. and Pseudomonas marginalis pv. marginalis (Brown) Stevens

(Ren et al., 2000; Charron and Sams, 2002).

Cauliflower mosaic virus and cabbage black ring spot are two viruses of Brassica

vegetables that are transmitted by the aphid species Myzus persicae Sulz. and Brevicoryne

brassicae L. (Broadbent, 1957). Another virus that affects B. rapa (turnip) crops is turnip yellow

mosaic virus, transmitted by flea beetles in the genus Phyllotreta (Broadbent, 1957).

Brassica crops are subject to herbivory by insects. Many species in the genus Lepidoptera

are herbivores of Brassica crops in Canada. Major pests of Brassica vegetable crops in Ontario

include caterpillar pests, such as imported cabbageworm (Pieris rapae L.), cabbage looper

(Trichoplusia ni Hübner), diamond-back moth (Plutella xylostella L.) and cabbage maggot (Delia

radicum L.). Mammestra configurata, also known as bertha armyworm, is a major oilseed crop

defoliator. Flea beetles (order Coleoptera, genus Phyllotreta) are important leaf feeders,

especially on seedlings (Lamb, 1989). Lygus spp. (L. elisus, L. lineolaris) and aphids

(M. persicae and B. brassicae) cause damage but generally have a minor impact on yield (Lamb,

1989). Swede midge (Contarinia nasturtii Keiffer) is an important pest of Brassica crops in

Ontario and an emerging pest in parts of western Canada. Native to Europe and Asia, adults

emerge from soil cocoons in mid May until mid June, with peak emergence occurring in the first

Page 14: Effect of Plasmodiophora brassicae ... - University of Guelph

  7  week of June (Allen et al., 2008). A second generation emerges 7-21 days later, depending on

climatic conditions, and there are generally four to five overlapping generations throughout the

season. Control of Swede midge relies on careful timing of insecticide application, based on

pheromone traps and field scouting (Hallett et al., 2009).

1.2 Clubroot of Brassica 1.2.1 Host range and history of clubroot

Clubroot is caused by the obligate Rhizarian pathogen Plasmodiophora brassicae

Woronin. Severe clubbing symptoms reduce the uptake of nutrients and water in Brassica crops

(Burki et al., 2010; Cavalier-Smith et al., 2013). Clubroot can infect most or all cultivated and

non-cultivated species of the family Brassicaceae, including all of the crops previously listed

(Sections 1.1.1, 1.1.2), and cruciferous weeds, such as Capsella bursa-pastoris (L.) Medik.

(shepherd's-purse), Sinapsis arvensis (L.) DC. (wild mustard), Diplotaxis muralis L. (annual

wall-rocket), Sisymbrium officianale (L.) Scop. (hedge mustard) and Arabidopsis thaliana (L.)

Heynh. (Buczacki and Ockendon, 1979). Clubroot was introduced to North America from

Europe, as a result of cross-Atlantic transport of livestock and clubroot-infected animal fodder. In

the 1920s, clubroot was reported in British Columbia, Ontario, Quebec and the Maritimes

(Drayton et al., 1926). Cabbage, cauliflower and rutabaga were the three main crops affected at

that time. Clubroot was first confirmed in canola fields in Alberta in 2003 (Tewari et al., 2005),

in Saskatchewan in 2009 (Dokken-Bouchard et al., 2010) and in Manitoba in 2013 (Strelkov and

Hwang, 2014). Previously, there were unpublished reports of clubroot in vegetable gardens in

western Canada (Tewari et al., 2005). In the United States, clubroot was confirmed on canola in

North Dakota in 2013 (Chittem et al., 2014).

Page 15: Effect of Plasmodiophora brassicae ... - University of Guelph

  8  The characteristic symptoms of clubroot are abnormal swelling of the roots caused by

hyperplasia and hypertrophy (Karling, 1968). Nutrient and water transport are progressively

inhibited as the clubs develop (Voorrips et al., 2003). This leads to the first noticeable,

aboveground symptoms of affected plants, which are wilting, yellowing and stunting. In severe

infections, plants die prematurely (Karling, 1968). Clubroot reduces both the yield and the quality

of the crop (Wallenhammar et al., 1999; Strelkov et al., 2005; Tewari et al., 2005; Dixon, 2009).

1.2.2 Causal agent and taxonomy

The causal agent of clubroot is P. brassicae, described by Mikhail Woronin in 1878.

P. brassicae is an obligate, biotrophic, soil-borne pathogen that reproduces in the cytoplasm of

the host plant cell (Karling, 1968; Williams and McNabola, 1967). The pathogen can survive

without a host as a dormant resting spore in soil for many years (Wallenhammar, 1996).

P. brassicae is classified in the order Plasmodiophorales, which consists of plant pathogens with

a multinucleate vegetative form, called a plasmodium (Karling, 1968). Characteristics of this

order include life stages of both resting spores and motile zoospores with two whiplash flagella

of unequal length. The family Plasmodiophoraceae consists of nine genera and 35 species. The

genera are distinguishable by the organization of cysts and sporangia. All of the species in the

genus Plasmodiophora have a plasmodial stage within the cell of the host plant, and they all

cause malformation due to hypertrophy (Karling, 1968). As scientific knowledge of clubroot has

developed over the past 100 years, the taxonomic classification of P. brassicae has evolved.

Plasmodiophorales, now classified in the kingdom Chromista, were previously classified as fungi

and protozoans (Cavalier-Smith, 2013). One example of a structural difference that distinguishes

Page 16: Effect of Plasmodiophora brassicae ... - University of Guelph

  9  chromists from fungi is the production of flagellated zoospores. Chromists have tubular ciliary

hairs, which protozoans lack (Cavalier-Smith, 1993).

1.2.3 Disease cycle

The lifecycle of P. brassicae begins with the germination of resting spores, followed by

two infection phases (Tommerup and Ingram, 1971). Resting spores are round and 2.8-5.0 µm in

diameter (Ayers, 1944; Ingram and Tommerup, 1972). Filamentous materials connect young

resting spores, while mature resting spores develop spines (Ikegami et al., 1978). During resting

spore formation, the nucleus and nucleoli become much smaller and undergo meiosis I and II,

characterized by two nuclear divisions (Garber and Aist, 1979). Cell walls of resting spores

contain chitins, proteins and lipids, and are resistant to microbial degradation (Buczacki and

Moxham, 1983). Resting spores are uni-nucleate and contain no nucleoli, but do contain

complete organelles (Williams and McNabola, 1967). The decay of clubs is aided by soil micro-

organisms (Ingram and Tommerup, 1972). The decay releases resting spores into the soil. Resting

spores can survive in a dormant state for a long period of time. The half-life of a soil’s infective

capacity has been calculated at 3.6 years in Sweden (Wallenhammar, 1996) and 4.4 years in

western Canada (Hwang et al., 2013). Under favourable conditions, resting spores germinate to

produce primary zoospores. Germination can take place with or without a host plant present

(Ingram and Tommerup, 1972; Friberg, 2005), but the frequency of germination is increased in

the presence of a compatible host (Macfarlane, 1970). Environmental influences on germination

include pH, soil moisture, soil temperature, and calcium ion concentration (Kageyama and

Asano, 2009). The likelihood of germination also increases with age and the level of decay of

spore-containing clubs (Macfarlane, 1970).  

Page 17: Effect of Plasmodiophora brassicae ... - University of Guelph

  10  Root hair infection (RHI) occurs when primary zoospores move through the soil via films

of water and then penetrate root hairs of plants (Kageyama and Asano, 2009). Ayers (1944)

reported that primary zoospores are 2.8–5.9 µm in length. Primary zoospores are spindle-shaped

or ovate, with two flagella of unequal length (Ayers, 1944). The flagella are angled at 45 degrees

from each other, and consist of two micro-tubules and nine outer doublets (Aist and Williams,

1971). Zoospores attach to the cell wall of host root hairs in preparation for encystment. Flagella

coil around the zoospore body so that the zoospore is slightly flattened against the host cell wall,

and the axonemes are retracted so that the sheath membrane is no longer surrounding them (Aist

and Williams, 1971). During encystment, an enlarged adhesorium is formed by swelling of a

cylindrical structure called a satchel, which is 700 nm in length and packed with Golgi vesicles

and condensed cytoplasm (Aist and Williams, 1971). The adhesorium punctures the cell wall of

the root hair following adhesion to the cell wall (Aist and Williams, 1971). Host wall degradation

does not appear to be a factor in root hair penetration (Aist and Williams, 1971). RHI can occur

within 24 hours of resting spore germination (Dobson and Gabrielson, 1983). Penetration of the

cell wall takes 2.5-3.5 h. After a zoospore has entered the host cell, it becomes amoeboid and

forms a multinucleate thallus (Ayers, 1944). As thalli develop, they become detached from the

point of penetration in the root hair cell wall. Primary plasmodia undergo synchronous mitotic

divisions to form zoosporangia, each containing 4-16 secondary zoospores (Ayers, 1944; Ingram

and Tommerup, 1972). At favourable temperature conditions of 20-25°C, zoosporangial numbers

increase significantly between 3 and 4 days (Dobson and Gabrielson, 1983). At the optimum

temperature of 25°C, secondary zoospores are fully developed after approximately 5 days, and

are then released into the soil (Sharma et al., 2011a). There are no clubbing symptoms or crop

losses due to colonization of root hairs (Macfarlane, 1952; Kageyama and Asano, 2009).

Page 18: Effect of Plasmodiophora brassicae ... - University of Guelph

  11  Cortical infection, or secondary infection, follows RHI (Donald et al., 2008). Secondary

zoospores enter and infect root cortical cells (Kageyama and Asano, 2009). Aist and Williams

(1971) suggest that a similar process to the penetration of root hairs may occur in cortical

infection. The size of secondary zoospores after release is 9.6-14.4 µm in diameter (McDonald et

al. 2014). The mean diameter of encysted zoospores is 21.7 µm. Secondary zoospores are much

smaller when located inside root hairs (2.7-5.0 µm in very early stages, 3.3-4.5 µm in early

stages, and on average 8 µm later on) (Ikegami et al., 1978). The size difference between

secondary zoospores located in root hairs vs. released may be caused by higher osmotic or

physical pressure on secondary zoospores located inside root hairs (McDonald et al., 2014). In

the very early stage of development, secondary zoospores are spherical, with a smooth surface

and a membranous envelope (Ikegami et al., 1978), while in later stages, the surface of the

zoospore becomes rougher and the shape may range from spherical to ovate. Ikegami et al.

(1978) proposed that morphological differences between secondary zoospores may be related to

the nutrients available within the host cell tissue. Secondary zoospores are initially uni-nucleate

(McDonald et al., 2014), which indicates that fusion of zoospores may not be required for

secondary infection to take place, as was suggested in previous studies (Ingram and Tommerup,

1972). Under favourable   conditions of 20-25°C temperature and high soil moisture, cortical

infection by secondary zoospores can occur in 3 hours (Dobson and Gabrielson, 1983).

Clubs are formed on plant roots, as the result of redirection of carbohydrates from the rest

of the plant to infected roots caused by the pathogen (Evans and Scholes, 1995). Cellular

multiplication, elongation, hypertrophy and hyperplasia take place at this stage, leading to visible

clubbing symptoms (Karling, 1968; Deora et al., 2012). Morphological changes are linked to the

disruption of plant regulators such as cytokinins and auxins, and an increase in host metabolism

Page 19: Effect of Plasmodiophora brassicae ... - University of Guelph

  12  (Muller and Hilgenberg, 1986; Ludwig-Muller, 1993; Jameson, 2000). Production of amino

acids, proteins, starches and lipids all increase during club development. Increases in

glucosinolate content, myrosinase and nitrilase also occur during club formation (Ludwig-Muller

et al., 1999; Grsic-Rausch et al., 2000).

Plasmodia form an envelope with seven layers and uniform thickness, mostly made up of

cells from the host, which surrounds each plasmodium during vegetative growth (Williams and

McNabola, 1967). Vegetative plasmodia consist of many spherical bodies with smooth surfaces,

which together create a sponge-like structure (Ikegami et al., 1978). Growth stops when

sporogenesis begins (Williams and McNabola, 1967). At this time, the envelope that surrounds

each plasmodium gets thinner, shedding its outer layers (Williams and McNabola, 1967). By the

time the vegetative growth period is ending, the plasmodia take up most of the host cell, and the

cytoplasm has been forced to the periphery of the cell (Williams and McNabola, 1967).

Eventually the clubs begin to decrease their production of sugars, RNA and starch, and the

plasmodia divide into millions of resting spores. The transition to resting spores happens very

quickly (Williams and McNabola, 1967). As the roots break down and decay, the plant loses

what is left of its ability to take in water and nutrients and can no longer support itself (Williams

and McNabola, 1967; Voorrips et al., 2003).  

 

1.2.4 Pathotypes

Pathogen specialization in P. brassicae, historically known as ‘races,’ are now called

‘pathotypes’ (Williams, 1966; Ayers, 1972). Pathotype is based on virulence, whereas race

categories are based on genes, alleles or chromosome structure (Sturhan, 1985). In Canada, the

Williams pathotype classification system is typically utilized (Williams, 1966). In this

Page 20: Effect of Plasmodiophora brassicae ... - University of Guelph

  13  classification system, some populations may have only one phenotype present for virulence,

while others have several phenotypes in different frequencies (Sturhan, 1985).

Pathotypes 2, 3, 5, and 6 are most prevalent in Canada. Pathotype 6 is most common in

Ontario, and also occurs in British Columbia and Quebec (Reyes, 1974; Williams, 1966).

Pathotype 2 is most prevalent in Quebec, but is also found in the Atlantic provinces (Ayers,

1972). Pathotypes 3 and 5 have been found on canola in Alberta since 2003, and presumably the

other prairie provinces (Ayers, 1972; Cao et al., 2007). In addition, studies of single-spore

isolates indicate that pathotype 8 may be present near Edmonton, Alberta and Orton, Ontario

(Xue et al., 2008). In 2013, a novel strain of the pathogen, initially called 5x, was identified near

Edmonton, Alberta. A range of genotypes of the pathogen have since been identified in fields

where resistance has been overcome. The various genotypes that were subsequently identified

were virulent against all commercially available canola cultivars in Canada (Cao et al., 2015;

S. Strelkov, unpublished). The new pathotypes are referred to collectively as pathotype X until a

more definitive identification is possible.

1.2.5 Incompatible Interactions

An incompatible interaction takes place between a host plant that is resistant and a

virulent strain of the pathogen. In most incompatible interactions with P. brassicae, clubroot

symptoms do not develop and resting spores are not produced (Ludwig-Muller et al., 1999; Feng

et al., 2012a). Root hair infection can occur on host and non-host plants (Dixon, 2006).

Secondary plasmodia have been observed in clubroot-resistant canola cultivars (Donald et al.,

2008; Gludovacz, 2013; McDonald et al., 2014). However, pathogen development is inhibited

before completing its lifecycle and clubroot symptoms are not observed (Morgner et al., 1995;

Page 21: Effect of Plasmodiophora brassicae ... - University of Guelph

  14  Hwang et al., 2011b). During this stage, auxin levels increase in resistant plants (Ludwig-Muller

et al., 1993). Park et al. (2007) suggested that an antagonistic interaction between growth

hormones such as auxin and salicylic acid may lead to reductions in plant growth during disease

resistance.

 

1.3 Factors influencing infection and development  

The small but abundant resting spores of P. brassicae are readily transported anywhere that

contaminated soil or water can go. For example, spores can be moved on Brassica seedlings for

transplanting, hand tools, trucks, farm machinery, livestock, compost, harvest bins, boots, gloves

and other items of clothing of farm workers. Transportation of livestock fodder, such as turnip,

may also transport the disease (Donald et al., 2006). Clubroot spread from one localized area to

an entire region in the Holland Marsh, Ontario, following flooding associated with Hurricane

Hazel in 1954 (Conners et al., 1956). This indicates that flooding may be responsible for

movement of resting spores up to 8 km. Resting spores may also be present in airborne dust and

may be dispersed up to 2 km by wind-mediated soil erosion. However, it is not known whether

those spores remain viable and it is unlikely that enough windborne resting spores travel

distances that are long enough to establish new infestations (Rennie et al., 2015).

The physical condition of the soil has a great influence on clubroot incidence and severity.

Important influences on clubroot development include soil moisture, temperature, pH,

concentration of resting spores, and the level of host plant resistance or susceptibility (Murakami

et al., 2002; Dixon, 2006). Micronutrient levels such as calcium and boron can also have an

effect on clubroot development (Webster and Dixon 1991a, 1991b).

Page 22: Effect of Plasmodiophora brassicae ... - University of Guelph

  15  1.3.1 Soil moisture

High soil moisture is favourable for clubroot development (Karling, 1968). Soil moisture

is generally measured as gravimetric water content (mass of water per mass of soil) or volumetric

water content (volume of water per volume of soil) (Or and Wraith, 2002). Resting spore

germination is delayed at soil moisture levels below 30% moisture content (Macfarlane, 1952),

though continuous levels of soil moisture are not required. Short intervals of high moisture

content following heavy rains are sufficient for infection, especially in the 2-3 weeks following

seeding (Thuma et al., 1983). Moisture levels affect the motility of zoospores within the soil

(Colhoun, 1973). However, pathogen development can be lower in saturated soils, when all

available pore spaces are filled with water (Gludovacz, 2013). There is a positive correlation

between total rainfall and clubroot incidence and severity (Gossen et al., 2012). Cortical infection

requires higher levels of soil saturation than is required for root hair infection (Dobson et al.,

1982). The interaction between soil moisture and temperature also plays a key role in clubroot

development (Donald et al., 2006).

 

1.3.2 Temperature

Temperature has a major influence on clubroot development, with a soil temperature of

18-25°C being most favourable for pathogen development, especially in the first 2-3 weeks of

plant growth (Colhoun, 1953; Buczacki et al., 1978; McDonald and Westerveld, 2008; Sharma et

al., 2011a, 2011b; Gossen et al., 2012). Soil temperature at 5-cm depth has a strong correlation

with clubroot severity and incidence (McDonald and Westerveld, 2008). Air temperature may

also be used as an effective indicator of clubroot development, in particular during the last 10

days before harvest for fast growing Asian Brassica vegetable crops such as Shanghai pak choy

Page 23: Effect of Plasmodiophora brassicae ... - University of Guelph

  16  (McDonald and Westerveld, 2008). The optimal temperature for germination of resting spores is

also 25°C, the temperature at which activity of the gene product Pro1 is highest and other

germination factors are stimulated (Dixon, 2009; Feng et al., 2010).  Clubroot development is

very slow at air temperatures below 17°C and above 30°C (Sharma et al., 2011a, 2011b). The

role of temperature in infection and symptom development has been confirmed in seeding date

trials that provide a range of temperatures for clubroot development under field conditions

(Gossen et al., 2012). Seeding date also affected clubroot development in short-season Brassica

vegetables, such as Shanghai pak choy, with the lowest incidence in May and September

plantings when the mean air temperature was lower during the early stages of plant growth

(Gossen et al., 2012). Although canola is a long-season crop with limited seeding date flexibility,

research has shown that seeding 1-3 weeks earlier in spring can decrease clubroot severity

(Hwang et al., 2011a, 2012a) because plants can become established during temperatures that are

too low for clubroot development.

 

1.3.3 Soil pH

Increased soil pH is directly correlated with decreased clubroot incidence and severity,

particular when soil pH is above 7.0 or 7.2 (Webster and Dixon, 1991a, 1991b; Donald and

Porter, 2004). Indeed, clubroot mitigation recommendations for vegetable crops often

recommend raising soil pH to 7.2 or above (OMAFRA, 2010). With adequate soil moisture and

optimal temperatures for clubroot development, however, moderately severe clubroot symptoms

developed up to a pH of 8 (Gossen et al., 2013). At temperatures below 17°C, there is a weaker

correlation between pH level and clubroot incidence and severity.

 

Page 24: Effect of Plasmodiophora brassicae ... - University of Guelph

  17  1.3.4 Resting spore concentration

There is a positive correlation between resting spore levels in the soil and clubroot

severity in susceptible Brassica crops (Dixon, 2006; Hwang et al., 2011a, 2012b). Depending on

environmental conditions, an inoculum concentration of 1000 spores g-1 dry soil is considered to

be the threshold for clubroot development in most susceptible cultivars (Murakami et al., 2002;

Donald and Porter, 2009; Faggian and Strelkov, 2009). The effect of resting spore concentration

on the severity of clubroot symptoms varies depending on the interaction with other

environmental factors, such as pH (Colhoun, 1953). Resting spores were distributed throughout

the soil profile to depths of 102 cm (Cranmer, 2015), although a previous report did not find

resting spores deeper than 45 cm below the soil surface with more than 97% of spores in the top

0-5 cm of soil (Kim et al., 2000).

1.4 Disease management  1.4.1 Cultural control

No single method provides substantial, consistent clubroot reduction, so an integrated

approach is essential for effective clubroot management (Donald et al., 2006; Diederichsen et al.,

2009). Recommended practices for sanitation and prevention of pathogen spread include nursery

hygiene for transplanted crops, avoidance of contaminated irrigation sources, and prevention of

cross-contamination by the movement of farm equipment from infested to non-infested fields

(Donald et al., 2006). Liming is a form of cultural control that has been commonly recommended

against clubroot in Brassica vegetable crops (Colhoun, 1953; Murakami et al., 2002). Small

increases in pH can suppress clubroot. However, at temperature and moisture levels that are

optimal for clubroot development, changes in pH are less effective at suppressing clubroot

Page 25: Effect of Plasmodiophora brassicae ... - University of Guelph

  18  (Gossen et al., 2013). Furthermore, liming is not an economically viable solution for canola

producers due to the high cost (Howard et al., 2010), and the effect can be inconsistent from year

to year (McDonald et al., 2004; McDonald and Westerveld, 2008). Increasing the pH level may

also have negative impacts on uptake of other nutrients (Hwang et al., 2008; Gossen et al., 2012)

and so may not be suitable for other crops in a rotational system (Hildebrand and McRae, 1998).

Boron and calcium have been found to suppress clubroot in Brassica vegetables (Webster

and Dixon, 1991a, 1991b), with a positive correlation between the rate of drench application of

boron and clubroot suppression (Deora et al., 2011). Field trials in high organic content soil

(Holland Marsh, Ontario) demonstrated that boron applied at a rate of 4 kg B ha-1 reduced

clubroot severity by 64% relative to the control (Deora et al., 2011). This was the most effective

rate with no phytotoxic effects, which occurred at rates above 2 kg B ha-1 in sand under

controlled conditions and 48 kg B ha-1 or greater in field conditions (Deora et al., 2011).

Phytotoxic effects of boron include chlorosis and necrosis. Boron is active against both the

primary and secondary stages of infection (Webster and Dixon, 1991b; Deora et al., 2011).

However, drench applications of boron are not considered to be economical for canola producers

due to the high water volume required for root drenches. Other application techniques may be

feasible, for example in combination with other fertilizers at sowing (Deora et al., 2011).

Calcium cyanamide (nitrogen and calcium oxide) has been used to suppress clubroot for

approximately 60 years, in addition to its uses as a slow-release nitrogen fertilizer and herbicide.

Calcium cyanamide (Perlka, 50% calcium oxide, 19.8% nitrogen, 1.5% magnesium oxide) was

effective in suppressing clubroot on Asian Brassica vegetables (McDonald et al., 2004). High

levels of limestone or wood ash (± quintozene fungicide) also suppressed clubroot, but the

Page 26: Effect of Plasmodiophora brassicae ... - University of Guelph

  19  authors concluded that soil amendments to change pH were not a practical or economically viable

option for canola producers (Hwang et al., 2011c).

Calcium application can delay pathogen maturation and reduce the total pathogen biomass

in root hair cells. At pH 5.5, a delay in pathogen development was observed only at the highest

rate of calcium application, compared to the control, while at pH 6.5, a delay was observed at all

rates of calcium (Webster and Dixon, 1991a). The effectiveness of calcium amendments

decreased as the pH increased above 6.5 to 8.0 (Donald and Porter, 2004).  

Production of susceptible crops in a short- to mid-term (2-5 year break from canola) crop

rotation is not a viable management option. After a single year of cropping with a susceptible

cultivar in a soil already infested with 1 x 106 resting spores per mL, the increase in spore load

was more than 60-fold (Hwang et al., 2013). In fields with high clubroot infestation, it took 17.3

years for the pathogen level to decrease below a detectable level (Wallenhammar, 1996). There

are, however, benefits to implementing short-term rotations in combination with clubroot-

resistant cultivars. A break from canola for two or more years reduced the concentration of

P. brassicae resting spores in the soil up to 10-fold and increased yield in resistant canola (Peng

et al., 2013). The concentration of P. brassicae resting spores declined by 98% from the initial

concentration following a 2-year break from susceptible canola (Peng et al., 2015). Increasing

cropping diversity in canola production in the absence of P. brassicae results in a yield increase

of 22% in a crop rotation with canola 1 in 6 years (Harker et al., 2014).

1.4.2 Synthetic fungicides and surfactants

Numerous fungicides have been tested against clubroot, including fluazinam (3-chloro-N-

[3-chloro-2,6-dinitro-4-(trifluoromethyl)phenyl]-5-(trifluoromethyl)-2-pyridinamine) (Allegro®

Page 27: Effect of Plasmodiophora brassicae ... - University of Guelph

  20  500F, ISK Biosciences Corporation), which is registered for management of clubroot on Brassica

vegetables in Canada, and cyazofamid (4-chloro-2-cyano-N,N-dimethyl-5-p-tolylimidazole-1-

sulfonamide) (Ranman® 400SC ISK Biosciences Corporation). Cyazofamid directly inhibits

resting spore germination, which in turns inhibits both root hair infection and pathogen

development (Mitani et al., 2003). Inhibition of root hair infection increases as the incubation

period with cyazofamid increased from 1 to 10 days (Mitani et al., 2003). Fluazinam disrupts the

oxidative phosphorylation metabolic pathway in mitochondria (Guo et al., 1991). Fluazinam and

cyazofamid provided 100% clubroot control on Shanghai pak choy under controlled conditions

(Adhikari, 2010). However, fluazinam and cyazofamid were not effective under high pathogen

populations in field trials (Tanaka et al., 1999; Saude et al., 2012; Peng et al., 2014), possibly due

to insufficient coverage over the entire soil volume, or degradation / leaching from the root zone

(Peng et al., 2014).

There are currently no fungicides registered in Canada for management of clubroot on

canola, and drench application would not be economically viable for canola producers (Howard

et al., 2010). Opportunities for fungicidal control, however, could include seed treatments or

fumigants, used only to treat localized areas (Donald and Porter, 2009), such as low-lying areas

with a higher spore concentration due to increased moisture, compaction and / or activity from

equipment carrying contaminated soil. Various surfactants have been evaluated for use against

clubroot (Humpherson-Jones, 1993; Hildebrand and McRae, 1998) though none have been

registered in Canada, possibly due to phytotoxicity (Howard et al., 2010).

Page 28: Effect of Plasmodiophora brassicae ... - University of Guelph

  21  1.4.3 Biological control and biofungicides

There are no proven, cost-effective biological controls for clubroot management at the

current time. However, there are some promising biological control agents (BCA) and

biofungicides (Peng et al., 2011; Kasinathan, 2012; Lahlali et al., 2013). A BCA is a living

organism that can be used to suppress pathogens by natural competition, parasitism and / or as a

microbial antagonist. A biofungicide is either a living organism or a compound derived from a

living organism that is toxic to fungal pathogens (Pal and McSpadden Gardener, 2006).

Biological control agents have been identified that cause induced systemic resistance

(ISR) to clubroot and other diseases. These BCAs protect the plant for a longer period of time

than synthetic fungicides (Peng et al., 2011). Microorganisms with activity against clubroot

include fungi such as Trichoderma, and bacteria such as Streptomyces (Cheah et al., 2000).

Commercial biofungicides with activity against clubroot available in Canada include Mycostop

(S. griseoviridis) (Verdera Oy, Finland), Actinovate (S. lydicus) (Natural Industries, USA), Root

Shield (T. harzianum) (BioWorks Inc. USA), Prestop (Gliocladium catenulatum syn.

Clonostachys rosea f. catenulate) (Verdera Oy) and Serenade (Bacillus subtilis QST713) (Bayer

CropScience, Germany).

Bacillus subtilis effectively colonized canola roots, a critical component to induced

resistance and antibiosis (Lahlali et al., 2013). B. subtilis resulted in small reductions in clubroot

severity as a soil drench and as a seed treatment in some, but not all, studies under controlled

conditions (Peng et al., 2011). Drench applications of Serenade and Prestop were moderately

effective in canola under controlled conditions (Kasinathan, 2012). Both Serenade and Prestop

were more effective than indigenous BCA isolates and as effective as synthetic fungicides under

low pathogen pressure (Peng et al., 2011). Neither Serenade nor Prestop effectively suppressed

Page 29: Effect of Plasmodiophora brassicae ... - University of Guelph

  22  clubroot in soil with high concentration of resting spores in field trials (Kasinathan, 2012; Peng et

al., 2014). Adequate soil moisture and distribution are required for B. subtilis to effectively

suppress clubroot. Depending on temperature in field conditions, primary and secondary

infections could take place over a period of 3 weeks in spring. Therefore, it can be challenging to

ensure adequate coverage of B. subtilis during the period of potential infection (Adhikari, 2010;

Peng et al., 2014).

Initial testing indicated that Actinovate was only effective at low inoculum levels, while

Mycostop was more consistently effective across inoculum levels (Adhikari, 2010). The efficacy

of each of these biofungicides varied greatly among trials and sites (Peng et al., 2014).

Acremonium alternatum is an endophytic fungus that enters the host plant through roots

and colonizes root and leaf cells. When applied to Arabidopsis thaliana, it reduced clubroot

severity by 50% and delayed development of P. brassicae (Jaschke et al., 2010). Heteroconium

chaetospira (Grove) M.B. Ellis is another endophyte that reduced clubroot severity by 52-97% in

field trials (Narisawa et al., 2000). A high rate of H. chaetospira reduced severity on canola

exposed to low levels of the pathogen, but was not effective when pathogen pressure was high.

When applied in combination with P. brassicae inoculation, H. chaetospira increased the activity

of phenylalanine ammonia lyases (PAL), and the expression of genes involved in the biosynthesis

of jasmonic acid, ethylene, auxin and PR-2 proteins (Lahlali et al., 2014). In broccoli

(B. oleracea var. italica), stimulation of defense response with salicylic acid activated

pathogenesis-related (PR) genes and reduced clubroot symptoms (Lovelock et al., 2013).

Two organic plant growth stimulants, Fructigard and PlasmaSoil (TILCO, Biochemie),

consisting of algal extracts, amino acids and phosponate, are effective against clubroot on napa

cabbage (B. rapa subsp. pekinensis) and canola (B. napus), but ineffective on Arabidopsis

Page 30: Effect of Plasmodiophora brassicae ... - University of Guelph

  23  thaliana. Plants with induced resistance exhibited the following differences: cells were more

densely packed in secondary phloem, large cells were retained in root cortex, and the outer

protective cell layer was thicker and impermeable (Kammerich et al., 2014). Individual

components of the formulation did not have the same effect.

1.4.4 Host plant resistance

 Resistant cultivars are an important component of effective, integrated clubroot control in

the Brassicaceae family (Diederichsen et al., 2009). Clubroot resistance genes have been found

in B. napus, B. oleracea and B. rapa (Hirai, 2006; Piao et al., 2009). The CRa gene in B. rapa

was the first clubroot resistant gene to be identified (Ueno et al., 2012). European fodder turnips

(B. rapa ssp. rapifera) carried multiple genes for resistance (Delourme et al., 2012). It is

common, however, for a single dominant resistance gene to be introduced to another crop, for

example in B. napus Swedish winter oilseed rape cultivar Tosca (Delourme et al., 2012). A single

major gene generally provides resistance to one specific pathotype, and resistance has been

overcome over time (Kuginuki et al. 1999; Dederichsen et al., 2009; Ueno et al., 2012). In

B. napus, resistance genes have been named Pb-Bn and PbBn. (Piao et al. 2009). To avoid

confusion, Piao et al. (2009) suggested standardized loci nomenclature for B. napus (PbBn),

B. rapa (PbBr) and B. oleracea (PbBo).

Canola cultivars with high levels of resistance to clubroot that are currently commercially

available in Canada include 45H29 (Pioneer Hi-Bred, Mississauga, ON), D3152 (DuPont

Canada, Mississauga, ON), Proven 9558C (Viterra, Regina, SK), 1960 and CS 2000 (Canterra

Seeds, Winnipeg, MB), 73-67 and 73-77 (Monsanto, Winnipeg, MB), 6056 CR (Brett Young,

Winnipeg, MB), V12-3 (Cargill Limited, Winnipeg, MB) and 14H1176 (Syngenta Canada, Inc.,

Page 31: Effect of Plasmodiophora brassicae ... - University of Guelph

  24  Guelph, ON). Resistance to pathotypes 2, 3, 5 and 6 was demonstrated in four commercial

cultivars of canola. These cultivars all had uniform responses at both the root hair and cortical

infection stages (Deora et al., 2013). Cultivar 45H21 (Pioneer Hi-Bred, Mississauga, ON) was

found to be resistant to pathotype 6 only. Resistance to pathotype 3 is currently essential for

commercial cultivars in many areas of Alberta, and effective resistance to the new strains of

P. brassicae in western Canada will be essential for clubroot management in this region over the

next decade.

Integrated management strategies for clubroot can improve the durability of genetic

resistance by reducing selection pressure. Durability of resistance refers to the effectiveness of

genetic resistance over time during prolonged and widespread use in the presence of the pathogen

under favourable environmental conditions (Johnson, 1984). In smaller pathogen populations, the

ability to adapt to plant host resistance is more limited because the likelihood of new genetic

variations arising from mutations or recombination is smaller (Fisher, 1930). Population size may

be reduced in a number of ways: decreasing the initial inoculum load, routine reductions, such as

fungicide applications, or limiting the movement of genotypes among pathogen populations,

which decreases the spread of mutant alleles and genotypes (McDonald and Linde, 2002).  More

durable resistance can be developed through efforts focused on developing broad-spectrum

resistance (Piao et al., 2009). Broad resistance to clubroot is defined as resistance against the

majority of P. brassicae pathotypes (Fu et al., 2011). This strategy could incorporate disrupting

pathogen selection for virulence by rotating cultivars with different major resistance genes over

time, when available or pyramiding several resistance genes in one cultivar (McDonald and

Linde, 2002; Ueno et al., 2012). For example, the new canola cultivar PV580GC (Crop

Page 32: Effect of Plasmodiophora brassicae ... - University of Guelph

  25  Production Services Canada, High River, AB), registered in October 2015, will be the first

multigenic clubroot-resistant canola cultivar available in Canada (CPS Canada, 2015).

1.4.5 Metabolic cost of resistance

A metabolic cost of disease resistance is the negative effect on growth and development

that occurs when a plant’s resources are reallocated toward defense in response to pathogen

recognition (Brown and Rant, 2013). It is also called a fitness cost of disease resistance (Brown

and Rant, 2013). Continuous low levels of expression of a major resistance gene can cause

slightly decreased productivity in plants even in the absence of the pathogen, although more

significant fitness costs are incurred when presence of the pathogen leads to induction of a

resistance response (Bergelson and Purrington, 1996). Induced resistance is the activation of

previously inactive resistance mechanisms, rather than the fabrication of new modes of resistance

that did not previously exist in that plant host, and relies on signals effectively triggering the

plant’s defensive response against an attacking pathogen (Van Loon et al., 1998). The cost of

disease resistance can be seen as a trade-off, because there is a greater cost of not expressing a

defense mechanism in the presence of a virulent pathogen.

In canola, root hair infection plays an important role in recognition and induction of

resistance to P. brassicae (Feng et al., 2012a). The metabolic cost of clubroot resistance may

result in decreasing vegetative growth or seed production, or delayed maturity (Hwang et al.,

2011a; Deora et al., 2012b; Peng et al., 2014). In western Canada, delayed maturity of canola is a

particular concern because it can result in increased susceptibility to frost damage, with

associated loss in quality as well as yield.

Page 33: Effect of Plasmodiophora brassicae ... - University of Guelph

  26  1.5 Summary and objectives  

Clubroot is an economically important disease of canola and Brassica vegetable crops in

Canada. Fungicides and biofungicides are not currently used in commercial canola production

because results have been inconsistent and they are not economically viable. Therefore, genetic

resistance is essential for clubroot management. The mechanisms of resistance to clubroot are not

well understood, and resistance has often been quickly overcome by the pathogen. Previous

research under controlled conditions has indicated that there may be a metabolic cost of

resistance to clubroot. In addition, the effect of resting spore concentration on the severity of

clubroot symptoms varies depending on the interaction with other environmental factors. Also,

plant growth and yield increase in response to increased diversity in crop rotation, compared to

continuous canola. More research on the effect of short- to mid-term (2-5 year break from canola)

crop rotation on concentration of resting spores in the soil and on metabolic cost of resistance

could have important implications for the development of integrated clubroot management

systems for canola.

The objectives of this research were to:

1. Examine the effect of concentration of P. brassicae resting spores on growth and

development of clubroot-resistant canola and Brassica vegetables, to identify if there is a

metabolic cost of resistance.

2. Examine the effect of crop rotation on the rate of decline of resting spores over time and

the concentration of resting spores in soil.

3. Examine the interaction between crop rotation and inoculation with P. brassicae on

growth of clubroot-resistant canola.

Page 34: Effect of Plasmodiophora brassicae ... - University of Guelph

  27  4. Assess the interaction between pH and inoculum concentration of P. brassicae on

clubroot incidence and severity of susceptible canola.

The following hypotheses were tested:

1. Plant height and biomass are reduced and development is delayed by increasing

concentration of resting spores in resistant plants.

2. The concentration of resting spores in soil declines exponentially with increasing length

of break interval following a susceptible canola crop.

3. The interaction of crop rotation and inoculation with P. brassicae results in a greater cost

of resistance to clubroot in soil with a history of continuous canola compared to a more

diverse crop rotation.

4. At low concentrations of resting spores, clubroot incidence is reduced at pH >7.0

compared with pH 6.0 to 7.0, but there is no effect of pH at high concentrations of spores.

Page 35: Effect of Plasmodiophora brassicae ... - University of Guelph

  28  CHAPTER TWO

 EFFECT OF RESTING SPORE CONCENTRATION ON GROWTH OF

CLUBROOT-RESISTANT BRASSICA CROPS

2.1 Introduction  

Inoculum concentration in soil is an important factor influencing infection by soil-borne

pathogens (Richardson and Munnecke, 1964). Disease severity increases with increasing

inoculum concentration in many susceptible plant-pathogen interactions, including Fusarium wilt

of lettuce (Lactuca sativa L.) caused by Fusarium oxysporum f. sp. lactucae (Scott et al., 2012),

root rot in lentil (Lens culinaris Medikus) caused by Rhizoctonia solani Kühn (Chang et al.,

1998) and pre-emergence damping off in field pea (Pisum sativum L.) caused by Pythium

irregulare Buisman (Richardson and Munnecke, 1964). Reduction of inoculum concentration in

soil is the focus of many disease management strategies, including soil solarization, fumigation

and crop rotation with non-host or bait crops. For example, soil solarization reduces fusarium wilt

of lettuce by up to 91% (Matheron and Porchas, 2010). Fumigation with methyl bromide plus

chloropirin reduces Verticillium dahliae Kleb. microsclerotia by 93% (Short et al., 2015) and

crop rotation reduces inoculum concentration of F. oxysporum f. sp. lactucae by 86% within 12

months in the absence of a host crop (Scott et al., 2012).

Genetic resistance is essential for management of many soil-borne pathogens, including

F. graminearum in wheat (Triticum aestivum L.) (Buerstmayer et al., 2009), Sclerotinia

sclerotiorum (Lib.) de Bary in soybean (Glycine max (L.) Merr.) (Cober et al., 2003) and

P. brassicae in canola (Cao et al., 2009). However, resistance to plant pathogens can be

associated with a metabolic cost to the plant, also called the fitness cost of disease resistance.

This can result in a negative association between disease resistance and other traits, such as plant

Page 36: Effect of Plasmodiophora brassicae ... - University of Guelph

  29  maturation and yield. The negative effect of the reallocation of resources from growth to defense

during the disease resistance response can be quantified using measurable indicators of plant

fitness. Bergelson and Purrington (1996) reviewed 88 published studies and found that overall

about 50% of plant species surveyed had a significant fitness cost of disease resistance.

The effect of inoculum concentration on resistant crops is generally not well understood

because fitness costs of disease resistance are challenging to identify and to quantify. There may

be several physiological and genetic factors contributing to the overall cost of resistance (Brown

and Rant, 2013). Yield is determined by numerous traits and genetic components (Campbell and

Kondra, 1978; Foulkes et al., 2000). Indicators such as vegetative growth, leaf area and pod

development are useful in the assessment of cost of resistance because they make important

contributions to yield of canola (Campbell and Kondra, 1978; Freyman et al., 1973; Krogman

and Hobbs, 1975, Chongo and McVetty, 2001). A fitness cost can be incurred either from the

presence of the resistance gene, or from expression of defense by the resistant plant upon

recognition of the pathogen (Brown and Rant, 2013). Constitutive defenses, which are always

present in the plant, have continuous low levels of expression. This leads to a mean fitness cost of

3.5% reduced productivity compared to susceptible plants, averaged across many different crops

(Bergelson and Purrington, 1996). Induced resistance is believed to incur less metabolic cost

overall, compared to constitutive resistance, because the costs are reduced (though not

eliminated) in the absence of pests (Purrington, 2000; Walters et al., 2013). The overall cost of

resistance can be seen as a trade-off, because there is a greater cost of not expressing a defense

mechanism in the presence of a virulent pathogen.

The trade-off between plant growth and defense against pathogens involving the

jasmonante-signaling pathway is a widely conserved defense strategy in angiosperms. It has been

found that gibberellic acid activates defense and also mediates a delay in the degradation of

Page 37: Effect of Plasmodiophora brassicae ... - University of Guelph

  30  growth-response proteins in rice (Oryza sativa L.) and Arabidopsis thaliana (L.) Heynh. (Yang et

al., 2012). In Arabidopsis, the PMR6 gene codes for a protein required for susceptibility to

powdery mildew (Erysiphe cichoracearum DC.) and may also influence plant growth and

development and pectin degradation (Vogel et al., 2002). A pmr6-1 mutant with loss of function

for the PMR6 gene (resistant to powdery mildew) is 23% smaller in rosette diameter and has

increased pectin in cell walls relative to the susceptible control. Pectin in cell walls may play a

role in resistance to powdery mildew (Vogel et al., 2002). Model plant-pathogen systems using

mutant constructs with loss of function for an identified resistance gene could improve

understanding of metabolic cost of resistance to clubroot.

In commercial canola cultivars, a major challenge for comparison of fitness costs is that

near-isogenic lines, differing only in presence or absence of resistant gene, are not available.

Instead, comparison of fitness must be conducted between induced and non-induced plants to

assess the overall costs of induced resistance (Purrington, 2000). Studies of other plant-pathogen

interactions demonstrate that induced resistance costs are commonly related to a hypersensitive

response (HR) (Boyd et al., 1995; Tian et al., 2003). However, in commercial canola cultivars,

HR is not the mechanism of clubroot resistance (Deora et al., 2013).

In response to clubroot infection, resistant canola (B. napus) roots form a ring of

concentrated reactive oxygen species (ROS) in the inner cortex. ROS accumulation is a response

to wounding, which reinforces cell wall proteins and limits P. brassicae invasion of vascular

tissues during secondary colonization (Deora et al., 2013). Enzymes that metabolize ROS, such

as copper / zinc superoxide dismutase cytochrome c oxidase, decreased in susceptible canola

within the first 12 hours after inoculation, and then increased 24-72 hours after inoculation (Cao

et al., 2008). Indole-3-acetic acid (IAA) increases continually in resistant cultivars of napa

cabbage (B. rapa subsp. pekinensis) (Ludwig-Muller et al., 1993) and canola (Feng et al., 2012).

Page 38: Effect of Plasmodiophora brassicae ... - University of Guelph

  31  Root hair infection plays an important role in recognition and induction of resistance to

P. brassicae in host plants (Siemens et al., 2002; Feng et al., 2012; McDonald et al. 2014).

Clubroot symptoms do not develop and resting spores are not produced during an incompatible

interaction (Ludwig-Muller et al., 1999; Feng et al., 2012). However, root hair infection may take

place on host, non-host and resistant plants (Feng et al., 2012).

Increasing the resting spore concentration of P. brassicae from 1 x 105 to 1 x 108 spores

g-1 soil increases clubroot severity and decreases plant height and seed yield in susceptible canola

(Brassica napus L.) cultivars (Hwang et al., 2011a). In a recent study, four clubroot-resistant

canola cultivars (45H29, 73-67, 73-77 and Proven 9558C) inoculated with P. brassicae exhibited

reductions in plant height, number of flowers and pods, and delayed progression from the

vegetative to reproductive stage (Deora et al., 2013). These findings were consistent with two

previous studies showed that seedling emergence, height and yield were negatively correlated

with increasing resting spore concentration in a resistant canola cultivar (Hwang et al., 2011a;

Deora et al., 2012b). In western Canada, delayed maturity can lead to late canola crops, which

may then be damaged by frost.

In a susceptible cultivar, 1000 spores g-1 soil is commonly cited as the infection threshold

(Donald and Porter, 2009). However, the effect of resting spore concentration on the severity of

clubroot symptoms varies depending on the interaction with other environmental factors, such as

pH (Colhoun, 1953). Root hair infection (RHI) declines above pH 6.5 (Myers and Campbell,

1985; Webster and Dixon, 1991; Donald and Porter, 2004; Gossen et al., 2013). Development of

primary plasmodia is inhibited at pH ≥7.2 prior to secondary zoospore release (Myers and

Campbell, 1985). The interaction between pH and resting spore concentration is not well

understood. Maximum root hair infection in slightly acidic soil (pH 6-6.5) occurs at temperatures

near 25°C (Gossen et al., 2013). Severe clubbing occurs at resting spore concentrations as low as

Page 39: Effect of Plasmodiophora brassicae ... - University of Guelph

  32  103 at pH 6.3, but resting spore concentration is positively correlated with clubroot incidence in

alkaline soil (pH 7.8) (Colhoun, 1953). Increased understanding of this interaction could make it

easier to predict clubroot severity and improve recommendations regarding use of resistant

cultivars or other management practices.

Clubroot management that is sustainable for several years requires effective strategies for

reducing the concentration of resting spores in soil. To date, the implementation of cultural,

biological and chemical strategies have been ineffective (Ahmed et al., 2011) or not cost-

effective (Porter et al., 1991; White and Buczacki, 1977). In fields with a high initial level of

inoculum, the concentration of resting spores may remain above the infection threshold for a

susceptible crop for many years (Wallenhammer, 1996), even though there has been a significant

reduction in the number of resting spores (Peng et al., 2014).

There were two objectives of this research. The first was to identify and quantify the

metabolic cost of resistance to P. brassicae by examining the effect of the concentration of

resting spores of P. brassicae on growth and development of clubroot-resistant canola and

Brassica vegetable cultivars. It is hypothesized that increasing inoculum concentration will

reduce plant growth in canola, napa cabbage and cabbage plants and delay development in

canola. The second objective was to assess the interaction between pH and resting spore

concentration of P. brassicae on clubroot incidence and severity in susceptible canola. It is

hypothesized that pH will not affect clubroot incidence at high concentrations of resting spores,

but clubroot incidence at low concentrations of resting spores will be reduced at pH >7.0

compared with pH 6.0 to 7.0.

Page 40: Effect of Plasmodiophora brassicae ... - University of Guelph

  33  2.2 Materials and Methods 2.2.1 Controlled environment study – canola

Growth room studies were conducted using canola cultivar 45H29 (Pioneer Hi-Bred Ltd,

Chatham, ON), which is resistant to pathotype 6 (Deora et al., 2013) of P. brassicae (Williams’

system). The seeds were sown in tall plastic pots (conetainers) filled with soil-less media

(Sunshine Mix #4, Sun Gro Horticulture Canada Ltd., Agawam, MA). Two seeds were planted

per pot and thinned to one seedling per pot. Four replicates of 10 pots each of the clubroot-

susceptible canola line ACS-N39 (AAFC, Saskatoon SK) were included as a susceptible control

(Gludovacz, 2013). The plants were maintained at 25°/20° C day/night, with 16-hr photoperiod

and 65% relative humidity, and fertilized with 1 g L-1 N-P-K (20-20-20) and 1 g L-1 magnesium

sulphate solution at 2–3 day intervals. The plants were watered with deionized water adjusted to

pH 6.0 using commercial white vinegar. Each 9-day-old seedling was inoculated with 5 mL of

resting spore suspension of P. brassicae pathotype 6 in the following treatments: 0, 1 x 104,

1 x 105, 1 x 106, 1 x 107 and 1 x 108 spores mL-1. Inoculum was prepared following a standard

protocol (Sharma et al., 2011). Briefly, clubbed roots were washed and soaked in deionoized

water, and 10 g of gall was homogenized in 300 mL deionized water for 2 min in an electronic

blender. The mixture was filtered though eight layers of cheesecloth and a haemocytometer was

used to estimate the resting spore concentration. The resulting spore suspension was diluted to

create the required treatment concentration.

In each replicate, the height of each plant was assessed at 7-day intervals from 2 to 6

weeks after inoculation (WAI). In repetition 2, the height measurement for week 5 was missed by

mistake. At 10 weeks after planting, plants were harvested and weighed, and roots were assessed

for clubroot incidence (%) and severity using the standard 0–3 rating scale, where 0 = no

Page 41: Effect of Plasmodiophora brassicae ... - University of Guelph

  34  clubbing, 1 < 1/3 of root clubbed, 2 = 1/3–2/3 of roots clubbed and 3 > 2/3 of roots clubbed

(Strelkov et al., 2006). A Disease Severity Index was calculated according to Crete et al. (1963):

DSI = ∑ [(c [(class no.)(no. of plants in each class)] x100 [(total no. plants per sample)(no. classes - 1)]

Plant height was  measured from hypocotyl (at soil surface) to shoot apex. Plant height

measurements from the controlled environment studies were compared at individual time-points

and also for overall growth over time. Growth over time was analyzed using area under the

growth stairs (AUGS). The original model for this analysis is area under the disease progress

curve (AUDPC), which is commonly used as a quantitative measure of disease over time (Van

der Plank, 1963). It is calculated from multiple disease assessments over time (Jeger and

Viljanen-Rollinson, 2001). In the current study, plant height assessments were used instead of

disease assessment at each time-point. In area under the growth curve (AUGC), the first and last

time-points are only weighted at 50%, because the trapezoid only extends in one direction from

the midpoint. Area under the Growth Stairs (AUGS) was used in this study, which provides equal

weighting of all time-points to summarize season-long growth into a single value, calculated as

follows:

AUGS =Y!   +  Y!!!

2  !!!

!!!

×   t!!! −   t! +Y! +  Y!

2  ×  t! − t!n− 1  

where Yi is plant height in cm at the ith observation, ti is time in days after inoculation at the ith

observation, and n is the total number of observations (Simko and Piepho, 2012).

Page 42: Effect of Plasmodiophora brassicae ... - University of Guelph

  35  2.2.2. Large pot studies – outdoors

Outdoor studies in large pots were conducted as a randomized complete block design

(RCBD) with four replicates per treatment at Bajar Farm, near the Muck Crops Research Station,

King, Ontario. The clubroot-resistant canola cv. 45H29 (Pioneer Hi-Bred, Caledon, ON) and

ACS-N39 (AAFC, Saskatoon, SK), a susceptible canola check, were seeded into 200-cell plug

trays on 16 July 2014 and grown in a greenhouse. The seedlings were hand transplanted into

plastic pots (30 cm × 27.5 cm dia.) containing dark grey gleysol-Granby sandy loam soil

(Westerveld, 2005) from the Bajar farm site on 06 August 2014, at 3 weeks after seeding. When

the trial was conducted in 2015, one repetition was seeded on 09 June and another on 24 June.

The seedlings were hand transplanted into pots containing 75% mineral soil from the Bajar farm

and 25% soil-less media (Sunshine Mix #4, Sun Gro Horticulture Canada Ltd., Agawam, MA) on

23 June and 08 July, at 2 weeks after seeding. The soil-less media was mixed into the field soil

sample using an electrical cement mixer.

The soil pH in each pot was lowered from 7.8 to 6.4 by application of 7.3 g 90% sulphur

chips, 2.9 g nitrogen sulphate, and 1.0 g magnesium sulphate and 400 mL water pH adjusted to

4.0 with phosphoric acid. The pots were watered once per week with water adjusted to pH 6.0

using commercial vinegar and N-P-K (20-20-20) solution. Inoculum was prepared as described

previously (Sharma et al., 2011). Each pot was inoculated at transplanting with 450 mL of spore

suspension of pathotype 6 to produce a final concentration of 0, 1 x 103, 1 x 104, 1 x 105, 1 x 106

and 1 x 107 resting spores g-1soil in the top 15 cm of the pot. In 2014, each pot was inoculated at

6 weeks before transplanting and again at transplanting. In 2015, each pot was inoculated one

time at transplanting.

The height of each resistant plant was assessed at weekly intervals starting at 2 WAI. At 8

WAI, the developmental stage (vegetative, bud, flowering, pod development) of each plant was

Page 43: Effect of Plasmodiophora brassicae ... - University of Guelph

  36  assessed. Plants were then harvested, weighed and roots assessed for clubroot incidence (%) and

severity using a standard 0-3 rating scale, as previously described.

Each experimental unit consisted of two pots with five plants per pot, except for the

treatments in the fourth replicate in 2014, which only had one pot per unit because there were not

enough seedlings for transplant. Biomass assessments included five plants per experimental unit.

Resistant canola plants that exhibited clubbing symptoms were removed from the biomass

assessments in the second repetition in 2015, but were included in all of the plant height

assessments in both years and in the biomass assessments for the first repetition in 2015.

2.2.3 Controlled environment study of canola, cabbage and napa cabbage

Three clubroot-resistant canola cultivars, 45H29, 73-67 (Monsanto Canada Inc.,

Winnipeg, MB) and 73-77 (Monsanto Canada Inc.), and the susceptible line ACS-N39 were

sown in tall plastic conetainers filled with mineral soil from Elora Research Station (Elora, ON).

In companion studies, three clubroot-resistant napa cabbage cultivars, China Gold (Sakata Seed

Corporation, Morgan Hill, CA), Yuki (Sakata Seed Corporation) and Emiko (Bejo Seeds Inc.,

Oceano, CA), and the susceptible cultivar Mirako (Bejo Seeds Inc.) and three clubroot-resistant

cabbage cultivars, Kilaherb (Syngenta Seeds, Inc., Minnetonka, MN), Kilaton (Syngenta Seeds,

Inc.), and Tekila (Syngenta Seeds, Inc.), and the susceptible cultivar Bronco (Bejo Seeds Inc.)

were assessed.

For each study, two seeds were planted per pot and thinned to one seedling per pot. Four

replicates of 10 pots were included for each treatment. The plants were maintained at 25°/20° C

day/night, with 16-hr photoperiod and 65% relative humidity, and fertilized with 1 g L-1 N-P-K

(20-20-20) and 1 g L-1 magnesium sulphate at 2–3 day intervals. The plants were watered with

deionized water adjusted to pH 6.0 using commercial white vinegar. Each 10-day-old seedling

Page 44: Effect of Plasmodiophora brassicae ... - University of Guelph

  37  was inoculated with 5 mL of resting spore suspension of 1 x 106 resting spores mL-1 of

P. brassicae pathotype 6. Inoculum was prepared as described above. In each replicate, the height

of each plant was assessed at 7-day intervals starting 2 WAI. Each study was repeated with the

following modifications: 25% soil-less media (Sunshine Mix #4, Sun Gro Horticulture Canada

Ltd., Agawam, MA) was incorporated into the soil blend with 75% mineral soil, and the mixture

was treated with water adjusted to pH 4.0 using phosphoric acid to lower the pH of the soil at 2

weeks prior to planting.

2.2.4 Field trials

Field trials were conducted to compare the growth of clubroot-resistant canola and napa

cabbage cultivars at two adjacent sites at the Muck Crops Research Station, Holland Marsh,

Ontario. Both sites were on organic soil, a typic humisol-muck (McDonald et al., 2008), and

received the same agronomic treatments (tillage and fertilization) and had a similar crop history,

but differed in concentration of resting sproes of P. brassicae (Table 2.1).

In one trial, three clubroot-resistant canola cultivars, 45H29, 73-67 and 73-77, and the

susceptible line ACS-N39 were sown on 26 June, 2014. In a companion trial, three clubroot-

resistant napa cabbage cultivars, China Gold, Yuki and Emiko, and the susceptible cultivar

Mirako were sown on 16 July, 2014. Each trial was laid out in a randomized complete block

design with four replicates. Both crops were direct seeded at a rate of approximately 18 seeds per

m of row using an Earthway push seeder fitted with an Earthway 1002-9 disc. Each canola plot

was 6 m in length and each napa cabbage plot was 3 m in length. Napa cabbage plants were

thinned to approximately 25 cm apart within rows. Each plot consisted of a single row, spaced

~26 cm apart. In each replicate, a representative sample of 10 consecutive plants starting 2-3

plants in from the beginning of the row was measured repeatedly at 7-day intervals starting 4

Page 45: Effect of Plasmodiophora brassicae ... - University of Guelph

  38  weeks after seeding. Plant growth was assessed in canola by measuring plant height from the

hypocotyl to the shoot apex. Napa cabbage plant growth was assessed by measuring the length of

the third and fourth youngest leaves. AUGS was calculated as previously described.

Table 2.1 Soil properties of field trial sites at the Muck Crops Research Station at Holland

Marsh, Ontario in 2014 and 2015.

Site ID Spores

g-1 soil pH Soil analysis Texture

P K Mg OM

2014-Low 7 x 105 6.3 67 240 381 78 Muck

2014-High 7 x 106 6.5 52 195 408 75 Muck

2015-Low 3 x 106 6.1 79 291 493 77 Muck

2015-High 1 x 107 6.3 84 100 463 79 Muck

2015-BDL1 BDL 6.0 112 548 495 82 Muck

1BDL = Below Detection Limit

At 9 weeks after planting, the developmental stage (vegetative, bud, flowering, pod

development) of each canola plant was assessed. Plants were harvested and weighed, and roots

were assessed for clubroot incidence (%) and severity using a standard 0-3 rating scale, as

previously described.

Both trials were repeated in 2015, with the addition of an adjacent site in each trial with

inoculum concentration below detectable levels (Table 2.1). All plots were 6 m in length, and

each plot for both canola and napa cabbage was seeded on 09 July 2015 to allow for comparison

Page 46: Effect of Plasmodiophora brassicae ... - University of Guelph

  39  across crop species. Also, the susceptible napa cabbage cultivar Mirako was replaced with

Suzuko B-2961 (Bejo Seeds, Inc.) due to discontinuation of Mirako.

2.2.5 Controlled environment pH study

 Growth room studies were conducted using the susceptible line ACS-N39 in a factorial

design with two factors. One factor was inoculum concentration (0, 1 x 103, 1 x 105, 1 x 107

spores mL-1) and the other factor was pH (5.5, 6.0, 6.5, 7.0, 7.5). The experimental design was a

RCBD with three replicates. Three plants were grown per experimental unit as outlined in the

protocol by Kasinathan (2012) for assessment of root hair colonization. Plant culture followed the

protocol outlined by Kasinathan (2012), which were adapted from Donald and Porter (2004).

Seeds were sown in moist sand in Petri dishes at 20°C with 16-h photoperiod. When the

cotyledons were fully expanded, individual seedlings were transplanted into 50-mL Falcon tubes

(Fisher Scientific, Markham, ON) containing autoclaved, non-calcareous sand at pH 6.5

(Hutcheson sand, Hutcheson Sand Mixes, Huntsville, ON). The plants were fertilized with

15:15:18 N:P:K and ammonium sulphate. A stock solution was prepared using 40 g of N:P:K and

20 g of ammonium sulphate in 1 L of water. A 5-mL aliquot of stock solution was added to 1 L

of deionized water and the pH was adjusted before watering as described below. The plants were

watered with deionized water adjusted to the desired pH using biological buffers (Myers and

Campbell, 1985): PIPES [piperazine-N, N’-bis-(2-ethanesulfonic acid,) monosodium salt, mono

hydrate], MES [2-(N-morpholino) ethanesulfonic acid, sodium salt] and HEPES [N-2-

hydroxyethylpiperazine-N’-2-ethanesulphonic acid] (Robiot Canada, Toronto, ON). Stock

solutions were prepared and stored at 5°C. The pH meter (Hanna instruments, Woonsocket, RI)

was calibrated before each use, using standard buffer solutions (Fisher Scientific, Nepean, ON).

Page 47: Effect of Plasmodiophora brassicae ... - University of Guelph

  40  To achieve the desired pH of the sand growth medium, solution of the target pH was added to

each Falcon tube. The water that drained from the Falcon tubes was collected and the pH of the

drain solution was measured for each treatment. Based on this result, the pH of the watering

solution was adjusted up or down and the watering solution was re-applied until the desired pH

was attained in the drain solution. Drainage holes were made at the bottom of each Falcon tube to

avoid water logging.

Each 10-day-old seedling was inoculated with 1 mL of resting spore suspension.

Inoculum was prepared following the protocol outlined by Sharma et al. (2011) as described

above. Plants were harvested at 8, 15 and 22 days after inoculation (DAI). Roots (including root

hairs) were sampled from all three plants in one experimental unit for one biological replicate.

Clubroot incidence and severity were assessed using the 0–3 scale previously described. The

roots were rinsed with deionized water, followed by surface sterilization in 10% bleach, rinsed

three times in deionized water to remove debris from the root surface, cut into 1-cm pieces,

frozen in liquid nitrogen, and stored at -80oC until use, as described in Lahlali et al. (2011). This

study was not repeated.

2.2.6 Statistical analysis

All analyses were conducted using SAS version 9.3. The data for the controlled

environment study and large pot outdoor study were analyzed in a mixed model analysis of

variance using single-degree of freedom contrasts using PROC MIXED and PROC GLM. Spore

concentration was a fixed effect and block and repetition were random effects.

The data for field trials and studies of canola, napa cabbage and cabbage were analyzed as

a mixed model analysis of variance using single degrees of freedom and the slice option to

Page 48: Effect of Plasmodiophora brassicae ... - University of Guelph

  41  examine the interaction between cultivar and site (spore concentration) with PROC MIXED and

PROC GLM. Levene’s test was also used to test for homogeneity of variance across sites.

Variance in plant height, biomass and maturity of resistant canola across sites within each year

was homogeneous (Tables A1.35-37, A1.41-43, A1.50-52), so comparisons were made among

sites for those response variables. Variance in plant height was homogenous across sites in 2015

but not in 2014. Therefore, plant height, biomass and maturity provide a more robust indication

of cost of resistance to clubroot at various concentrations of resting spores than plant height in the

2014 field trial. There was homogeneity of variance in CI and DSI for susceptible canola among

sites within year. Variance was not homogeneous across years for any variable, so each year was

analyzed separately. Means separation were conducted using Tukey’s test at P = 0.05 level of

significance. The fixed effects in the mixed model analysis were cultivar, crop and spore

concentration and random effects were block and repetition.

The data for the pH study were analyzed in a mixed model analysis of variance, with

spore concentration and pH as fixed effects, and block and repetition as random effects. Clubroot

incidence and disease severity index for the pH treatments at each concentration of resting spores

were compared using Tukey’s procedure as described by Kasinathan (2012).

2.3 Results 2.3.1 Controlled environment study

There were high rates of clubroot incidence and severity in the susceptible control and no

clubroot symptoms in resistant canola plants in both repetitions under controlled conditions

(Table 2.2). One additional repetition was conducted, but it was not included in the analysis due

to inconsistent plant growth and signficant block effect (P <0.0001) (see data in appendix).

Page 49: Effect of Plasmodiophora brassicae ... - University of Guelph

  42  Table 2.2 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible

canola ACS-N39 inoculated with 1 x 106 spores mL-1 under controlled conditions1.

Repetition CI DSI

1 100 ns2 71 b3

2 100 95 a 1 N.B. There were no clubbing symptoms in any resistant plants (data not shown).

2 ns = not significant. 3 Columns with the same letter do not differ at P = 0.05, based on Tukey’s multiple means

comparison test.

Plant height of canola cv. 45H29 (resistant) at 6 WAI declined in a quadratic relationship

(y = -0.086x2 + 0.129x + 31.7, R² = 0.21) with increasing concentration of resting spores

(P = 0.005) across the two repetitions (Fig. 2.1). Plant height was reduced by approximately 13 %

at conecentrations > 1 x 106 spores mL-1 compared to the non-inoculated control at 6 WAI, based

on the regression equation. There was no repetition effect on plant height at 6 WAI and therefore

data were analyzed across repetitions.

The biomass of resistant canola (dry shoot weight) declined in a weak linear relationship

(y = 0.189x + 6.47, R2 = 0.07, P = 0.03) with increasing inoculum in the first repetition. The

proportion of resistant canola plants at the seedpod development stage at 8 WAI declined in a

quadratic relationship (y = -1.63x2 + 9.37x + 78.2, R2 = 0.60) with increasing concentration of

resting spores (P = 0.007) in the first repetition (Fig. 2.2). The regression was not significant for

biomass or maturity in the second repetition.

Page 50: Effect of Plasmodiophora brassicae ... - University of Guelph

  43  

Figure 2.1 Plant height of clubroot-resistant canola at 6 weeks after inoculation with increasing

concentrations of Plasmodiophora brassicae resting spores under controlled conditions (two

repetitions, n=8, P = 0.005).

   

Figure 2.2 Plant maturity at harvest in canola grown in controlled conditions in response to

increasing concentration of Plasmodiophora brassicae resting spores (P = 0.007) in the first

repetition.

y  =  -­‐0.086x2  +  0.129x  +  31.7  R²  =  0.21,  P  =  0.005  

0  

5  

10  

15  

20  

25  

30  

35  

40  

0   1   2   3   4   5   6   7   8  

Plant  height  (cm

)  

Resting  spore  concentration,  log  scale  (spores  mL-­‐1)  

y  =  -­‐1.63x2  +  9.37x  +  78.2  R²  =  0.60,  P  =  0.008  

0  10  20  30  40  50  60  70  80  90  100  

0   1   2   3   4   5   6   7   8  

Plants  at  pod  stage  (%

)  

Resting  spore  concentration,  log  scale  (spores  mL-­‐1)  

Page 51: Effect of Plasmodiophora brassicae ... - University of Guelph

  44  2.3.2 Large pot studies – outdoors

Clubroot incidence and severity in the susceptible control was low in 2014, but high in

both repetitions in 2015 (Table 2.3). There were no clubbing symptoms in resistant canola in

2014, but there were low levels of clubroot incidence and severity in both repetitions in 2015

(Table 2.3). There was no effect of inoculation on plant height, maturity or biomass among the

treatments in 2014 or 2015 (data not shown).

In the 2 weeks following transplanting of canola in 2014, mean air temperature was 25°C

maximum and 12°C minimum. In 2015, mean air temperature was 23°C maximum and 15°C

minimum in the first repetition and 27°C maximum and 14°C minimum in the second repetition.

Table 2.3 Clubroot incidence (CI) and severity (disease severity index, DSI) in canola breeding

line ACS-N39 (susceptible check) and resistant cultivar 45H29 in an outdoor trial using large

pots near Bradford, ON, 2014 and 2015.

ACS-N39 45H29

Assessment CI DSI CI DSI

2014 35 b1 16 c 0 b 0 ns2

2015 Repetition 1 98 a 78 b 8 ab 6

2015 Repetition 2 100 a 97 a 15 a 12 1 Means in a column followed the same letter do not differ at P = 0.05, based on Tukey’s multiple

means comparison test. 2 ns = not significant.

Page 52: Effect of Plasmodiophora brassicae ... - University of Guelph

  45  2.3.3 Controlled environment study – canola, napa cabbage and cabbage

 Clubroot incidence and severity were low in canola and moderate in napa cabbage and

cabbage in the first repetition of the study under controlled conditions. Clubroot symptoms were

low in cabbage and moderate in canola and napa cabbage in the second repetition (Table 2.4).

In resistant napa cabbage, inoculation with P. brassicae resting spores reduced leaf length

at 5 WAI by 11.4% (±6.5) compared with the non-inoculated control (P = 0.005) across

repetitions, but there was no effect of inoculation on plant height of resistant canola or leaf length

of resistant cabbage. There were no differences in the timing of maturity in resistant canola or in

biomass of any crop species (data not shown).

Table   2.4 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible

canola, napa cabbage and cabbage inoculated with 1 x 106 spores mL-1 P. brassicae, under

controlled conditions. 1

Repetition

Canola Napa cabbage Cabbage

CI DSI CI DSI CI DSI

1 18 b2 6 b 39 a 20 ns3 52 a 20 a

2 63 a 35 a 30 ab 21 23 b 10 ab 1 N.B. There were no clubbing symptoms in the non-inoculated control or in any resistant plants

(data not shown). 2 Columns with the same letter do not differ at P = 0.05 based on Tukey’s multiple means

comparison test. 3 ns = not significant.

Page 53: Effect of Plasmodiophora brassicae ... - University of Guelph

  46  2.3.4 Field trials

 The concentration of resting spores site at the low spore concentration site in 2014 was

90% lower than the concentration of resting spores at the high spore concentration site. However,

spore concentrations at both sites were much higher than the infection threshold for susceptible

cultivars (103) (Faggian and Strelkov, 2009). In 2014, there was 100% CI and DSI in the

susceptible canola at both sites. Susceptible napa cabbage had 23% CI and 13% DSI at the low

spore concentration site, and 90% CI and 44% DSI at the high spore concentration site (Table

2.5). There were no clubbing symptoms in resistant canola or napa cabbage in 2014.

Table 2.5 Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible

canola and napa cabbage grown in field soil at Muck Crops Research Station, 2014.

Canola Napa cabbage

Site ID Spores

(g-1 soil)

CI DSI CI DSI

2014-High 7 x 106 100 ns1 100 ns 90 a2 44 a

2014-Low 7 x 105 100 100 23 b 13 b 1 N.B. There were no clubbing symptoms in the non-inoculated control or in any resistant plants

(data not shown). 2 ns = not significant. 3 Columns with the same letter do not differ at P = 0.05, based on Tukey’s multiple means

comparison test.

In 2015, there was 100% CI in the susceptible canola cultivar at the high and low spore

concentration sites, and 45% at a site with spore concentration below detectable limits. DSI in

susceptible canola was 70% at the high spore concentration site, 65% at the low spore

Page 54: Effect of Plasmodiophora brassicae ... - University of Guelph

  47  concentration site and 19% at the site with spore concentration below detectable limits (Table

2.6). Resistant canola in 2015 at all three sites developed low levels of clubroot symptoms (Table

2.6). There were no clubbing symptoms in any napa cabbage cultivars in 2015. This indicates that

in these conditions there was no susceptible control for napa cabbage. There were no clubroot

symptoms observed in cultivar Suzuko B-2961 following inoculation with clubroot pathotype 6

in a follow-up assessment under controlled conditions, although the seed company did not state

that this cultivar was clubroot-resistant.

Table 2.6. Clubroot incidence (CI) and severity (disease severity index, DSI) in susceptible and

resistant canola cultivars, grown in field soil at sites with high, lower and no measurable

concentration of resting spores at the Muck Crops Research Station, 2015.

Susceptible Resistant

Site ID Spores

(g-1 soil)

CI DSI CI DSI

2015-High 1 x 107 100 a1 70 a 6 ns2 3 ns

2015-Low 3 x 106 100 a 65 a 3 1

2015-BDL < 1000 45 b 19 b 3 1 1 Columns with the same letter do not differ at P = 0.05 based on Tukey’s multiple means

comparison test. 2 ns = not significant.

BDL = below detection limit.

 

The plant height of the resistant canola cultivars in 2014 was 30% (±1.3) shorter at the

high spore concentration site compared with the low spore concentration site at 8 weeks after

seeding (WAS) (P < 0.0001, Fig. 2.4). The biomass of resistant canola was 43% (±14.3) lower at

sites with high concentrations of resting spores (P = 0.006, Fig. 2.4). The proportion of resistant

plants at the pod development stage of plant maturity was reduced by 75% (±12.5) at the high

Page 55: Effect of Plasmodiophora brassicae ... - University of Guelph

  48  spore concentration site compared with the low spore concentration site at 8 WAS (P < 0.0001,

Fig. 2.4). There was no effect of resting spore concentration on plant height, biomass or maturity

in resistant canola in 2015 (Fig. 2.5).

Page 56: Effect of Plasmodiophora brassicae ... - University of Guelph

  49  

                       Figure  2.3 Plant height of three resistant canola cultivars at the Muck Crops Research Station,

Ontario, at sites with high, lower and undetectable resting spore concentrations, 2014 and 2015.

0  10  20  30  40  50  60  70  80  90  100  110  

4   5   6   7   8  

0  10  20  30  40  50  60  70  80  90  100  110  

4   5   6   7   8  

Plant  height  (cm

)  

Weeks  after  seeding  

0  10  20  30  40  50  60  70  80  90  100  110  

4   5   6   7   8  Weeks  after  seeding  

0  10  20  30  40  50  60  70  80  90  100  110  

4   5   6   7   8  

Plant  height  (cm

)  

Weeks  after  seeding  

0  10  20  30  40  50  60  70  80  90  100  110  

4   5   6   7   8  

Plant  height  (cm

)   45H29  73-­‐67  73-­‐77  

2014-H

2015-L 2015-BDL

2015-H

2015-H

2014-L

Page 57: Effect of Plasmodiophora brassicae ... - University of Guelph

  50  

Figure  2.4 Plant height at 8 weeks after seeding (WAS), biomass at harvest and maturity at 8

WAS for resistant canola cultivars grown in field soil with 7 x 105 spores g-1 soil (Low) and

7 x 106 spores g-1 soil (High) at Muck Crops Research Station, 2014. Bars with the same letter do

not differ at P = 0.05 based on Tukey’s multiple means comparison test.

Figure 2.5 Plant height and maturity at 8 weeks after seeding (WAS), biomass at harvest for

resistant canola cultivars grown in field soil with <1000 spores g-1 soil (BDL), 3 x 106 spores g-1

soil (Low) and 1 x 107 spores g-1 soil (High) at Muck Crops Research Station, 2015. Bars with the

same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison test.

a

b

0  

10  

20  

30  

40  

50  

60  

70  

Plan

t hei

ght (

cm)

Plant Height

Low High

a

b

0  

100  

200  

300  

400  

500  

600  

Shoot  w

eight    of  10  plants  (g)  

Biomass  

Low High

a

b

0  

10  

20  

30  

40  

50  

60  

70  

Plants  at  pod  stage  (%

)    

Maturity

Low      High  

a a a

0  

20  

40  

60  

80  

100  

120  

Plant  height  (cm

)  

Plant Height

BDL  Low    High  

a a

a

0  

100  

200  

300  

400  

500  

600  

Shoot  w

eight  10  plants  (g)  

Biomass  

BDL  Low  High  

a a a

0  10  20  30  40  50  60  70  80  90  100  

Plants  at  pod  stage  (%

)  

Maturity

BDL  Low  High  

Page 58: Effect of Plasmodiophora brassicae ... - University of Guelph

  51  Leaf length of resistant napa cabbage in was 31% (±15.5) shorter in 2014 and 22% (±1.0)

shorter in 2015 at the high spore concentration site relative to the low spore concentration site at

8 WAS (P < 0.0001). There was no reduction in leaf length between the low spore concentration

stie and site with spore concentration below detectable limits in 2015 (Fig. 2.6). Biomass of napa

cabbage was 34% (±15.9) lower at the high spore concentration site compared with the low spore

concentration site at harvest in 2014 (P = 0.007). There were no differences in biomass in 2015,

but the trend in biomass was similar to that in leaf length (Fig. 2.7).

Figure 2.6 Leaf length and biomass of napa cabbage at sites with 7 x 105 spores g-1 soil (Low)

and 7 x 106 spores g-1 soil (High) in field trials at Muck Crops Research Station, Ontario, 2014.

Bars with the same letter do not differ at P = 0.05 based on Tukey’s multiple means comparison

test.

 

a

b  

0  

5  

10  

15  

20  

Leaf  length  (cm)  

Leaf length

 

 Low      High  

a

b  

0  100  200  300  400  500  600  700  800  900  1000  

Dry  shoot  weight  (g)  

Biomass

 Low      High  

Page 59: Effect of Plasmodiophora brassicae ... - University of Guelph

  52  

Figure 2.7 Leaf length and biomass of napa cabbage at sites with <1000 spores g-1 soil (BDL), 3

x 106 spores g-1 soil (Low) and 1 x 107 spores g-1 soil (High) in field trials at the Muck Crops

Research Station, Holland Marsh, Ontario, 2015. Bars with the same letter do not differ at

P = 0.05 based on Tukey’s multiple means comparison test.

In the 2 weeks following seeding of canola in 2014, the total rainfall was 6 mm with no

more than 4 mm in any single day. Mean air temperature was 29°C maximum and 14°C

minimum. In contrast, in the 2 weeks following napa cabbage seeding in 2014, the total rainfall

was 37 mm, of which 35 mm occurred over 2 days. In 2015, both canola and napa cabbage were

seeded at the same date. In the 2 weeks following seeding, the total rainfall was 8 mm with no

more than 4 mm in any single day. Mean air temperature was 27°C maximum and 14°C

minimum.

2.3.5 pH study

At 8 days after inoculation (DAI), there were no clubbing symptoms on any plant. At

15 DAI, clubroot severity increased as the concentration of resting spores increased (P < 0.0001),

but there was no interaction or main effect of pH. At 22 DAI, there was an interaction effect

a a  

b

0  

5  

10  

15  

20  

25  

30  

Leaf  length  (cm)  

Leaf length

 

BDL      Low      High  

a a  

a

0  

200  

400  

600  

800  

1000  

1200  

1400  

Dry  shoot  weight  (g)  

Biomass

 

BDL      Low      High  

Page 60: Effect of Plasmodiophora brassicae ... - University of Guelph

  53  (P = 0.03) for CI response to spore concentration and pH, but no interaction for DSI. There was

an increase in clubroot severity with increasing spore concentration (P < 0.0001) and an effect of

pH on clubroot severity (P = 0.01). Inoculation with 1 x 103 spores mL-1 at pH 6.0-7.0 resulted in

increased clubroot incidence relative to 1 x 103 spores mL-1 at pH 5.5 and 7.5, at 22 DAI. In

contrast, there was no effect of pH on clubroot incidence at 1 x 105 and 1 x 107 spores mL-1 (Fig.

2.8). At harvest (35 DAI), increasing concentration of resting spores increased CI and DSI

(P < 0.0001). There was no interaction between pH and inoculum concentration and no effect of

pH on CI or DSI (Fig. 2.8). There was no interaction for DSI at any time-point (data not shown).

Figure 2.8 Clubroot incidence (CI) in susceptible canola grown under controlled conditions at a

range of pH (5.5 to 7.5) and concentrations of resting spores of P. brassicae at 22 days after

inoculation (DAI).

       

0  10  20  30  40  50  60  70  80  90  100  

0.0   1x10^3   1x10^5   1x10^7  

Clubroot  incidence  (%

)  

Spores  mL-­‐1  

5.5   6   6.5   7   7.5  pH

Page 61: Effect of Plasmodiophora brassicae ... - University of Guelph

  54  2.4 Discussion  

This study examined the hypothesis that there was a cost of resistance to P. brassicae.

The hypothesis was assessed in field trials, in large pots outdoors and under controlled

conditions. Overall, there was a metabolic cost of resistance at high spore concentrations in

canola and napa cabbage, but the results were inconsistent in pots, whether outdoor or under

controlled conditions. In the field trials, plant growth of canola and napa cabbage was generally

reduced and maturity was delayed at high spore concentrations.

The biomass of resistant canola was 43% lower when spore concentration was increased

by 10-fold (from 7 x 105 to 7 x 106 spores g-1) in the 2014 field trial at the Muck Crops Research

Station. A similar trend was observed for plant height, although variance was heterogeneous

across sites, indicating that there may have been an influence on plant height in addition to

P. brassicae. Crop development was also delayed, with the proportion of plants at the pod stage

reduced by 75% at 8 weeks after seeding. However, there was no effect of resting spore

concentration on plant height, biomass or maturity in 2015.

There were similar concentrations of resting spores in the sites with high spore

concentrations in 2014 (7 x 106 spores g-1) and 2015 (1 x 107 spores g-1). However, clubroot

severity was much lower in 2015 (DSI 70) relative to 2014 (DSI 100). Rainfall and air

temperature in the 3 weeks after seeding were similar in both years. Soil moisture may have had a

greater effect than rainfall on clubroot development in muck soil (Thuma et al., 1983; Cranmer,

2015). However, soil moisture was not recorded in this trial. Another difference between years

was that damage from Swede midge (Contarinia nasturtii) was higher in 2014 than in 2015. It is

possible that the metabolic cost of resistance to clubroot is larger when other biotic stresses are

present.

Page 62: Effect of Plasmodiophora brassicae ... - University of Guelph

  55  Plant height of canola cv. 45H29 (resistant) decreased by 12-14% at very high resting

spore concentrations (1 x 107 and 1 x 108), but not at 1 x 106 spores ml-1 under controlled

conditions. Similarly, biomass declined and maturity was delayed at 6 WAI in one of two

repetitions. However, there was no effect of resting spore concentration on plant height, biomass

or maturity in outdoor pot studies in 2014 or 2015. Clubroot incidence and severity were low in

the susceptible control in 2014. This was likely related to late transplanting of canola, so the

mean nighttime temperatures (12°C) were not warm enough for the development of clubroot

symptoms (McDonald and Westerveld, 2008; Gossen et al., 2012; Cranmer, 2015). In 2015,

canola seedlings were transplanted earlier in the season and were slightly younger when

transplanted (2- rather than 3-wks-old). Warmer nighttime air temperature (14-15°C) followed

inoculation, and clubroot incidence and severity were high. Previous research has demonstrated

that cool temperatures can slow pathogen colonization and reduce clubroot symptom

development (Gossen et al., 2012). Also, inoculation of older seedlings can result in slightly

lower levels of infection (Hwang et al., 2011b). However, low temperature was likely the most

important factor in the low levels of clubroot in 2014.

In general, reductions in plant height of resistant canola were apparent at 5 to 6 weeks

after inoculation (WAI) in indoor studies and 8 weeks after seeding (WAS) in field trials. In a

previous study under ideal conditions, restriction of pathogen development occurred from 2-4

weeks after inoculation (Deora et al., 2013). In the current study, that time-point corresponded

with initiation of flowering. From flowering through pod development, canola is highly sensitive

to abiotic stresses such as drought (Champolivier and Merrien, 1996). Abiotic and biotic stresses

may activate similar pathways for plant response. For example, it has been proposed that

synthesis of jasmonic acid (JA) from linolenic acid in a lipoxygenase-dependent process can be

initiated by abiotic and biotic stresses in soybean (Glycine max (L.) Merr.) (Creelman and Mullet,

Page 63: Effect of Plasmodiophora brassicae ... - University of Guelph

  56  1995). Thus, information about the timing of drought response could be helpful for understanding

response to P. brassicae during the flowering stage of development in canola.

More than one crop species was included in the current study to examine the effect of

resting spore concentration on plant growth of different Brassica species. Growth of napa

cabbage declined in sites with a high spore concentration in field trials and under controlled

conditions. In field conditions with a high concentration of resting spores, leaf length of resistant

napa cabbage grown was 31% shorter in 2014 and 22% shorter in 2015 relative to a site with a

lower spore concentration. This was consistent with one of two studies under controlled

conditions, where leaf length was 11% shorter in resistant napa cabbage inoculated with 1 x 106

spores mL-1. Growth of canola declined in sites with a high spore concentration in field trials.

However, under controlled conditions, canola did not exhibit reduction in plant height or biomass

at 1 x 106 spores mL-1 and cabbage did not exhibit reduction in leaf length or biomass (data not

shown).

Brassica vegetables have a variety of different sources of, and mechanisms of, resistance

to clubroot (Rahman et al., 2014; Chu et al., 2015; Gludovacz et al., 2015). It is possible that

there are differences in the metabolic cost of resistance to P. brassicae among species, which

may contribute to more consistent observation of cost of resistance in napa cabbage compared

with canola. It is also possible that leaf length may be more strongly correlated with cost of

resistance than plant height, because leaf area affects photosynthetic capacity, whereas plant

height does not (Freyman et al., 1973; Krogman and Hobbs, 1975; Campbell and Kondra, 1978).

Seed yield of canola was not assessed in the current study because canola was seeded late in the

season to optimize environmental conditions for pathogen development, such as warmer air and

soil temperature (Kasinathan et al., 2011a,b; Gossen et al., 2013; Cranmer, 2015). In addition,

Page 64: Effect of Plasmodiophora brassicae ... - University of Guelph

  57  there were other challenges to yield assessment at the Muck Crops Research Station, incuding

small plot size, damage from insect pests (Swede midge) and birds feeding on canola seed.

Field trials were arranged to allow for comparisons between resistant cultivars and the

susceptible line within one site. In 2015, one site with a spore concentration below the detection

limit was added as a clubroot-free control. Even though low levels of clubroot developed at that

site, it had a much lower (>1000-fold) concentration of resting spores relative to other sites.

The studies involving inoculated field soil generally did not exhibit a cost of resistance

response. In some indoor studies, plant height was slightly shorter in inoculated plants than the

non-inoculated control, but the differences were not significant. The effect of resting spore

concentration on resistant plants may have been influenced by displacement of the field soil into

pots. In a previous study, compaction and saturation of soil in pots resulted in increased clubroot

symptoms compared with drained soil (Kasinathan, 2012). However, the effect of soil saturation

on clubroot development is complex. Results from other studies have indicated that excessive soil

moisture may suppress pathogen development (Thuma et al., 1983; Cranmer, 2015), which may

be related to over-dilution of the spore concentration in soil or downward movement of inoculum

to the bottom of the pot. Conversely, much of the inoculum may have been trapped close to the

soil surface and never reached plant roots.

In the current study, clubroot symptoms were higher in the second repetition of the indoor

study, when 25% soil-less media was incorporated into field soil and a lower pH was maintained

using acetic acid. Soil at pH 6.0-6.5, which is favourable for infection and clubroot development,

may lead to increased fitness costs in resistant cultivars compared with soil pH >7.0, at a specific

concentration of resting spores. Furthermore, plant-to-plant variability within the containers may

have reduced the ability to separate out the cost of resistance from other variance in plant growth.

One possible reason that the field trials provide a much stronger and more consistent response to

Page 65: Effect of Plasmodiophora brassicae ... - University of Guelph

  58  clubroot is that only a relatively small number of resting spores were applied to the pots in the

inoculated trials. These may be highly concentrated in small areas of the root system, and it is

possible that large parts of the root system in older plants were not subject to continued infection

(Hwang et al., 2011b). However, infection can occur on new roots that develop throughout the

season, although young plants are more susceptible to infection and the rate of infection

decreases over time (Hwang et al., 2012, Fei et al., 2016).

There were several unexpected challenges in field trials. In the 2015 field trial, the napa

cabbage cultivar Suzuko B-2961 was seeded in place of the susceptible control Mirako, which

was seeded in 2014 but had been discontinued by the seed company. However, this new cultivar

turned out to be resistant to pathotype 6. Thus, susceptible canola ACS-N39 was used as a

substitute for the susceptible control for napa cabbage in 2015.

There was an unexpected increase from no clubroot symptoms in resistant canola in 2014

to the low levels of clubroot incidence and severity observed in 2015, but not in resistant napa

cabbage. High concentrations of resting spores in soil can result in selection of more aggressive

isolates of the pathogen (Kiyosawa, 1982; LeBoldus et al., 2012). It is possible that a shift in the

pathotype is occurring at Muck Crops Research Station that affects the resistance in the canola

cultivars but not napa cabbage. Alternatively, there may be have been susceptible off-types or

susceptible volunteer canola from a previous crop.

Susceptible canola grown in 2015 with 65-70 DSI was much shorter than susceptible

canola grown in 2015 with 100 DSI. These results support a previous report that plant height of

susceptible canola decreased with increasing clubbing symptoms, from 74 to 100 DSI (Hwang et

al., 2011). However, damage from swede midge (Bardner et al., 1971) may also have contributed

to the shorter plant height of susceptible canola in 2014.

Page 66: Effect of Plasmodiophora brassicae ... - University of Guelph

  59  The metabolic basis for a cost of resistance may be explained based on previous research

investigating the stages of pathogen development during which the resistance response takes

place and the metabolic pathways involved in host resistance to clubroot. Recognition and

induction of resistance takes place during root hair infection (McDonald et al., 2014). Plant host

recognition takes place when pathogen recognition receptors (PRRs) recognize conserved

pathogen-associated molecular patterns (PAMPs). Defense mechanisms may be activated by

endogenous signals that result from effectors released by the pathogen, or by microbial activity

that trigger the appropriate responses (Pal and McSpadden-Gardener, 2006). Pathogen

development is then halted or greatly restricted during cortical infection, which occurs 22-28

days after germination of resting spores under optimal conditions (Deora et al., 2012). In the

current study, this was 1-2 weeks before reductions in plant growth were observed under

controlled conditions. Although the cost of resistance is observed later, there may be an effect on

resource allocation occurring within the plant at an earlier stage that is not observable using the

indicators of growth and development in the current study. Also, there may be new infections

occurring later in the season on new root growth.

A recent study demonstrated up-regulation of signaling and metabolism of jasmonic acid

and ethylene, callose deposition and indole-containing compounds in inoculated plants of

clubroot-resistant B. rapa carrying the Rcr1 gene (Chu et al., 2015). Salicylic acid was not

elevated, and auxin and a chitinase-like protein were down-regulated. Clubroot-resistant B. napus

canola responded to secondary infection with P. brassicae by forming a concentrated reactive

oxygen species (ROS) ring in the inner cortex (Deora et al., 2013), which may contribute to

structural reinforcement of the host cell wall. ROS generation takes place in cellular

compartments such as mitochondria or chloroplasts, and involves several enzymes (Abel and

Page 67: Effect of Plasmodiophora brassicae ... - University of Guelph

  60  Hirt, 2004). ROS mediates ABA-induced stomatal closure and can react with proteins to decrease

enzyme activity (Moller, 2001). The mitochondrial electron transport chain (ETC) is a major site

of ROS production, which under normal conditions is responsible for oxidative phosphorylation,

the metabolic pathway that converts NADH into energy that is used for plant growth (Moller,

2001). Stress-induced co-expression of specific AOX (alternative oxidase) and NDH (NADH

dehydrogenase) genes in Arabidopsis (Clifton et al., 2005) may play a role in decreasing the

respiration function and leaf growth of plants in response to abiotic stress (Sevilla et al., 2015).

However, a decrease in classical pathway activity may have a greater effect than the alternative

pathway on decreasing respiration and plant growth (Sevilla et al., 2015). The effect of ROS

production on plant respiration and growth is not well understood, but provides one possible

explanation for the reduction in plant growth that occurred in the current study in resistance to

clubroot. Investigating the effect of P. brassicae on gene expression in a model plant host such as

Arabidopsis using a mutant construct with loss of function for an identified resistance gene could

improve understanding of metabolic cost of resistance to clubroot.

The data generally support the hypothesis that increasing spore concentration would

reduce plant growth and delay development in Brassica crops. However, the differences were not

quantifiable under low pathogen pressure or conditions of high variability in plant growth for

other reasons, such as soil compaction or saturation.

In the study of the effect of spore concentration and pH on clubroot incidence and

severity, susceptible canola inoculated with 1 x 103 spores mL-1 at pH 6.0-7.0 resulted in higher

clubroot incidence compared with 1 x 103 spores mL-1 at pH 5.5 and 7.5, at 22 days after

inoculation. In contrast, there was no effect of pH on clubroot incidence at 1 x 105 and 1 x 107

spores mL-1. This interaction was observed at 22 DAI but not at any other sample dates. These

results support previous reports by Colhoun (1953) that there is a strong influence of pH on

Page 68: Effect of Plasmodiophora brassicae ... - University of Guelph

  61  clubroot severity at low concentrations of resting spores, but not at high spore concentrations.

Previous studies have reported that 1000 spores g-1 soil is the infection threshold for susceptible

canola (Donald and Porter, 2009; Faggian and Strelkov, 2009). These results support the

hypothesis that pH 5.5 and >7.0 reduced clubroot incidence at low concentrations of resting

spores but not at concentrations ≥ 1 x 105 spores mL-1, although the differences only became

apparent at 22 days after inoculation. However, these results are based on only one repetition, and

therefore should be treated as preliminary until the study can be repeated. Study repetition should

include 10 plants per experimental unit instead of three plants to provide a more robust

assessment of clubroot incidence and disease severity. Plant root tissues collected at 8, 15 and 22

DAI have been stored for future examination of pathogen colonization of root tissue.

Quantification of pathogen DNA using quantitative polymerase chain reaction (qPCR) would

provide an assessment of pathogen colonization in host plant root tissue.

Increased understanding of the effect of resting spore concentration at various pH levels

will make it easier to predict clubroot severity and yield loss at specific field sites by providing a

more precise assessment of the risk of clubroot.

In summary, field studies indicated that there is a cost of resistance in canola and napa

cabbage. Reductions in plant height of resistant canola ranged from 12-14% indoors to 30% in

one field trial. Biomass was also reduced and maturity was delayed in some but not all studies.

Napa cabbage leaf length was reduced by 22-31% in field trials and biomass was reduced by 34%

in one field trial. Growers need to be aware that growing a resistant canola cultivar in a field with

a high concentration of resting spores results in a reduction in yield and delayed maturity, even if

no clubroot symptoms develop.

Page 69: Effect of Plasmodiophora brassicae ... - University of Guelph

  62  

CHAPTER THREE

DECLINE IN RESTING SPORES AND EFFECT OF CROP ROTATION FOLLOWING A

SUSCEPTIBLE CROP

3.1 Introduction  

Inoculum concentrations of soil-borne pathogens can increase rapidly in the presence of a

susceptible host (Gilligan et al., 1996; Hwang et al., 2013). High inoculum concentrations pose

an important threat to disease management strategies that rely heavily on genetic resistance. High

inoculum concentrations in soil can accelerate the breakdown of genetic resistance via selection

of more aggressive or virulent isolates of the pathogen (Kiyosawa, 1982; LeBoldus et al., 2012).

Furthermore, metabolic costs associated with disease resistance can lead to delayed development

and reduced plant growth and yield in canola (Tian et al., 2003; Hwang et al., 2011a).

Soil-borne pathogens such as Fusarium oxysporum f. sp. lactuca (Scott et al., 2012) and

Verticillium dahliae (Vos et al., 2012) can survive in soil in the absence of a host for 15 years or

more. The longevity of a soil-borne pathogen depends on many factors, including environmental

conditions, presence or absence of a host, and the morphological and chemical characteristics of

survival structures (McKeen and Wensley, 1961; Macfarlane, 1970; Coley-Smith et al., 1990).

Persistence in soil, and subsequent disease development following a long rotation, often lead

growers to conclude that crop rotation is not an effective method for reducing pathogen

populations.

The reduction in inoculum concentration of some soil-borne plant pathogens has been

studied in detail. Inoculum of V. dahliae in soil declined by 96% from the initial concentration

Page 70: Effect of Plasmodiophora brassicae ... - University of Guelph

  63  after 7 years of rotation with non-host crops. However, the inoculum concentration that remained

after 7 years was still three times the infection threshold (Ben-Yephet and Szmulewich, 1985).

Inoculum concentration of F. oxysporum f. sp. lactuca declined by 86% after a 12-month break

following a susceptible crop. In a resistant cultivar, however, F. oxysporum colonizes the root

cortex and produces inoculum, even though disease symptoms are not present (Scott et al., 2014).

Resistant plants contribute even more inoculum to the soil than susceptible plants, due to their

larger size.

For P. brassicae, a field survey was conducted on 190 field sites that were not uniform for

soil composition, structure, tillage practices, weed management, or cropping system during the

period of break from canola (Wallenhammar, 1996). Also, Brassica weeds were present in all of

the fields, some fields had adjacent areas highly infested with P. brassicae, and treatments were

not replicated to account for the variation in the initial concentration of resting spores. It took

about 17.3 years for the concentration of resting spores in a highly infested soil to decline below

the detection limit in Sweden. This study was conducted to a high standard using the technologies

available, but factors such as potential inflow of inoculum and increase on weeds complicate

interpretation of the data.

In Canada, resting spore concentration of P. brassicae declined by 98% from the initial

spore concentration of 3 x 106 spores g-1 soil following a 2-year break from susceptible canola, in

a recent study at the AAFC Research Farm at Normandin, Quebec (Peng et al., 2015). Thus, the

half-life of individual resting spores may be much shorter than the half-life of the infective

capacity of the soil (Wallenhammar, 1996; Hwang et al., 2013). Despite the decline in a large

number of resting spores, the resulting spore concentration (6 x 104 spores g-1 soil) after a 2-year

break from susceptible canola remained above the infection threshold for susceptible cultivars

Page 71: Effect of Plasmodiophora brassicae ... - University of Guelph

  64  (Peng et al., 2015), which is estimated to be 1000 spores g-1 soil (Donald and Porter, 2009;

Faggian and Strelkov, 2009).

Resting spores of P. brassicae are well-adapted to harsh environmental conditions

(Mafarlane, 1970). Spines covering the surface of the spore are enmeshed in fibres that are

approximately 5 nm in diameter (Buczacki et al., 1979). The protein that makes up the outermost

layers of the resting spore wall protects the inner chitin layer from natural chitinolytic processes

in the soil (Buczacki and Moxham, 1983). Environmental factors that affect germination include

pH, soil moisture, soil temperature, and the presence of calcium ions (Kageyama and Asano,

2009). Germination can take place with or without a host plant present (Ingram and Tommerup,

1972; Friberg, 2005), but the germination rate increases with the age and the level of decay of

spore-containing clubs (Macfarlane, 1970). The likelihood of germination also increases in the

presence of a compatible host (Macfarlane, 1970), such as susceptible canola volunteers. These

can persist in fields for many years and dramatically reduce the effectiveness of crop rotation

(Harker et al., 2014). Also, many of the common weeds in western Canada are susceptible to

clubroot, such as wild mustard (B. kaber L.), white mustard (B. hirta L.) shepherd’s purse

(Capsella bura-pastoris L.) and stinkweed (Thlaspi arvense L.).

From 1980 to 2000, more diverse crop rotations and reduced tillage were introduced in

western Canada as a strategy to reduce pest problems and decrease production risks. However,

from 2000 to the present, shorter rotations have become increasingly common, such as canola-on-

canola (no break from canola) and cereal-and-canola (1-year break from canola) (Kutcher et al.,

2013). Despite recommendations of canola 1 in 4 years (Rimmer et al., 2003) or canola 1 in 3

years (Cathcart et al., 2006), a shorter rotation is popular because of the high commodity price

and economic returns of canola relative to cereals and pulses (Strelkov et al., 2011). The

beneficial effect of crop rotation is associated with decreased allelopathy from plant residues

Page 72: Effect of Plasmodiophora brassicae ... - University of Guelph

  65  from previous canola crops, decreased presence of other canola pathogens, decreased weed

pressure, and improved soil structure and nutrition due to diverse rooting depths and nutrient

uptake of non-Brassica crops (Harker et al., 2014).

The effect of cropping rotation, especially of resistant canola cultivars, on resting spore

concentration in soil is not well understood. Resistant canola contributes fewer spores to the soil

than susceptible canola (Wallenhammar et al., 2000). Two studies of resistant cultivars of

Brassica vegetables observed a reduction in the concentration of resting spores of P. brassicae

relative to a non-host crop or fallow treatment (Yamagishi et al., 1986; Murakami et al., 2001).

In a more recent study, conducted in tubs placed outdoors, a resistant canola cultivar had either

no effect or produced a small increase in spore concentration in soil compared to an unplanted

control (Hwang et al., 2013). These studies used counts of spores extracted from soil, but recent

studies indicate that such estimates can be highly variable (Cranmer, 2015; F. Al-Daoud, personal

communication). Yield of resistant canola following 2- to 4-year break from susceptible canola

can be up to 25% higher than yield of canola grown in shorter rotations (Peng et al., 2015).

Accurate quantification of the concentration of P. brassicae resting spores in soil after

various lengths of break following susceptible and resistant canola can improve our

understanding of the long-term effect of crop rotation on canola production and have important

implications for clubroot management. Real-time quantitative polymerase chain reaction (qPCR)

assessments are effective for quantification of pathogen spore levels in the soil (Faggian and

Strelkov, 2009). However, physical or chemical substances that occur naturally in the soil, such

as humic acids, phenolic compounds, clay particles, or heavy metals, can inhibit amplification of

the target DNA during the qPCR process (Volossiouk et al. 1995; Matheson et al. 2010). Sample

dilution is a common approach to dealing with inhibition during qPCR amplification (Hoshino

and Inagaki, 2012). This is important because levels of inhibitors in soil cannot be assumed to be

Page 73: Effect of Plasmodiophora brassicae ... - University of Guelph

  66  consistent across treatments of crop rotation, which may have varying degrees of organic matter

breakdown. The use of an internal control allows for estimation of the level of inhibition that is

present in each sample, and so more accurate estimates of spore concentration (Deora et al.,

2015).

The first objective of this research was to examine the effect of crop rotation on the

decline of resting spores over time and the concentration of resting spores in soil. It was

hypothesized that the concentration of resting spores in soil would decline exponetially with

increasing length of break interval following a susceptible canola crop. The second objective was

to examine the interaction between crop rotation and inoculation with P. brassicae on growth of

clubroot-resistant canola. The main effect of crop rotation was expected to increase plant height

as length of break from canola increases. The main effect of inoculation with P. brassicae was

expected to decrease plant height compared with a non-inoculated control. It was hypothesized

that the interaction of crop rotation and inoculation with P. brassicae would affect the metabolic

cost of resistance to clubroot and subsequent reductions in plant height of resistant canola. It was

therefore expected that plant height would be shortest in the treatment of soil with no crop

rotation and inoculation with P. brassicae.

Page 74: Effect of Plasmodiophora brassicae ... - University of Guelph

  67  

3.2 Materials and Methods

3.2.1 Decline in resting spores over time following susceptible canola

Soil samples were collected in the spring of 2014 from the Agriculture and Agri-Food Canada

(AAFC) research farm at Normandin, Québec (latitude 48°51’N, longitude 72°32’W). The soil is

a Labarre silt loam (Humic Gleysol) that consists of 8% sand, 70% silt and 22% clay and

approximately 4% organic matter (assessed separately). Soil composition and texture were

determined at the Agrifood Laboratories, Guelph, ON. The soil is naturally infested with

P. brassicae, identified as pathotype 2 (Strelkov et al., 2006) according to the Williams systems

of classifications (Williams, 1966). Small blocks (8 x 30 m) were selected based on cropping

history as part of a larger, long-term crop rotation study, where the rotation crops were canola,

barley, field pea and fallow. In fallow treatments weeds were controlled once a year (by

harrowing). However, fallow treatments were not weed-free throughout the entire season and

weeds may have included susceptible host species. Break intervals following susceptible canola

were 0, 1, 2, 3, 5, and 6 years (Table 3.1).

Resting spore concentration in soil was quantified after break intervals of 0, 1, 2, 3, 5, and

6 years following a susceptible canola crop. For each length of break, two plots were selected and

each plot was divided into two parts for subsampling, by collecting five soil cores (surface to 15-

cm depth) in a W pattern, with a core from each point in the W. The samples were dried at room

temperature and bulked. Each subsample was pulverized using a ceramic pestle and mortar

(CoorsTek Inc., Golden, CO), which were cleaned and autoclaved prior to each use to avoid

contamination. DNA was extracted using the PowerSoil DNA isolation kit (MO BIO

Laboratories Inc., Carlsbad, CA). Three biological replicates were conducted for each soil bulk,

with three technical replicates per biological replicate. For each biological replicate, 250 mg

Page 75: Effect of Plasmodiophora brassicae ... - University of Guelph

  68  

Table 3.1 Cropping rotation following susceptible canola in selected plots sampled from a long-term rotation study at AAFC research farm in Normandin, Quebec.

Years break

from canola Block Cropping history

0 1

2

n/a

n/a

1 1

2

Fallow (F)

F

2 1

2

F-F

F-F

3 1

2

F-F-F

F-F-F

5 1

2

Pea-Barley-F-F-F

Barley-Pea-F-F-F

6 1

2

Barley-Barley-Pea-F-F-F

Barley-Barley-Pea-F-F-F

of soil was added directly into microbead tubes provided with the kit. DNA extraction and

purification were performed according to the manufacturer’s instructions. The spore

concentration in each sample was determined using a multi-plexed TaqMan qPCR assay, with a

competitive internal positive control (CIPC) to minimize differential inhibition among the

samples (Deora et al., 2015). Amplification of a known quantity of the CIPC is used to estimate

the level of PCR inhibition during the DNA amplification process (Wang et al., 2007). The CIPC

selected for these assessments was the DNA of Green Fluorescent Protein (GFP) from Aequorea

victoria, a jellyfish (Deora et al., 2014). Forward (DC1F) and reverse (DC1mR) primers

amplified a 90-bp fragment of the internal transcribed spacer (ITS1) region of P. brassicae. A

Page 76: Effect of Plasmodiophora brassicae ... - University of Guelph

  69  TaqMan probe (PB1) with a 5’ end reporter dye of FAM and 3’ quencher of NFQ-MGB was

used. The CIPC probe, GFP1, was derived from a plasmid (pDSK-GFPuv1) that contains a

variant of gene coding for GPF (GFPuv1) (Wang et al., 2007). Both primers and probe were

designed for conventional and quantitative PCR amplification of P. brassicae, using a partial

P. brassicae 18S ribosomal RNA (rRNA) gene sequence (Hwang et al., 2011b).

Assays were conducted in a 96-well Step-One Real-Time PCR system (Applied

Biosystems, CA), according to the manufacturer’s instructions. Each sample was homogenized

using a mechanical bead rupter (OMNI International, Inc., Kennesaw, GA) for 1 min. at high

speed. A 10x dilution with molecular grade water (Hyclone Laboratories Inc., Logan, UT) was

made for each genomic DNA (gDNA) sample. A standard series was produced using gDNA

derived from 1 x 108 P. brassicae resting spores mL-1 diluted in 10-fold increments down to 102

resting spores mL-1 (Deora et al., 2015). The standard series was used to generate a standard

curve for qPCR assays (as described by Deora et al., 2015, and previously conducted for

Entrococcus bacteria by Lievens and Thomma, 2005). Two technical replicates were included in

each concentration of the standard series. A negative control sample (Hypure™Molecular

Biology Grade Water, Hyclone Laboratories Inc.) with two technical replicates was also included

in each assay. An additional standard series without GFP was included as a negative control for

GFP in each assay. Each 20 µl reaction consisted of 10 µl Taqman® Universal PCR Master Mix

(P/N 4304437, Applied Biosystems), 1.8 µl of 250 nM of P. brassicae primer pairs, 0.5 µl of 250

nM of P. brassicae probe, 0.5 µl of 250 nM CIPC probe (GFP1), 1.4 µl of molecular grade water,

2 µl internal control and 2 µl of target DNA. Thermal cycling conditions were: 50°C for 2 min.,

95°C for 10 min., 40 cycles of 95°C for 15 sec., followed by 62°C for 1 min. Estimates of spore

concentrations were generated as the output for Cq values. The detection limit was Cq > 35 and

Page 77: Effect of Plasmodiophora brassicae ... - University of Guelph

  70  the lowest value on the standard curve was 103 spores mL-1 (Cq = 35.3). Therefore, a Cq value of

35.1 was selected as a cut-off point, and values lower than the cut-off were considered to be not

accurately quantifiable data.

Historical weather data for the Normandin Research Farm was accessed from

Environment Canada at http://climate.weather.gc.ca. Mean monthly air temperature was

calculated from historical weather data over a 10-year period, as well as total monthly rainfall

during the growing period of canola at Normandin, Quebec, from 2007–2014.

3.2.2 Effect of resistant canola on the concentration of resting spores in soil

In 2014, soil samples were collected from a follow-up study for the study described in

Section 3.2.1 at AAFC Normandin Research Farm. The clubroot-resistant canola cultivar 45H29

(Pioneer Hi-Bred Ltd, Chatham, ON) was seeded in 20 plots in 2014. Each plot is one

experimental unit. There were three treatments. One treatment was seeded with canola cv. 45H29

where susceptible canola had been cultivated the previous year, with eight field replicates.

Another treatment was seeded with canola cv. 45H29 where susceptible canola had been grown

in 2012, followed by a 1-year break (fallow), also with eight field replicates. Another treatment

was seeded canola cv. 45H29 where susceptible canola had been grown in 2011, followed by a 2-

year break (fallow), with four field replicates. Each plot was sampled as described above in early

spring of 2014 (before seeding), again in fall 2014 (after harvest), and in spring 2015

(approximately 8 months after harvest). Resting spore concentration was estimated using the

extraction and qPCR protocol described above.

Page 78: Effect of Plasmodiophora brassicae ... - University of Guelph

  71  3.3.3 Controlled environment study – crop rotation and spore concentration

Experiments examining the relationship between the presence of P. brassicae resting

spores, soil type and crop rotation were conducted in a controlled environment study using field

soil with no history of clubroot. The study was conducted as a 2×2×2 factorial arranged as a

randomized complete block design with four replicates. The factors were: crop rotation (0- vs. 2-

year break from canola), inoculation (P. brassicae spore suspension vs. water-only control), and

soil type, represented by two locations (Scott, SK and Elora, ON) (Table 3.2). A Dark Brown

Chernozemic mixed Elstow and shallow Elstow loam soil (Brandt and Zentner, 1995) was

collected in June of 2014 from two treatments in a rotation study at the Agriculture and Agri-

Food Canada Research Farm near Scott, SK. A Guelph series loam (Orthic Gray Brown Luvisol)

(Knezevic et al., 1994) was collected in June of 2014 from two adjacent sites at Elora Research

Station of the University of Guelph near Elora, ON. At each location, one sampling site had no

break from canola (planted to canola in 2013) and the other site had a 2-year break from canola

(planted to canola in 2011). During the 2-year break from canola, the site at Scott was cropped

with wheat the first year and field pea the second year. The site at Elora was cropped with

soybean followed by maize.

Page 79: Effect of Plasmodiophora brassicae ... - University of Guelph

  72  Table 3.2 Comparison of soil locations used in crop rotation and inoculum trials, under

controlled environment in Guelph, Ontario, 2014.

Location Latitude,

Longitude

Years break

from canola

pH Nutrient analysis (ppm) OM1

(%) P K Mg Ca

Elora,

ON

43°64' N,

80°40' W

0 7.2 37 135 346 3186 4.0

2 7.3 23 147 426 3038 3.9

Scott,

SK

52°36' N,

108°84' W

0 5.1 46 391 265 1145 3.5

2 6.5 31 338 344 2151 4.5 1 OM = Organic matter content (%)

The clubroot-resistant canola cultivar 45H29 was sown in tall plastic ‘conetainer’ pots

(Stuewe Sons Inc., Corvallis, OR) filled with soil. Two seeds were planted per pot and thinned to

one seedling per pot. There were 10 plants per experimental unit and four replicates. The

clubroot-susceptible canola line ACS-N39 (AAFC, Saskatoon Research Centre, Saskatoon, SK)

was included as a susceptible control for each treatment. The plants were maintained at 25°/20°C

day/night, with 16-hr photoperiod and 65% relative humidity, and fertilized with N-P-K solution

(20-20-20) at 2–3 day intervals. The plants were watered weekly with deionized water adjusted to

pH 6.0 using commercial white vinegar to maintain a pH level between 6.0 and 6.5. Each 10-day-

old seedling was inoculated with 5 mL of resting spore suspension of 1 x 106 resting spores mL-1

of P. brassicae pathotype 6. Inoculum was prepared following the protocol outlined by Sharma et

al. (2011). Briefly, mature clubs were washed and soaked in deionoized water; 10 g of clubbed

root was homogenized in 300 mL deionized water for 2 min. in an electronic grinder. The

mixture was filtered through eight layers of cheesecloth and resting spore concentration was

estimated using a haemocytometer. The spore suspension was diluted to 1 x 106 resting spores

mL-1.

Page 80: Effect of Plasmodiophora brassicae ... - University of Guelph

  73  Plant height, measured from the soil surface to the shoot apex, was assessed for each plant

in each experimental unit at 7-day intervals from 2−6 weeks after inoculation (WAI). At 10

weeks after planting, plants were harvested and weighed, and roots were assessed for clubroot

incidence (%) and severity using the standard 0-3 rating scale, where 0 = no clubbing, 1 < 1/3 of

root clubbed, 2 = 1/3–2/3 of roots clubbed and 3 > 2/3 of roots clubbed (Strelkov et al., 2006). A

disease severity index (DSI) was calculated according to Crete et al. (1963):

DSI = ∑ [(c [class no.)(no. of plants in each class)]

x100 (total no. plants per sample)(no. classes - 1)

3.3.4 Statistical analysis

All of the statistical analyses were performed with SAS software (version 9.3 SAS

Institute, Cary, NC) with a type I error set at P = 0.05. Data were tested for normality using the

Shapiro-Wilk test of residuals and tested for outliers using Lund’s test. Regression was tested

using PROC REG using data-points from two parts per block and two blocks per treatment.

The concentration of resting spores was not normally distributed, based on the Shapiro-

Wilk statistic (W = 0.431, P < 0.0001). Data were log transformed to reduce the heterogeneity of

the variance. Tests of normality on log-transformed data confirmed that distribution was normal

(W = 0.946, P = 0.22). Lund’s test of outliers identified only one extreme observation in the

treatment for 0-year break from canola. Variability in the concentration of resting spores

increased in treatments with higher concentration of resting spores in soil, in particular for the 0-

break from canola treatment. The data point was retained in the analysis, even though the

Studentized residual was 2.90, which is higher than the critical value for n=24, q=1 (Lund, 1975).

The concentration of resting spores in the 2014 resistant canola trial was also not

normally distributed, based on the Shapiro-Wilk statistic (W = 0.904, P = 0.008). Data were not

Page 81: Effect of Plasmodiophora brassicae ... - University of Guelph

  74  normally distributed when log-transformed (W = 0.943, P = 0.001). However, parametric

analyses were continued to remain consistent with analysis of all other data in the study.

Height measurements were collected for comparison at individual time-points and also to

examine overall growth over time. Growth over time, using plant height assessments over time as

a measurement for growth, was summarized using ‘area under the growth stairs’ (AUGS) as

previously described (Richards, 1959; Simko and Piepho, 2012). The area under the curve is

divided into a series of trapezoids using a simple calculation called the trapezoidal rule (Simko

and Piepho, 2012).

3.3   Results  

3.3.1 Decline in resting spores over time following susceptible canola

Resting spores of P. brassicae were present in all plots and treatments sampled. The

concentration of resting spores following a susceptible canola cultivar declined over time in a

quadratic relationship, y = (1E + 07)e-0.759x, R2 = 0.65, as the length of break from the susceptible

canola crop increased (Fig. 3.1). Regressional analysis resulted in a linear relationship on log-

transformed data (P = 0.008). Resting spore concentration declined by 96.5% after a 1-yr break,

and 99.3% after a 2-yr break from canola, but then declined slowly compared to continuous

canola (1.3 x 108 spores g-1 soil).

Page 82: Effect of Plasmodiophora brassicae ... - University of Guelph

  75  

Figure 3.1 Decline in resting spore concentrations of Plasmodiophora brassicae over time

following susceptible canola at Normandin, Quebec, assessed in 2014.

3.3.2 Effect of resistant canola on the concentration of resting spores in soil

Following continuous susceptible canola, one crop of resistant canola increased the

concentration of resting spores in soil by 12-fold, from 2 x 108 to 3 x 109 spores g-1 soil (Table

3.3). However, an unexpectedly high level of clubroot symptoms developed on canola cultivar

45H29, which has previously been resistant to clubroot at the Normandin site (D. Pageau,

personal communication). Following a 1-year break from susceptible canola, one crop of resistant

canola had no significant effect on the concentration of resting spores in soil. Following a 2-year

break from susceptible canola, one crop of resistant canola increased the concentration of resting

spores in soil by nearly 8-fold from 2 x 106 to 1 x 107 spores g-1 soil (Table 3.4).

Large numbers of resting spores were lost between the fall of 2014 and the spring of

2015. The concentration of resting spores was reduced by 97% (±0.3) from 3 x 109 spores g-1 soil

in the fall of 2014 to 8 x 107 in the spring of 2015, when one crop of resistant canola followed

y  =  1E+07e-­‐0.759x  R²  =  0.65,  P  =  0.008  

1.0E+04  

1.0E+05  

1.0E+06  

1.0E+07  

1.0E+08  

0   1   2   3   4   5   6  

Log  scale  

Resting  spores  g-­‐1  soil  

Years  break  from  canola  

Decline in Resting Spores

Page 83: Effect of Plasmodiophora brassicae ... - University of Guelph

  76  continuous susceptible canola (Table 3.4). Similarly, resting spores were reduced by 99% (±0.2)

when resistant canola followed a 1-year break from susceptible canola and 90% (±1.0) when

resistant canola followed a 2-year break from susceptible canola (Table 3.4).

Prior to planting with canola cv. 45H29, the concentration of resting spores in soil

declined by 92% after a 1-year break and 99% after a 2-year break from canola, compared to no

break from susceptible canola (2 x 108 spores g-1 soil, Table 3.4). The concentration of resting

spores declined by >99% in treatments Fallow(F)-Resistant(R) and F-F-R, compared with

Susceptible(S)-R (3 x 109 spores g-1 soil) in soil samples collected in the fall of 2014. The

concentration of resting spores in soil declined by >98% in treatments F-R and F-F-R, compared

with S-R (8 x 107 spores g-1 soil) in soil samples collected in the spring of 2014.

After one crop of canola cv. 45H29, there was no difference between the F-R treatment

(1-yr break between susceptible and resistant canola) and the F-F-R treatment (2-yr break

between susceptible and resistant canola). Clubroot symptoms were observed in these fields,

indicating that spore concentration may have been influenced by susceptible weeds or off-types,

or a pathotype of P. brassicae that has overcome resistance of the host cultivar.

Page 84: Effect of Plasmodiophora brassicae ... - University of Guelph

  77  Table  3.3 Quantification of resting spore concentration using qPCR after crop rotation

treatments including fallow (F), resistant (R) and susceptible canola (S), sampled in spring and

fall of 2014 and spring of 2015 at Normandin, Quebec.

Treatment Inital1

(g-1 soil) SE2 Final3

(g-1 soil) SE

Spring 2014 S 5 x 106 2.0 x 106 2 x 108 1.8 x 108

F 1 x 107 5.7 x 106 2 x 107 3.6 x 106

F-F 8 x 105 1.8 x 105 2 x 106 4.3 x 105

Fall 2014 S-R 3 x 107 6.8 x 106 3 x 109 2.3 x 109

F-R 1 x 107 5.3 x 106 1 x 107 3.3 x 106

F-F-R 1 x 107 2.9 x 106 1 x 107 2.9 x 106

Spring 2015 S-R 2 x 107 1.2 x 107 8 x 107 5.6 x 107

F-R 5 x 104 3.7 x 104 2 x 105 1.2 x 105

F-F-R 5 x 104 2.3 x 104 1 x 106 7.8 x 105

1Initial= Initial spore concentration prior to adjustment of estimates for inhibition based on

amplification of the internal control. 2SE = Standard error. 3Final = Final spore concentrations (g-1 soil) following adjustment for the internal control

are indicated in bold.

3.3.3 Weather

At the site of the crop rotation study at Normandin, soil moisture and temperature were

generally conducive for clubroot development during the first 2 to 3 weeks after seeding (Table

A2.7). During the course of the portion of the long-term study examined in the current report

(2007-2014), weather conditions were generally near normal. Rainfall was lower than the 10-year

average by 12-36 mm from May to July 2007, higher by 17-57 mm from June to July 2008, lower

Page 85: Effect of Plasmodiophora brassicae ... - University of Guelph

  78  by 46 mm in July 2010, higher by 41 mm in May 2011, higher by 17-33 mm in May and June

2012, and higher by 16-39 mm in May and June 2013 (Table A2.7). Throughout the period of the

study, mean monthly air temperature varied up to 2°C cooler and 2°C warmer than the 10-year

average during the growing season (May to September, Fig. 3.2). Mean air temperature in May

was lower than 9°C in 2008 and 2009. Mean air temperature in May was 10°C in 2011 and 11°C

in 2010, 2012 and 2013.

Figure 3.2 Monthly rainfall (bars) and mean monthly air temperature (line) at Normandin,

Quebec, for May to September, 2007-2013.

3.3.4 Controlled environment – crop rotation and spore concentration

Clubroot incidence and severity in the susceptible control were higher in soil from Scott

than in soil from Elora (Table 3.5), likely associated with higher pH in the Elora soil.

0  2  4  6  8  10  12  14  16  18  20  

0  

50  

100  

150  

200  

250  

M   J   J   A   S  M   J   J   A   S  M   J   J   A   S  M   J   J   A   S  M   J   J   A   S  M   J   J   A   S  M   J   J   A   S   Monthly  air  temperature  (m

ean  °C)  

Monthly  rainfall  (mm)  

2007 2008 2009 2010 2011 2012 2013

Page 86: Effect of Plasmodiophora brassicae ... - University of Guelph

  79  Table  3.4 Clubroot incidence (CI) and severity (disease severity index, DSI) in canola breeding

line ACS-N39 (susceptible control) and the resistant cultivar 45H29, inoculated with P. brassicae

and grown under controlled conditions.

ACS-N39 45H29

Assessment CI DSI CI DSI

Scott soil, inoculated 90 a1 66 a 0 ns2 0 ns

Scott soil, not inoculated 0 c 0 c 0 0

Elora soil, inoculated 65 b 30 b 0 0

Elora soil, not inoculated 0 c 0 c 0 0 1 Means in a column followed by the same letter do not differ, based on Tukey’s multiple means

comparison test at P = 0.05. 2 ns = not significant.

Inoculation of the clubroot-resistant canola cultivar with resting spores of P. brassicae

reduced plant height by 7% (±3.3) at 4 WAI (P = 0.047), and 5 WAI (P = 0.038). Data from the

two repetitions of these dates were pooled because there were no significant repetition effects in

plant height (Tables A2.8-A2.9). There was a significant soil location effect at four out of five

time-points, as expected based on the broad differences in soil type, texture, pH and cropping

history.

Area Under the Growth Stairs (AUGS), calculated from plant height measured each week

from 2-6 WAI, indicated that overall growth of clubroot-resistant canola cultivars inoculated with

P. brassicae was reduced by 8% (±4.8) (P = 0.04, Table A2.11) across repetitions in soil from

both sites. In contrast, there was no effect of cropping rotation or interaction between inoculation

and crop rotation at either site (Fig. 3.3).

Page 87: Effect of Plasmodiophora brassicae ... - University of Guelph

  80  

Figure 3.3 Effect of crop rotation and inoculation with Plasmodiophora brassicae on plant

height (area under growth stairs, AUGS) of clubroot-resistant canola, under controlled conditions

in field soil from Elora, ON, and Scott, SK in 2014.

 The results for biomass (dry shoot weight) of resistant canola cultivars were more complex. The

biomass of resistant plants inoculated with P. brassicae was consistently lower than the non-

inoculated control (P = 0.004). For crop rotation, each site and repetition was analyzed separately

because there was a significant interaction between site, rotation and repetition (P = 0.001). In

Scott soil, a 2-year break from canola increased plant biomass by 41% compared with no break in

the first repetition (P = 0.002), but there was no effect in the second repetition (Table 3.6). The

trends were similar in repetition 2, but there was more variability in the treatment with 2-year

break from canola and therefore increased standard error. The opposite trend was observed in

Elora soil, where the 2-year break reduced biomass by 25% compared with no break (P = 0.003,

Table 3.6).

A A A A

0  

30  

60  

90  

120  

150  

180  

Elora, ON

Plan

t hei

ght (

AU

GS)

2-yr Break- Inoculated 2-yr Break- Water Control No Break- Inoculated No Break- Water Control

AB A

B

AB

0  

30  

60  

90  

120  

150  

180  

Scott, SK Pl

ant h

eigh

t (A

UG

S)

2-yr Break- Inoculated 2-yr Break- Water Control No Break- Inoculated No Break- Water Control

Page 88: Effect of Plasmodiophora brassicae ... - University of Guelph

  81  Table 3.5. Effect of crop rotation and inoculation with P. brassicae on biomass (dry shoot

weight, g) of clubroot-resistant canola, under controlled conditions in field soil from Elora, ON,

and Scott, SK, 2014.

Elora, ON Scott, SK

Repetition 1 Repetition 2 Repetition 1 Repetition 2

Spores mL-1

Biomass SE1 Biomass SE Biomass SE Biomass SE

No break

1x106 2.7 ab2 0.09 4.3 ns3 0.34 2.4 b 0.23 2.3 ns3 0.12

0 2.9 a 0.33 5.1 0.70 2.7 b 0.34 2.7 0.15

2-year break

1x106 2.0 b 0.24 3.0 0.50 3.3 ab 0.23 2.1 0.26

0 2.5 ab 0.26 3.7 0.64 4.0 a 0.27 2.8 0.70 1Means in a column followed by the same letter do not differ based on Tukey’s multiple means

comparison test at P = 0.05. 2ns = not significant. 3SE = standard error.

3.4 Discussion  

In the current study, the concentration of resting spores in soil following a susceptible

canola cultivar declined over the first 2 years following a susceptible canola crop. These results

support a recent study that observed a decline in resting spores at the AAFC Research Farm at

Normandin, QC, conducted at the same site as the current study but in two different years (Peng

et al., 2015). Similarly, the concentration of resting spores in soil declined in mini-plots

consisting of plastic tubs containing field soil placed outdoors near Edmonton, AB, where both

continuous resistant canola and continuous fallow reduced the concentration in resting spores in

soil relative to continuous susceptible canola (Hwang et al., 2013).

The current study indicated that the concentration of resting spores in soil declined by

97% after a 1-year break and 99% after a 2-year break from susceptible canola, but then declined

Page 89: Effect of Plasmodiophora brassicae ... - University of Guelph

  82  very slowly (<1%) over the next four years. Following a 6-year break from canola, the

concentration of resting spores was 4.1 x 105 spores g-1 soil. This is more than 400-fold greater

than the estimated threshold for uniform infection (1000 spores g-1 soil) in a susceptible cultivar

(Donald and Porter, 2009). Therefore, a viable management strategy for a location with a high

concentration of P. brassicae resting spores should include crop rotation in combination with a

clubroot-resistant cultivar. A management strategy that combines a crop rotation with ≥ 2-year

break from canola and use of a clubroot-resistant cultivar can reduce the concentration of resting

spores and fitness cost of resistance to clubroot. This would also contribute to a reduction in

selection pressure within the pathogen population, which could delay the erosion of resistance

(Kiyosawa, 1982; LeBoldus et al., 2012).

Estimates of the half-life of the infective capacity of resting spores have ranged from 3.6

years (Wallenhammar, 1996) to 4.4 years (Hwang et al., 2013). However, a decrease in ≥ 98% of

resting spores in the first 2 years following susceptible canola was observed in the current study

and in a recent study at the same site in two different years (Peng et al., 2015). It is possible that

within one population of resting spores, there may be differences in the half-life of individual

resting spores. These differences may be related to decreased microbial degradation of resting

spores that have increased protein structure in the outer spore wall (Moxham et al., 1983). Thus,

although ≥ 98% of resting spores die or disappear within the first 2 years, the half-life of the

remaining ≤ 2% of spores may be closer to 5 years or longer. Similarly, there could be large

differences in the pattern of decline in resting spores at different sites. For example, in a site

where clubroot has been present for decades and many susceptible hosts have been grown, it is

possible that the spore population in the soil is more adapted for protection and would have a

longer half-life of infective capacity than in an inoculated soil or a soil with a relatively new

Page 90: Effect of Plasmodiophora brassicae ... - University of Guelph

  83  infestation. However, it is also possible that there could be an increase over time in the

concentration of microbes that degrade P. brassicae resting spores, which would lead to a shorter

half-life of resting spores at sites where clubroot has been established for a longer period of time.

Clubbing symptoms were observed in a large number of canola plants in the field at

Normandin, Quebec, seeded with cultivar 45H29. This cultivar was selected because it had

previously demonstrated resistance against the clubroot pathotype at Normandin (D. Pageau,

personal communication). The trend of a decline in the concentration of resting spores as the

length of break interval increased from 0- to 2-year break from susceptible canola was also

observed in soil samples collected before and after one crop of 45H29 canola. However, there

was an 8- to 12-fold increase in the concentration of resting spores in the soil samples taken in

the fall of 2014 after harvest of 45H29, and the concentration of resting spores in soil increased

following 45H29 in two of three treatments in the study. In samples collected in the spring of

2015, resting spore concentration decreased by ≥ 90% overwinter. These results contradict

reports that resistant canola has either no effect or a small increase in spore concentration in soil

compared to an unplanted control (Wallenhammar et al., 2000; Hwang et al., 2013). In a field

trial near Edmonton, AB, a decline in resting spores was observed despite a 14% clubroot

incidence in the field planted with resistant canola (Hwang et al., 2013). Similarly, the

concentration of resting spores in soil was reduced by 36-45% following clubroot-resistant leafy

daikon cultivars (Raphanus sativus var. longipinnatus) in soil artificially inoculated with 106

spores mL-1 (Murakami et al., 2001). Clubroot severity on subsequent napa cabbage plants was

reduced compared to a fallow control. In another study, the concentration of resting spores was

reduced by 99.9% from the starting concentration of 5 x 106 spores mL-1 soil to 5 x 103 spores

mL-1 soil following four years of cultivation of kale (B. oleracea ssp. acephala) and turnip

Page 91: Effect of Plasmodiophora brassicae ... - University of Guelph

  84  (B. rapa ssp. rapa) (Yamagishi et al.,1986). In each of these previous trials, soil was artificially

inoculated with resting spores. As mentioned previously, it is possible that the viability of spores

applied by inoculation may decline at a different rate than spores under natural conditions. This

could explain why a much greater number of resting spores was contributed to the soil by a

resistant cultivar in an naturally infested field, compared with artificially inoculated soil.

However, the different trend observed in the current study may also be explained by the very high

level of clubroot symptoms on the resistant canola, which indicates that there was either a very

high percentage of susceptible canola volunteers or susceptible off-types, or that genetic

resistance was overcome by the pathogen. This latter possibility is being explored in another

study (B. Gossen, personal communication). Therefore, these results may not be indicative of

crop rotation with a resistant canola cultivar in a field where genetic resistance has not been

overcome. However, differences in experimental methods may have also played a role.

Significant reductions in resting spore concentration following a resistant cultivar were observed

in two studies in Japan where there is a different pathotype of P. brassicae than in Canada. In

addition, clubroot-resistant cultivars of B. oleracea and B. rapa may have a different mechanism

of resistance than clubroot-resistant B. napus. Finally, several of these studies estimated the

concentration of resting spores in soil using extraction from soil and direct counts. Estimates

using counts can be highly variable and often not correlated with results from qPCR (Cranmer,

2015).

In the current study, there was substantial variation in the amount of P. brassicae DNA

detected. Similar high variability has been reported in previous studies (Cranmer, 2015;

Wallenhammar et al., 2012), despite the steps taken to minimize variability. For example, each

subsample consisted of five cores rather than three, and cores were homogenized prior to DNA

extraction. Variability in resting spore counts increased at higher concentrations of resting spores

Page 92: Effect of Plasmodiophora brassicae ... - University of Guelph

  85  in soil, and was generally not normally distributed about the mean. Inhibition levels in the soil

samples from Normandin were low compared with other studies (Wallenhammar et al., 2012;

Cranmer, 2015), likely because there were comparatively low levels of organic matter in the soil

(~3.7%). One drawback to qPCR assessment of resting spore numbers is that viable spores

cannot be distinguished from non-viable spores (Faggian and Strelkov, 2009). As a result, qPCR

may overestimate the concentration of viable resting spores in soil relative to a bioassay.

Previous studies have reported that rainfall during the first 2 to 3 weeks after seeding is

correlated with clubroot incidence and severity (Thuma et al., 1983; Gludovacz, 2013). Canola is

normally seeded in late May or early June at Normandin, Quebec. In the current study, the most

recent susceptible canola crop was planted in the same year for each block within that treatment

of time break from canola. Therefore, it is conceivable that a year with higher than average

rainfall during the 2 weeks after seeding could lead to a higher initial spore concentration relative

to another year with lower rainfall. For example, monthly rainfall was lower than average by

12-36 mm from May to July 2007 and higher than average by 17-57 mm from June to July 2008.

However, the concentration of resting spores following a 6-year break from susceptible canola

(planted in 2007) was slightly higher than following a 5-year break from susceptible canola

(planted in 2008). In 2011, 2012 and 2013, rainfall was moderately higher than average in May

and June, which indicates that there was no substantial effect of rainfall on the inoculum level

prior to the break from susceptible canola. This supports the conclusions from recent studies

(Kasinathan 2012; Gossen et al., 2013; Cramner, 2015) that beyond certain minimal threshold

values of temperature and soil moisture, clubroot develops at high levels irrespective of weather

in temperate regions when inoculum concentration is high.

Quantification of resting spores of P. brassicae in soil can be valuable in evaluation of

clubroot management practices. Current crop rotation recommendations are canola one in four

Page 93: Effect of Plasmodiophora brassicae ... - University of Guelph

  86  years (3-year break from canola) on the Canadian prairies (Kutcher et al., 2013). A crop rotation

of 5 to 7 years (4- to 6-year break from canola) is recommended in Ontario for Brassica crops

including canola (OMAFRA, 2000).

The first objective of this research was assessed by quantifying the DNA of P. brassicae

in soil samples with break intervals of 0- to 6-years following susceptible canola, and in a

companion study with treatments of resistant canola following various break intervals of 0- to 2-

years following susceptible canola. The results provided strong support for the hypothesis that the

concentration of resting spores in soil declined exponentially with length of break interval

following a susceptible canola crop. The greatest reduction in concentration of resting spores was

observed after 1-year break from canola, with a further reduction after the second year of break.

One limitation of this study is that the treatments (different rotations) in Normandin are

not directly comparable due to different weather conditions in different years of the rotation. For

example, fallow for one treatment may be in a wet year, and fallow for another may be in a dry

year. The study design included three biological and three technical replications of soil samples

for qPCR amplification. However, replication of crop rotation treatments in the field was limited

to only two replicates (each with two sub-samples) and the trial in it entirety was not repeated.  

The second objective, to examine the interaction between crop rotation and inoculation

with P. brassicae on growth of clubroot-resistant canola, was assessed in a controlled

environment study. Resistant canola grown in soil from Scott, SK, was shorter in height,

indicating a greater cost of resistance than resistant canola grown in soil from Elora, ON. This

supports observations from Chapter 2 that the cost of resistance to P. brassicae causes greater

reductions in plant height when environmental conditions are more conducive to pathogen

development, including high soil moisture and more acidic soils. Susceptible canola grown in soil

from Scott, SK, had higher clubroot incidence and severity than susceptible canola grown in soil

Page 94: Effect of Plasmodiophora brassicae ... - University of Guelph

  87  from Elora, ON. This is likely related to the higher acidity of the soil from Scott, which makes

the site more conducive for clubroot development (Donald and Porter, 2004, Gossen et al., 2013;

Webster and Dixon, 1991a, 1991b).

No interaction between crop rotation and inoculation was observed. In Scott soil, plant

biomass increased by 42% in canola grown in soil following a 2-year break from canola, in one

repetition and a similar trend was observed in the second repetition although the difference was

not significant. These results are consistent with a previous study that investigated the effect of

crop rotation on canola in the absence of clubroot, where canola yield increased by 22% in a 1 in

6 year rotation (Harker et al., 2014). However, the decrease in plant biomass in canola grown in

Elora soil following a 2-year break from canola contradicts previous research that plant growth

increases with increasing diversity of crop rotation (Harker et al., 2014). This could be explained

by other differences in management of the Elora soil in addition to rotational diversity that were

not known, such as fertilizer or treatment with fungicide or insecticide that may have affected

microbial populations, or presence of other pathogens that were not consistent across rotational

treatments. Unlike in the location in Scott, SK, the sites at Elora Research Station were not

previously part of a replicated long-term crop rotation trial.

In the current study, field soil was used in conetainers to examine the effect of actual

cropping rotation that took place in a field rather than a simulated cropping history, for example

cycles of crops planted in tubs. Conducting crop rotation trials in controlled environments is not

recommended for future research. Plant growth was highly variable, likely due to disturbance of

the soil structure, disruption of microbial communities during drying and transport of soil, and

compaction from watering. This may explain why the results of the current study were

inconsistent and did not replicate previous results in field trials. Similar to the possible

Page 95: Effect of Plasmodiophora brassicae ... - University of Guelph

  88  differences between Elora soils, the differences between the soils from Elora, ON and Scott, SK

may be related to other soil micro-organisms that are pathogenic or beneficial to plant growth,

other environmental conditions such as pH, or agronomic practices such as crop type planted

during the break interval from canola (Table 3.2). One opportunity to improve crop rotation trials

for future research could be to investigate the effect of continuous resistant canola crops on plant

growth and yield, relative to resistant canola in rotation with fallow or non-host crops.

It was hypothesized that an additive effect of no crop rotation and P. brassicae would

reduce plant growth and development more in inoculated plants grown in soil with no break from

canola, compared with inoculated plants grown in soil with a 2-year break from canola. There

were several reasons to expect an additive effect, including allelopathy, presence of other root

pathogens, decreased soil nutrition, changes in soil structure, or even plant residues from

previous canola crops increasing the germination of resting spores. Some or all of these factors

were expected to contribute to reduced plant growth in the presence of P. brassicae in soils

previously cropped with continuous canola, compared with soils that had a 2-year break from

canola. However, this was not demonstrated in the current study. The effect of these factors was

likely limited because of the disruption of the soil, as discussed previously. As a result, there was

no main effect of crop rotation observed, and no interaction observed between crop rotation and

inoculation.

Examination of the effect of crop rotation on plant growth requires long-term planning

and implementation of a controlled experiment over many years in field plots uniform for pH,

soil structure, soil nutrient levels, fertilization and agronomic practices. In addition to the

challenges associated with a crop rotation trial, assessing the interaction between resting spore

concentration and crop rotation requires artificial inoculation of treatments to ensure that there is

a negative control with no resting spores. In the current study, an approach involving examination

Page 96: Effect of Plasmodiophora brassicae ... - University of Guelph

  89  of soil from field locations with no history of clubroot in a controlled environment facility was

selected instead of a field trial to avoid contamination of a field with a pathogen that did not yet

occur there.

An improvement to the method for the decline in resting spores study would be to

increase the soil sampling depth to 0-30 cm below the soil surface to include a greater proportion

of the total number of resting spores in the soil. Depending on soil type and tillage practices, the

concentration of resting spores in soil decreases with increasing depth (Cranmer, 2015). In a silt

loam soil at Bassano, AB, 54-65% of spores were observed in the top 0-16 cm at one site

following a 6-year break from susceptible canola and a 3-year break from resistant canola

(Cranmer, 2015). Low levels of resting spores were observed at 0-8 cm, and higher levels were

observed at 8-16 cm, 16-23 cm and 23-30 cm. This may indicate that few or no new resting

spores had been contributed to the soil near the surface in the past three or more years. In the

current study of a silt loam soil at Normandin, QC, samples were taken from 0-15 cm depth of

soil near the cortical zone of young roots (Gan et al., 2009), where the majority of clubroot

infection occurs. It is possible that an increased break interval from canola may provide an

opportunity for downward movement of resting spores rather than their death and disappearance.

The results reported here support the results of previous trials on crop rotation and

clubroot (Hwang et al., 2013; Peng et al., 2015). Research by Peng et al. (2015) assessed resting

spore concentrations in soil samples from the same site at AAFC Normandin Research Farm that

was assessed in the current study, but collected in two different years (2012, 2013). A break

interval ≥ 2-year following susceptible canola reduces the concentration of resting spores >90%

relative to 0- and 1-year break from canola (Peng et al., 2015). However, resting spore

concentration is not significantly different in soil samples following a 0- and 1-year break from

Page 97: Effect of Plasmodiophora brassicae ... - University of Guelph

  90  susceptible canola. In contrast, in the current study, a 97% reduction in the concentration of

resting spores was observed following 1-year break from susceptible canola relative to no break

from susceptible canola. Both studies quantified the concentration of resting spores in soil

samples using a qPCR protocol. The main difference in protocol was that an internal control was

included in the current study to adjust the final estimate of spore concentration based on

inhibition of pathogen DNA amplification. Using a protocol with an adjustment for internal

control resulted in a greater reduction in resting spore concentration from 0- to 1-year break

(97%) relative to the reduction observed from 0- to 1-year break without adjustment for internal

control (75%). It is likely that this was observed because inhibition was much higher in soil

samples collected following no break from canola, probably due to increased humic acids and

other organic compounds associated with plant residues. However, the difference may also be

related to limitations of the current study such as replication and repetition discussed above.

In conclusion, the results of this study indicate that when the initial concentration of

resting spores is high, increasing the length of break from canola up to 2 years can decrease the

fitness cost associated with clubroot resistance and increase the opportunity for high yield in a

resistant cultivar. In any case, where clubroot symptoms have been identified, a management

strategy should be implemented that includes a rotation interval ≥ 2 years break from canola

followed by planting of a clubroot-resistant cultivar.

Page 98: Effect of Plasmodiophora brassicae ... - University of Guelph

  91  CHAPTER FOUR

 GENERAL DISCUSSION

 Integrated clubroot management is necessary for the long-term viability of canola

production in Alberta and across the Canadian prairies. High concentrations of P. brassicae

resting spores in soil can reduce vegetative growth and seed production, delay the maturity of

canola and result in crop failure in susceptible canola. The primary strategy for clubroot

management in Canadian canola production is planting a clubroot-resistant cultivar, because

there are currently no viable cultural, biological and chemical strategies available to growers.

However, recent studies indicate that exposure to resting spores can result in failure to achieve

the yield potential of resistant cultivars (Hwang et al., 2013; Peng et al., 2015).

The current study investigated the effect of resting spore concentration and crop rotation on

growth of clubroot-resistant crops. The influence of spore concentration on growth of clubroot-

resistant cultivars was assessed in canola, napa cabbage and cabbage. Also, the effect of crop

rotation on resting spore concentration over time was examined. Another study investigated the

interaction of spore concentration and crop rotation on growth of clubroot-resistant canola under

controlled conditions. Finally, the interaction of spore concentration and pH on clubroot

incidence and development in susceptible canola was examined under controlled conditions.

This is the first study to specifically examine the cost of resistance to P. brassicae in

resistant cultivars of canola, napa cabbage and cabbage under a range of resting spore

concentrations. Resistant canola was 30% shorter, biomass was 43% lower and maturity was

delayed in the field site with higher spore concentration relative to a similar site with lower spore

concentration, in one of two years. The decrease in growth is consistent with previous studies

(Hwang et al., 2011a; Deora et al., 2012b), which indicated that there was a metabolic cost of

Page 99: Effect of Plasmodiophora brassicae ... - University of Guelph

  92  resistance to clubroot in canola under controlled conditions. In resistant napa cabbage, the cost of

resistance, indicated by leaf length, was consistent over both years for. Leaf length was reduced

by 31% in 2014 and 22% in 2015 at a site with high spore concentration. There were no

significant differences for cabbage.

A controlled environment study demonstrated that plant height of clubroot resistant canola

cv. 45H29 declined in a quadratic relationship with increasing concentration of resting spores.

Plant height was reduced by 12-14% compared to the non-inoculated control at 1 x 107 and

1 x 108 spores g-1 soil. Biomass of canola was also lower and maturity was delayed with

increasing concentration of resting spores in one of two repetitions of the trial.

This is the first study to examine the interaction of a range of pH levels with a range of

spore concentrations on clubroot incidence and severity in susceptible canola. It is known that

clubroot incidence and severity in susceptible cultivars varies depending on environmental

factors, such as temperature, pH and soil moisture (Cohoun, 1953; Sharma et al., 2011a, 2011b;

Gossen et al., 2013). Although it is known that the infection threshold for susceptible cultivars is

approximately 1000 spores g-1 soil (Donald and Porter, 2009; Faggian and Strelkov, 2009), the

interaction between spore concentration and environment is not well understood. The current

study indicated that inoculation with 1 x 103 spores ml-1 resulted in abundant clubbing symptoms

at pH 6.0-7.0, but fewer symptoms develop at lower or higher pH (5.5 or 7.5). These results

support previous reports that high pH suppresses clubroot severity at low concentrations of

resting spores, but is overwhelmed at high spore concentrations (Colhoun, 1953). This study

should be repeated. Future research could include extraction of DNA to investigate pathogen

colonization of plant root tissue, as well as extraction of RNA to investigate gene expression at

various time-points across a range of pH level and spore concentration. Increased understanding

of the interaction of spore concentration and pH could lead to more accurate predictions of

Page 100: Effect of Plasmodiophora brassicae ... - University of Guelph

  93  clubroot severity.

The current study demonstrated that large numbers of resting spores are lost from the

surface soil layers, where they are most effective at causing disease, in the 2 years following

susceptible canola. The concentration of resting spores in soil declined by 97% after a 1-year

break and 99% after a 2-year break from susceptible canola, but then declined very slowly (<1%)

over the next four years. These reults are consistent with previous reports (Hwang et al., 2013;

Peng et al., 2015). Although the concentration of resting spores after a 6-year break was still

much higher than the infection threshold for susceptible cultivars, there are benefits to reducing

spore concentration before seeding a resistant cultivar. Specifically, a crop rotation with a

≥ 2-year break from canola can reduce the concentration of resting spores and thus the fitness

cost of resistance to clubroot that would reduce the yield of the resistant canola crop. This may

also prolong the effectiveness of resistant cultivars by alleviating selection pressure with the

population of resting spores (Kiyosawa et al., 1982).

In the second year of this study, one crop of canola cv. 45H29 (resistant) was seeded in

field plots with break intervals of 0- to 2-years from susceptible canola. A 12-fold increase in

spore concentration when susceptible canola was followed by 45H29 is different from previous

reports of small changes or no difference following resistant canola (Wallenhammar et al., 2000;

Hwang et al., 2011). In the current study, a sampling date 8 months after harvest showed that

≥ 90% of resting spores in soil were lost. They may have disintegrated or moved downward more

than 15 cm from the soil surface over winter.

The presence of susceptible canola volunteers, weeds, or even a resistant cultivar can

increase resting spore germination (Macfarlane, 1970). For this reason, there has been speculation

that planting a resistant host could lead to germination of a large number of resting spores that do

Page 101: Effect of Plasmodiophora brassicae ... - University of Guelph

  94  not complete the lifecycle and thus reduce the overall concentration of resting spores in soil

(Murakami et al., 2001). In contrast, susceptible weeds can persist in a field over many years, and

dramatically reduce the effectiveness of crop rotation (Harker et al., 2014). In the current study,

the influence of growing one crop of resistant canola on the concentration of resting spores could

not be properly assessed because a very high level of clubroot symptoms developed on the

previously resistant canola cultivar 45H29 (D. Pageau, personal communication). This indicates

either a very high proportion of susceptible weeds / off-types, or that resistance in the canola

cultivar had been eroded.

A separate study is investigating the possibility that the pathogen has overcome clubroot

resistance at this site (B. Gossen et al., personal communication). Recently, cleaved amplified

polymorphic sequence (CAPS) markers have been developed to differentiate P. brassicae

populations and single-spore isolates for at least some of the new genotypes of the virulent

phenotype in Alberta, known as 5X (Cao et al., 2015). These markers could be used to analyze

the genetic similarity between the new pathotypes in Alberta with those from Normandin. Future

studies should investigate the influence of a resistant crop on spore concentration in soil in a field

trial with lower risk of the pathogen overcoming clubroot resistance during that growing season.

For example, this could include a field with slightly lower initial spore concentration, or a field

with no previous history of resistant canola.

There are several ways that the experiments described in this thesis could be improved.

First, clubroot severity was low to moderate in the susceptible control in studies of canola, napa

cabbage and cabbage under controlled conditions. There appeared to be problems with the

inoculum used in some of the controlled environment studies. The concentration of viable resting

spores may have been much lower than the estimation of 1 x 106 spores mL-1. Spore

concentration was estimated using haemocytometer spore counts, which cannot differentiate

Page 102: Effect of Plasmodiophora brassicae ... - University of Guelph

  95  between viable and non-viable resting spores. Also, the protocol used for extraction and

estimation of spores from soil, which was adapted by Sharma et al. (2012) from Dhingra and

Sinclair (1985), may overestimate the total number of resting spores because empty resting spore

shells, or possibly other debris, can be mistaken for spores of P. brassicae. A protocol is

currently being developed with improved filtration of the resting spore suspension prior to the

haemocytometer spore count (F. Al-Daoud, personal communication). Also, research is currently

underway to develop and test a protocol for quantification of viable P. brassicae resting spores

using propidium monoazide (PMA)-PCR (F. Al-Daoud, personal communication). PMA is a

photoreactive dye that is cell membrane-impermeable but can bind with DNA. It can modify non-

viable cells so that they are not amplified by PCR, while leaving the DNA of viable cells intact.

Other recent developments in quantification of resting spores in soil, including qPCR, droplet

digital PCR (ddPCR) and the loop mediated isothermal amplification (LAMP) assay, could

increase access to more accurate, more affordable and faster assessment of spore concentration,

relative to plant bioassays. Additional research should compare the accuracy of these methods of

quantification of resting spores in soil. In combination with improved site-specific spore

quantification, improved understanding of the relationship between spore concentration and seed

yield of canola under field conditions could aid growers and agronomists in selection of cultivars

and cropping intervals.

There are many challenges to investigating the cost of resistance in field trials. Finding a

suitable location to conduct the trial can be challenging. Field plots should have uniform soil

characteristics and nutrient levels, and differ only in spore concentration. However, introducing

pathogen inoculum to a field that is not currently infested would be detrimental to the goal of

reducing clubroot levels in soil. There is not currently a reliable method of eradicating resting

spores in soil. One potential option for future research on the metabolic cost of resistance would

Page 103: Effect of Plasmodiophora brassicae ... - University of Guelph

  96  be in the clubroot research nursery near Edmonton, AB. In the current study, the field trial was

located at Muck Crops Research Station in Holland Marsh, ON, where differences in spore

concentrations were already present within a small area of the field. The site with highest

concentration of resting spores is near the entrance to the field, where clubroot trials have taken

place in succession over many years, and also increased deposition of contaminated soil from

equipment entering the field may have occurred (Cao et al., 2009). Although research sites

nearby do not contain clubroot, equipment entering and leaving that site may have increased

clubroot due to compaction of soil from higher equipment traffic (Gossen et al., 2016).

In summary, the results of this study indicate that when the initial spore concentration in

soil is high, increasing the length of break from canola to 2 years can decrease the concentration

of resting spores, which in turn reduces the cost of resistance and maximizes the yield potential

of a resistant cultivar. A negative relationship was found between plant height and concentration

of resting spores in one of two field trials and under controlled conditions. Some of the variability

among trials could be due to reduced viability of some inoculum, variability among plants within

a study and environmental influences such as soil type, soil moisture or temperature. For canola

growers, a ≥ 2-year break from canola, in combination with a clubroot-resistant cultivar, is

recommended wherever clubroot occurs.

Page 104: Effect of Plasmodiophora brassicae ... - University of Guelph

  97  LITERATURE CITED

Abdelzaher, H.M.A. 2003. Biological control of root rot of cauliflower caused by Pythium ultimum var. ultimum using selected antagonistic rhizospheric strains of Bacillus subtilis. New. Zeal. J. Crop Hort. Sci. 31: 209–220    Apel, K. and Hirt, H. 2004. Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Ann. Rev. Plant Biol. 55: 373–99. Abukutsa, M.O. and Onyango, J.C. 2005. Conservation and seed production of African leafy vegetables at Maseno University botanic garden, Kenya. African Crop Science Conference Proceedings. 7: 1201–1204.    Adhikari, K. K. 2010. Effect of temperature, biofungicides and fungicides on clubroot of selected brassica crops. M.Sc. thesis, University of Guelph, Guelph, Ontario. Ahmed, H.U., Hwang, S-F., Strelkov, S.E., Gossen, B.D., Peng, G., Howard, R.J. and Turnbull, G.D. 2011. Assessment of bait crops to reduce inoculum of clubroot (Plasmodiophora brassicae) of canola. Can. J. Plant Sci. 91(3): 545–551. Aist, J.R. and Williams, P.H. 1971.  The cytology and kinetics of cabbage root hair penetration by Plasmodiophora brassicae. Can. J. Bot. 49: 2023-2034. Allen, J., Fraser, H. and Hallett, R. 2008. The swede midge – a pest of crucifer crops. OMAFRA Factsheet 08-007. Queen’s Printer of Ontario.    Al-Shehbaz, I.A., Beilstein, M.A. and Kellogg, E.A. 2006. Systematics and phylogeny of the Brassicaceae (Cruciferae): An overview. Plant Syst. Evol. 259: 89–120.    Appel, O. and Al-Shehbaz. I.A. 2003. Cruciferae. In Kubitzki, K. and Bayer, C. (eds.) The Families and Genera of Vascular Plants. Springer, Berlin. Volume 5: 75–174. Arabidopsis Genome Initiative. 2000. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature. 408(6814): 796–815.  Ayers, G.W. 1972. Races of Plasmodiophora brassicae infecting crucifer crops in Canada. Can. Plant Dis. Surv. 52: 77–81.  

Bardner, H.M., Edwards, C.A., Arnold, M.K. and Rogerson, J.P. 1971. The symptoms of attack by swede midge (Contarinia nasturtii) and effects on the yield of swedes. 14: 223–233.

Ben-Yephet, Y. and Szmulewich, Y. 1985. Inoculum levels of Verticillium dahliae in the soils of the hot semi-arid Negev region of Israel. Phytoparasitica, 13(3-4): 193–200.

Page 105: Effect of Plasmodiophora brassicae ... - University of Guelph

  98  Berg, T., Tesoriero, L. and Hailstones, D.L. 2005. PCR-based detection of Xanthomonas campestris pathovars in Brassica seed. Plant Pathol. 54(3): 416–427.    Bergelson, J. and Purrington, C.B. 1996. Surveying patterns in the cost of resistance in plants. American Naturalist. 536–558. Brandt, S.A. and Zentner, R.P. 1995. Crop production under alternate rotations on a Dark Brown Chernozemic soil at Scott, Saskatchewan. Can. J. Plant. Sci. 75: 789–794. Broadbent, D.E. 1957. Immediate memory and simultaneous stimuli. Q. J. Exp. Psychol. 9: 1–11.  

Brown, J.K.M. and Rant, J.C. 2013. Fitness costs and trade-offs of disease resistance and their consequences for breeding arable crops. Plant Pathol. 62: 83–95.

Buczacki, S.T. and Moxham, S.E. 1983. Structure of the resting spore wall of Plasmodiophora brassicae revealed by electron microscopy and chemical digestion. Transactions of the British Mycol. Soc. 81 (2): 221–231.   Buczacki, S.T. and Ockendon, J.G. 1979. Preliminary observations on variation in susceptibility to clubroot among collections of some wild crucifers. Ann. Appl. Biol. 92: 113–118.   Buczacki, S. T., Toxopeus, H., Mattusch, P., Johnston, T. D., Dixon, G. R. and Hobolth, L. A. 1975. Study of physiologic specialization in Plasmodiophora brassicae: Proposals for attempted rationalization through an international approach. Transactions of the British Mycological Society 65: 295–303. Buczacki, S.T., Ockendon, J.G. and Freeman, G.H. 1978. An analysis of some effects of light and soil temperature on clubroot disease. Ann. Appl. Biol. 88: 229–238.   Buerstmayr, H., Ban, T. and Anderson, J.A. 2009. QTL mapping and marker‐assisted selection for Fusarium head blight resistance in wheat: a review. Plant breeding. 128(1): 1–26. Burki, F., Kudryavtsev, A., Matz, M.V., Aglyamova, G.V., Bulman, S., Fiers, M. 2010. Evolution of Rhizaria: new insights from phylogenomic analysis of uncultivated protists. BMC Evol. Biol. 10(1): 377. Campbell, D.C. and Kondra, Z.P. 1978. Relationships among growth patterns, yield components and yield of rapeseed. Can. J. Plant Sci. 58(1): 87–93. Canola Council of Canada. 2011. The History of Canola [online] Available from  http://www.canolacouncil.org/oil-and-meal/what-is-canola/the-history-of-canola/  [accessed 12 December 2013].

Page 106: Effect of Plasmodiophora brassicae ... - University of Guelph

  99  Canola Council of Canada. 2013. Clubroot found in Manitoba canola. http://www.canolawatch.org/2013/09/25/clubroot-found-in-manitoba-canola/ [accessed 12 December 2013].  Canola Council of Canada. 2014. Clubroot pathogen shift: Management. http://www.canolawatch.org/2014/09/10/clubroot-pathogenshift-management/ Cao, T., Srivastava, S., Rahman, M.H., Kav, N.N.V., Hotte, N., Deyholos, M.K. and Strelkov, S.E. 2008. Proteome-level changes in the roots of Brassica napus as a result of Plasmodiophora brassicae infection. Plant Sci. 174:97–115.  Cao, T., Manolii, V. P., Hwang, S. F., Howard, R. J., and Strelkov, S. E. 2009. Virulence and spread of Plasmodiophora brassicae [clubroot] in Alberta, Canada. Can. J. Plant Pathol. 31: 321–329.    Cao, T., Hwang, S-F. and Strelkov, S.E. 2015. Differentiation of Plasmodiophora brassicae populations and single-spore isolates using CAPS markers. (Abstract) Botany 2015. Edmonton, AB. Cartea, M.E., Lema, M., Francisco, M. and Velasco, P. 2011. Basic information on vegetable brassica crops. In Genetics, Genomics and Breeding of Vegetable Brassicas. Sadowski, J. and Kole, C. (Eds). Enfield, NH: Science Publishers. P. 1–33.    Cathcart, R.J., Topinka, A.K., Kharbanda, P., Lange, R., Yang, R.C. and Hall, L.M. 2006. Rotation length, canola variety and herbicide resistance system affect weed populations and yield. Weed Sci. 54(4): 726–734. Cavalier-Smith, T. 1993. Kingdom Protozoa and its 18 phyla. Microbiol. Rev. 57(4): 953–994.    Cavalier-Smith, T. 2013. Symbiogenesis: mechanisms, evolutionary consequences, and systematic implications. Ann. Rev. of Ecology, Evolution, and Systematics. 44(1): 145–172. Champolivier, L. and Merrien, A. 1996. Effects of water stress applied at different growth stages to Brassica napus L. var. oleifera on yield, yield components and seed quality. Euro. J. Agr. 5(3): 153–160. Chapman, J.F., Daniels, R.W. and Scarisbrick, D.H. 1984. Field studies on 14 C assimilate fixation and movement in oil-seed rape (B. napus). J. Agr. Sci. 102(1): 23–31. Charron, C.S. and Sams, C.E. 2002. Impact of Glucosinolate Content in Broccoli (Brassica oleracea (Italica Group)) on Growth of Pseudomonas marginalis, a Causal Agent of Bacterial Soft Rot. Plant Dis. 86:629–632.    Cheah, L. H., Veerakone, S., and Kent, G. 2000. Biological control of clubroot on cauliflower with Trichoderma and Streptomyces spp. N Z. Plant Prot. Soc. 53: 18–21.

Page 107: Effect of Plasmodiophora brassicae ... - University of Guelph

  100  Chittem. K., Mansouripour, S.M., Del Rio Mendoza, L.E. 2014. First Report of Clubroot on Canola Caused by Plasmodiophora brassicae in North Dakota. Plant Dis. 98(10): 1438.

Chongo, G. and McVetty, P.B.E. 2001. Relationship of physiological characters to yield parameters in oilseed rape (Brassica napus L.). Can. J. Plant Sci. 81(1).

Chu, M., Song, T., Falk, K.C., Zhang, X., Liu, X., Chang, A. et al. 2015. Fine mapping of Rcr1 and analyses of its effect on transcriptome patterns during infection by Plasmodiophora brassicae. BMC genomics. 15(1): 1166. Clifton, R., Lister, R., Parker, K.L., Sappl, P.G., Elhafez, D., Millar, A.H. et al. 2005. Stress-induced co-expression of alternative respiratory chain components in Arabidopsis thaliana. Plant. Mol. Biol. 58: 193–212. Cober, E.R., Rioux, S., Rajcan, I., Donaldson, P.A. and Simmonds, D.H. 2003. Partial resistance to white mold in a transgenic soybean line. Crop Sci. 43(1): 92–95. Coley‐Smith, J.R., Mitchell, C.M. and Sansford, C.E. 1990. Long‐term survival of sclerotia of Sclerotium cepivorum and Stromatinia gladioli. Plant Pathol. 39(1): 58–69. Colhoun, J. 1952. Factors affecting the incidence of club root disease of Brassicae. Nature 169: 21–22.   Colhoun, J. 1953. A study of the epidemiology of club-root disease of Brassicae. Ann. Appl. Biol. 40: 262–283.    Colhoun, J. 1973. Effects of environmental factors on plant disease. Ann. Rev. Phytopathol. 11: 343–364. Conners, I.L., Shoemaker, R.A., and Creelman, D.W. 1956. Thirty-sixth annual report of the Canadian plant disease survey. Can. Plant Dis. Surv. 36: 52–53. Cranmer, T.J. 2015.  Vertical distribution of Plasmodiophora brassicae resting spores in soil and the effect of weather conditions on clubroot development. M.Sc. thesis, University of Guelph, Guelph, Ontario.  Creelman, R.A. and Mullet, J.E. 1995. Jasmonic acid distribution and action in plants: Regulation during development and response to biotic and abiotic stress. Proc. Natl. Acad. Sci. 92: 4114–4119. Crête, R., Laliberté, J. and Jasmin, J.J. 1963. Lutte chimique contre la hernie, Plasmodiophora brassicae Wor., des cruciferes en sols mineral et organique. Can. J. Plant Sci. 43(3): 349–354. Deblonde, P.M.K. and Ledent, J.F. 2001. Effects of moderate drought conditions on green leaf number, stem height, leaf length and tuber yield of potato cultivars. Eur. J. Agr. 14(1): 31–41.

Page 108: Effect of Plasmodiophora brassicae ... - University of Guelph

  101   Dekhiujzen, H.M. 1979. Electron microscope studies on the root hairs and cortex of a susceptible and a resistant variety of Brassica campestris infected with Plasmodiophora brassicae. Europ. J. Plant Pathol. (Neth. J. Plant Pathol.) 85: 1–17.    Delourme, R., Barbetti, M.J., Snowdon, R., Zhao, J. and Manzanares-Dauleux, M.J. 2012. In Edwards, D., Batley, J., Parkin, I. and Kole, C. (Eds.) Genetics, Genomics and Breeding of Oilseed Brassicas. Enfield, NH: Science Publishers. 276–318.    Deora, A., Gossen, B.D., Walley, F. and McDonald, M.R. 2011. Boron reduces development of clubroot in canola. Can. J. Plant Pathol. 33(4): 475-484. Deora, A., Gossen, B.D. and McDonald, M.R. 2012. Infection and development of Plasmodiophora brassicae in resistant and susceptible canola cultivars. Can. J. Plant Pathol. 34(2): 239–247.   Deora, A., Gossen, B.D. and McDonald, M.R. 2013. Cytology of infection, development and expression of resistance to Plasmodiophora brassicae in canola. Ann. Appl. Biol. 163: 56–71.   Deora, A., Gossen, B.D., Hwang, S.F., Pageau, D., Howard, R.J., Walley, F. and McDonald, M.R. 2014. Effect of boron on clubroot of canola in organic and mineral soils and on residual toxicity to rotational crops.  Can. J. Plant Sci. 94: 109 ︎ –118. Deora, A., Gossen, B.D., Amirsadeghi, S. and McDonald, M.R. 2015. A multiplex qPCR assay for detection and quantification of Plasmodiophora brassicae in soil. Plant Dis. 99(7): 1002–1009. De Vos, M. 2006. Herbivore-induced resistance against microbial pathogens in Arabidopsis. Mol. Plant Microbe In. 19: 1431–1443.    Dhingra, O.D. and Sinclair, J.B. 1985. Detection and estimation of inoculum. In Basic plant pathology methods. Dhingra, O.D. and Sinclair, J.B. (Eds.). Florida: CRC Press. 67–79. Diederichsen, E., Frauen, M., Linders, E. G. A., Hatakayema, K., and Hirai, M.  2009. Status and perspectives of clubroot resistance breeding in crucifer crops. J. Plant Growth Regul. 28: 265–281.    Dixon, G. R. 2006. The biology of Plasmodiophora brassicae Wor. A review of recent advances. Acta Hort. 706: 271–282.    Dixon, G. R. 2007. Vegetable Brassicas and Related Crucifers. UK: CABI.    Dixon, G. R. 2009. Plasmodiophora brassicae in its environment. Journal of Plant Growth Regulators. 28: 212–228.    Dobson, R. L., and Gabrielson, R. L. 1983. Role of primary and secondary  

Page 109: Effect of Plasmodiophora brassicae ... - University of Guelph

  102  zoospores of Plasmodiophora brassicae in the development of clubroot in Chinese  cabbage. Phytopathol. 73: 559–561.  Dobson, R.L., Gabrielson, R.L. and Baker, A.S. 1982. Soil water matric potential requirements for root-hair and cortical infection of Chinese cabbage by Plasmodiophora brassicae. Phytopathol. 72: 1598-1600.    Dokken-Bouchard, F.L., Bouchard, A.J., Ippolito, J., Peng, G., Strelkov, S.E., Kirkham, C.L., and Kutcher, H.R. 2010. Detection of Plasmodiophora brassicae in Saskatchewan, 2008. Can. Plant Dis. Surv. 90: 126.    Donald, E.C., and Porter, I.J. 2004. A sand—solution culture technique used to observe the effect of calcium and pH on root hair and cortical stages of infection by Plasmodiophora brassicae. Australas. Plant Path. 33: 585–589.    Donald, E.C., and Porter, I.J. 2009. Integrated control of clubroot. J. Plant Growth Regul. 28: 289–303.    Donald, E.C., Jaudzems, G., and Porter, I. J. 2008. Pathology of cortical invasion by Plasmodiophora brassicae in clubroot resistant and susceptible Brassica oleracea hosts. Plant Pathol. 57: 201–209.    Donald, E.C., Porter, L.J., Faggian, R., and Lancaster, R.A. 2006. An integrated approach to the control of clubroot in vegetable Brassica crops. Acta Hort. 706: 283– 300.    Drayton, F.L. 1926. A summary of the prevalence of plant diseases in the Dominion of Canada 1920-1924. Ottawa: F.A. Acland. Vol. 71. Evans, J.L. and Scholes, J.D. 1995. How does clubroot alter the regulation of carbon metabolism in its host? Asp Appl. Biol. 42: 125–132.    Faggian, R., and Strelkov, S. E. 2009. Detection and measurement of Plasmodiophora brassicae. J. Plant Growth Regul. 28: 282–288.    Fei, W., Feng, J., Rong, S., Strelkov, S., Gao, Z. and Hwang, S-F. 2016. Infection and gene expression of the clubroot pathogen Plasmodiophora brassicae in resistant and susceptible canola cultivars. Plant Dis. (In press) Feng, J., Hwang, R.U. and Hwang, S.F. 2010. Molecular characterization of a serine protease Pro1 from Plasmodiophora brassicae that stimulates resting spore germination. Mol. Plant Pathol. 11, 503–12.    Feng, J., Hwang, S.F. and Strelkov, S.E. 2012a. Analysis of expressed sequence tags derived from a compatible Plasmodiophora brassicae-canola interaction. Can. J. Plant Pathol. 34: 562–574.  

Page 110: Effect of Plasmodiophora brassicae ... - University of Guelph

  103  Feng, J., Xiao, Q., Hwang, S.F, Strelkov, S.E. and Gossen, B.D. 2012b. Infection of canola by secondary zoospores of Plasmodiophora brassicae produced on a nonhost. Eur. J. Plant Pathol. 132: 309–315.    Feng, J., Hwang, S.F. and Strelkov, S.E. 2013a. Assessment of gene expression profiles in primary and secondary zoospores of Plasmodiophora brassicae by dot blot and real-time PCR. Microbiol. Res. 168: 518–524.    Feng, J., Hwang, S.F. and Strelkov, S.E. 2013b. Studies into primary and secondary infection processes by Plasmodiophora brassicae on canola. Plant Pathol. 62: 177–183.  Fisher, R.A. 1930. The genetical theory of natural selection. Clarendon Press, Oxford. Fitt, B.D.L., Evans, N., Howlett, B.J. and Cooke, M. 2006. Sustainable strategies for managing Brassica napus (oilseed rape) resistance to Leptosphaeria maculans (Phoma stem canker). Eur. J. Plant Pathol. 114(1): 1–1.    Foulkes, M.J., Paveley, N.D., Worland, A., Welham, S.J., Thomas, J. and Snape, J.W. 2006. Major genetic changes in wheat with potential to affect disease tolerance. Phytopathol. 96: 680–688. Franzke, A., Lysak, M.A., Al-Shehbaz, I.A., Koch, M.A., Mummenhoff, K. 2011. Cabbage family affairs: the evolutionary history of Brassicaceae. Trends in Plant Sci. 16: 108–116. Freyman, S., Charnetski, W.A. and Crookston, R.K. 1973. Role of leaves in the formation of seed in rape. Can. J. Plant Sci. 53(3): 693–694. Friberg, H., Lagerlöf, J. and Rämert, B. 2005. Germination of Plasmodiophora brassicae resting spores stimulated by a non-host plant. Eur. J. Plant Pathol. 113: 275–281.    Fu, J., Liu, H., Li, Y., Yu, H., Li, X., Xiao, J. and Wang, S. 2011. Manipulating broad-spectrum disease resistance by suppressing pathogen-induced auxin accumulation in rice.  Plant Physiol. 155(1): 589–602.    Gan, Y.T., Campbell, C.A., Janzen, H.H., Lemke, R., Liu, L.P., Basnyat, P. and McDonald, C.L. 2009. Root mass for oilseed and pulse crops: Growth and distribution in the soil profile. Can. J. Plant Sci. 85: 883–893. Garber, R. C. and Aist, J. R. 1979. The ultrastructure of mitosis in Plasmodiophora brassicae (Plasmodiophorales). J. Cell Sci. 40: 89–110.   Gilligan, C.A., Simons, S.A. and Hide, G.A. 1996. Inoculum density and spatial pattern of Rhizoctonia solani in field plots of Solanum tuberosum: effects of cropping frequency. Plant Pathol. 45(2): 232–244.

Page 111: Effect of Plasmodiophora brassicae ... - University of Guelph

  104  Gludovacz, T. 2013. Clubroot in canola and cabbage in relation to soil temperature, plant growth and host resistance. M.Sc. thesis, University of Guelph, Guelph, Ontario.   Gossen, B.D, Adhikari, K.K.C. and McDonald, M.R. 2012. Effect of seeding date on development of clubroot in short-season Brassica crops, Canadian Journal of Plant Pathology. 34(4): 516–523.    Gossen, B.D., Kasinathan, H., Cao, T., Manolii, V., Hwang, S-F., Peng, G., Cao, T. and McDonald, M.R. 2013. Interaction of pH and temperature affect infection and symptom development of Plasmodiophora brassicae in canola. Can. J. Plant Pathol. 35(3). 294–303. Gossen, B.D., Kasinathan, H., Deora, A., Peng, G. and McDonald, M.R. 2016. Effect of soil type, organic matter content, bulk density and saturation on clubroot severity and biofungicide efficacy. Plant Pathol. In press.  Grsic-Rausch, S., Kobelt, P, Siemens, J.M., Bischoff, M. and Ludwig-Muller, J. 2000. Expression and Localization of Nitrilase during Symptom Development of the Clubroot Disease in Arabidopsis. Plant Physiol. 122(2): 369–378.    Guo, Z., Miyoshi, H., Komyoji, T., Haga, T. and Fujita, T. 1991. Uncoupling activity of a newly developed fungicide, fluazinam [3-chloro-N-(3-chloro-2,6-dinitro-4- trifluoromethylphenyl)-5-trifluoromethyl-2-pyridinamine]. Biochimica et Biophuysica Acta. 1056: 89–92. Hallett, R.H., Chen, M., Sears, M.K. and Shelton, A.M. 2009. Insecticide Management Strategies for Control of Swede Midge (Diptera: Cecidomyiidae) on Cole Crops. J. Econ. Entomol. 102(6): 2241–2254. Harker, K.N., O’Donovan, J.T., Turkington, T.K., Blackshaw, R.E., Lupwayi, N.Z., Smith, E.G., et al. 2015. Canola cultivar mixtures and rotations do not mitigate the negative impacts of continuous canola. Can. J. Plant Sci. 95(6): 1085–1099.    Hayward, A. 2012. Introduction – Oilseed Brassicas. In Edwards, D., Batley, J., Parkin, I. and Kole, C. (Eds.) Genetics, Genomics and Breeding of Oilseed Brassicas. Enfield, NH: Science Publishers. 1–13.    Health Canada. 2003. Low Erucic Acid Rapeseed (Lear) oil derived from canola-quality Brassica juncea (L.) CZERN. Lines PC 97-03, PC98-44 and PC98-45. [online] Available from  http://www.hc-sc.gc.ca/fn-an/gmf-agm/appro/low_erucic-faible_erucique-eng.php [accessed 12 December 2013].    Hildebrand, P. D. and McRae, K. B. 1998. Control of clubroot caused by Plasmodiophora brassicae with non-ionic surfactants. Can. J. Plant Pathol. 20: 1–11.    Hirai, M. 2006. Genetic analysis of clubroot resistance in Brassica crops. Breeding Sci. 56: 223–229.  

Page 112: Effect of Plasmodiophora brassicae ... - University of Guelph

  105    Hoshino, T. and Inagaki, F. 2012.  Molecular  quantification  of  environmental  DNA  using  microfluidics  and  digital  PCR. Syst. Appl. Microbiol. 35: 390–395.   Howard, R. J., Strelkov, S. E., and Harding, M. W. 2010. Clubroot of cruciferous crops - new perspective on an old disease. Can. J. Plant Pathol. 32: 43–57.    Humpherson-Jones, F. 1993. Effect of surfactants and fungicides on clubroot (Plasmodiophora brassicae) of Brassicas. Ann. Appl. Biol. 122: 457–465.    Humpherson-Jones, F.M. and Phelps, K. 1989. Climatic factors influencing spore production in Alternaria brassicae and Alternaria brassicicola. Ann. Appl. Biol. 114: 449–458.    Hwang, S-F, Strelkov, S.E, Turnbull, G.D., Manolii, V., Howard, R.J. and Hartman, M. 2008. Soil treatments and amendments for management of clubroot on canola in Alberta. Can. J. Plant Pathol. 90: 410 (Abstract)    Hwang, S.F., Ahmed, H.U., Zhou, Q., Strelkov, S.E., Gossen, B.D., Peng, G. and Turnbull, G.D. 2011a. Influence of cultivar resistance and inoculum density on root hair infection of canola (Brassica napus) by Plasmodiophora brassicae. Plant Pathol. 60: 820–829.   Hwang, S-F., Ahmed, H.U., Strelkov, S.E., Gossen, B.D., Turnbull, G.D., Peng, G., Howard, R.J. 2011b. Seedling age and inoculum density affect clubroot severity and seed yield in canola. Can. J. Plant Sci. 91: 183–190.    Hwang, S.-F., Strelkov, S.E., Gossen, B.D., Turnbull, G.D., Ahmed, H.U., and Manolii, V.P. 2011c. Soil treatments and amendments for amelioration of clubroot on canola. Canadian Journal of Plant Pathology. 91: 999–1010.    Hwang, S-F., Cao, T., Xiao, Q., Ahmed, H., Manolii, V.P., Turnbull, G. et al. 2012a. Efficacy of fungicide, seeding date and seedling age on clubroot severity, seedling emergence and yield of canola. Can. J. Plant Sci. 92(6): 1175–1186.    Hwang, S-F., Ahmed, H.U., Zhou, Q., Rashid, A., Strelkov, S.E., Gossen, B.D. 2012b. Assessment of the impact of resistant and susceptible canola on Plasmodiophora brassicae inoculum potential. Plant Pathol. 61: 945–952.    Hwang, S-F., Ahmed, H.U., Zhou, Q., Rashid, A., Strelkov, S.E., Gossen, B.D. et al. 2013. Effect of susceptible and resistant canola plants on Plasmodiophora brassicae resting spore populations in the soil. Plant Pathol. 62: 404–412. Hwang, S.F., Howard, R.J., Strelkov, S.E., Gossen, B.D. and Peng, G. 2014. Management of clubroot (Plasmodiophora brassicae) on canola (Brassica napus) in western Canada. Can. J. Plant Pathol. 36: 49–65.

Page 113: Effect of Plasmodiophora brassicae ... - University of Guelph

  106  Hwang, S-F., Ahmed, H.U., Zhou, Q-X., Turnbull, G.D., Strelkov, S.E., Gossen, B.D., and Peng, G. 2015. Effect of host and non-host crops on Plasmodiophora brassicae resting spore concentrations and clubroot of canola. Plant Pathol. 64(5): 1198–1206.

Ignatov, A.M., Artemyeva, A.N. and Hida, K. 2010. Origin And Expansion Of Cultivated Brassica Rapa In Eurasia: Linguistic Facts. Acta Hort. 867: 81–88.  

Ikegami, H., Mukobata, H. and Naiki, T. 1978. Scanning electron microscopy of Plasmodiophora brassicae in diseased root cells of turnip and Chinese cabbage (Studies on the Clubroot of Cruciferous Plants III). Ann. Phytopathol. Soc. Jpn. 44:456–464.    Ingram, D. S. and Tommerup, I. C. 1972. The life history of Plasmodiophora brassicae Woron. Proc. R. Soc. London Ser. B. Biological Sciences 180: 103–112.   Jameson, P.E. 2000. Cytokinins and auxins in plant-pathogen interactions - An overview. Plant Growth Regul. 32: 369–380.    Jaschke, D., Dugassa-Gobena, D., Karlovsky, P., Vidal, S. and Ludwig-Muller, J. 2010. Suppression of clubroot (Plasmodiophora brassicae) development in Arabidopsis thaliana by the endophytic fungus Acremonium alternatum. Plant Pathol. 59: 100–111.    Jeger, M.J. and Viljanen-Rollinson, S.L.H. 2001. The use of the area under the disease-progress curve (AUDPC) to assess quantitative disease resistance in crop cultivars. Theor. Appl. Gen. 102(1): 32–40. Jensen, B.D., Vaerbak, S., Munk, L. and Anderson, S.B. 1999. Characterization and inheritance of partial resistance to downy mildew, Peronospora parasitica, in breeding material of broccoli, Brassica oleraceaconvar. botrytis var. italic. Plant Breeding. 118(6): 549–554.    Johnson, R. 1984. A Critical Analysis of Durable Resistance. Annual Review of Phytopathol. 22: 309–330.    Kageyama, K., and Asano, T. 2009. Life cycle of Plasmodiophora brassicae. J. Plant Growth Regul. 28: 203–211.    Kammerich, J., Beckmann, S., Scharafatb, I. and Ludwig-Muller, J. 2014. Suppression of the clubroot pathogen Plasmodiophora brassicae by plant growth promoting formulations in roots of two Brassica species. Plant Pathol. 63: 846–857. Karling, J. S. 1968. The Plasmodiophorales. Hafner Publishing Company, Inc., New York.   Kasinathan, H. 2012. Influence of pH, temperature, and biofungicides on clubroot of canola. M.Sc. thesis, University of Guelph, Guelph, Ontario. Kim, C.H., Cho, W.D., Kim, H.M. 2000. Distribution of Plasmodiophora brassicae causing clubroot disease of Chinese cabbage in soil. Plant Dis. Res. 6(1): 27–32.

Page 114: Effect of Plasmodiophora brassicae ... - University of Guelph

  107   Kiyosawa, S. 1982. Genetics and epidemiological modeling of breakdown of plant disease resistance. Ann. Rev. Phytopathol. 20(1): 93–117. Knezevic, S.Z., Weise, S.F. and Swanton, C.J. 1994. Interference of Redroot Pigweed (Amaranthus retroflexus) in Corn (Zea mays). 42: 568–573. Krogman, K.K. and Hobbs, E.H. 1975. Yield and morphological response of rape (Brassica campestris L. cv. Span) to irrigation and fertilizer treatments. Can. J. Plant Sci. 55(4): 903–909. Kuginuki, Y., Yoshikawa, H. and Hirai, M. 1999. Variation in virulence of Plasmodiophora brassicae in Japan tested with clubroot-resistant cultivars of Chinese cabbage (Brassica rapa L. ssp. pekinensis). Eur. J. Plant Pathol. 105: 327–332.    Kutcher, H.R., Brandt, S.A., Smith, E.G., Ulrich, D., Malhi, S.S. and Johnston, A.M. 2013. Blackleg disease of canola mitigated by resistant cultivars and four-year crop rotations in western Canada. Can. J. Plant Pathol. 35(2): 209–221. Lahlali, R.L., Peng, G., McGregor, L., Gossen, B.D., Hwang, S.F. and McDonald, M.R. 2011. Mechanisms of the biofungicide Serenade (Bacillus subtilis QST713) in suppressing clubroot. Biocontrol Science and Technology, 21:11, 1351–1362  Lahlali, R.L., Peng, G., Gossen, B.D., McGregor, L., Yu, F.Q., Hynes, R.K. et al. 2013. Evidence that the biofungicide Serenade (Bacillus subtilis) suppresses clubroot on canola via antibiosis and induced host resistance. Phytopathol. 103(3): 245–254.    Lahlali, R., McGregor, L., Song, T., Gossen, B.D., Narisawa, K. and Peng, G. 2014. Heteroconium chaetospira induces resistance to clubroot via upregulation of host genes involved in jasmonic acid, ethylene and auxin biosynthesis. PLoS ONE 9(4): e94144.    Lamb, R.J. 1989. Entomology of Oilseed Brassica Crops. Annual Review of Entomol. 34: 211–229.    LeBoldus, J.M., Manolii, V.P., Turkington, T.K. and Strelkov, S.E. 2012. Adaptation to Brassica host genotypes by a single-spore isolate and population of Plasmodiophora brassicae (clubroot). Plant Dis. 96(6): 833–838. Lovelock, D.A., Donald, C.E., Conlan, X.A. and Cahill, D.M. 2013. Salicylic acid suppression of clubroot in broccoli (Brassica oleracea var. italica) caused by the obligate biotroph Plasmodiophora brassicae. Australas. Plant Path. 42, 141–53. Ludwig-Muller, J. 1993. Concentrations of indole-3-acetic acid in in plants of tolerant and susceptible varieties of Chinese cabbage infected with Plasmodiophora brassicae Woron. New Phytol. 125(4): 763–769.    

Page 115: Effect of Plasmodiophora brassicae ... - University of Guelph

  108  Ludwig-Muller, J., Bennett, R.N., Kiddle, G., Ihmig, S., Ruppell, M. and Hilgenberg, W. 1999. The host range of Plasmodiophora brassicae and its relationship to endogenous glucosinolate content. New Phytol. 141(3): 4443–4458.    Lund, R.E. 1975. Tables for an approximate test for outliers in linear models. Technometrics. 17(4): 473–476. Major, D.J., Bole, J.B. and Charnetski, W.A. 1978. Distribution of photosynthates after 14CO2 assimilation by stems, leaves, and pods of rape plants. Can. J. Plant Sci. 58(3): 783–787. Manzanares-Dauleux, M.J., Divaret, I., Baron, F. and Thomas, G. 2001. Assessment of biological and molecular variability between and within field isolates of Plasmodiophora brassicae. Plant Pathol. 50: 165–173.    Matheron, M.E. and Porchas, M. 2010. Evaluation of soil solarization and flooding as management tools for Fusarium wilt of lettuce. Plant Dis. 94(11): 1323–1328. Matheson, C.D., Gurney, C., Esau, N., Lehto, R. 2010. Assessing PCR inhibition from humic substances. The Open Enzyme Inhibition Journal. 3: 38–45.  McDonald, B.A. and Linde, C. 2002. Pathogen population genetics, evolutionary potential, and durable resistance. Annu. Rev. Phytopathol. 40: 349–379.    McDonald, M.R., Kornatowska, B. and McKeown, A.W. 2004. Management of clubroot of Asian Brassica crops grown on organic soils. Acta Horticulturae. 635: 25–30. McDonald, M.R. and Westerveld, S.M. 2008. Temperature prior to harvest influences the incidence and severity of clubroot on two Asian Brassica vegetables. HortScience 43: 1509–1513.    McDonald, M.R., Vander Kooi, K.D. and Westerveld, S.M. 2008. Effect of foliar trimming and fungicides on apothecial number of Sclerotinia sclerotiorum, leaf blight severity, yield, and canopy microclimate in carrot. Plant Dis. 92(1): 132–136.  McDonald, M.R., Sharma, K., Gossen, B.D., Deora, A., Feng, J. and Hwang, S.F. 2014. The role of primary and secondary infection in host response to Plasmodiophora brassicae. Phytopathol. 104(10): 1078–1087.    McKeen, C.D. and Wensley, R.N. 1961. Longevity of Fusarium oxysporum in soil tube culture. Science. 134(3489): 1528–1529. Macfarlane, I. 1952. Factors affecting the survival of Plasmodiophora brassicae Wor. in the soil and its assessment by a host test. Ann. Appl. Biol. 39: 239–256.    

Page 116: Effect of Plasmodiophora brassicae ... - University of Guelph

  109  Macfarlane, I. 1970. Germination of resting spores of Plasmodiophora brassicae. Trans. Br. Mycol. Soc. 55: 97–112.    Mitani, S. Sugimoto, K., Hayashi, I., Takii, Y., Ohshima, T. and Matsuo, N. 2003. Effects of cyazofamid against Plasmodiophora brassicae Woronin on Chinese cabbage. Pest Management Science. 59(3): 287–293. Møller, I.M. 2001. Plant mitochondria and oxidative stress: electron transport, NADPH turnover, and metabolism of reactive oxygen species. Ann. Rev. Plant Biol. 52(1): 561–591. Morgner, M. and Sacristan, M.D. 1995. Quantitative determination of colonization by Plasmodiophora brassicae (Wor.) in Brassica napus and Brassica oleracea. Cruciferae Newsletter 17: 80–1.  Muller, P. and Hilgenberg, W. 1986. Isomers of zeatin and zeatin riboside in clubroot tissue: evidence for trans-zeatin biosynthesis by Plasmodiophora brassicae. Physologia Plantarum. 66(2): 245–250. Munkholm, L.J., Heck, R.J. and Deen, B. 2013. Long-term rotation and tillage effects on soil structure and crop yield. Soil and Tillage Research. 127: 85–91. Murakami, H., Tsushima, S., Akimoto, T. and Shishido, Y. 2001. Reduction of spore density of Plasmodiophora brassicae in soil by decoy plants. J. Gen. Plant Pathol. 67(1): 85–88. Murakami, H., Tsushima, S. and Shishido, Y. 2002. Factors affecting the pattern of the dose response curve of clubroot disease caused by Plasmodiophora brassicae. Soil Biol. Biochem. 32(11-12): 1637–1642.  

Myers, D.F. and Campbell, R.N. 1985. Lime and the control of clubroot of crucifers: effects of ph, calcium, magnesium and their interactions. Phytopathol. 75(6): 670–673.

Narisawa, K., Ohiki, K.T., Hasiba, T. 2000. Suppression of clubroot and Verticillium yellows in Chinese cabbage in the field by the root endophytic fungus, Heterconium chaetospira. Plant Pathol. 49: 141–146.    Ontario Ministry of Agriculture, Food, and Rural Affairs (OMAFRA). 2000. Vegetable production recommendations. Publ. 363. Queen’s Printer for Ontario, Toronto. Or, D. and Wraith, J.M. 2002. Soil physics companion: Soil water content and water potential relationships. CRC Press Boca Raton. 49–84. Pal, K.K. and McSpadden Gardener, B. 2006. Biological Control of Plant Pathogens. The Plant Health Instructor.  

Page 117: Effect of Plasmodiophora brassicae ... - University of Guelph

  110  Palafox-Carlos, H., Ayala-Zavala, J.F. and Gonzalez-Aguilar, G.A. 2011. The Role of Dietary Fiber in the Bioaccessibility and Bioavailability of Fruit and Vegetable Antioxidants. J. Food Sci. 76(1): R6–R15. Park, J.E., Park, J.Y., Kim, Y.S., Staswick, P.E., Jeon, J., Yun, J. et al. 2007. GH3-mediated auxin homeostasis links growth regulation with stress adaptation response in Arabidopsis. J. Biol. Chem. 282(13): 10036–10046. Peng, G., Lahlali, R., Hwang, S.F., Hynes, R.K., McDonald, M.R., Pageau, D. et al. 2011. Control of clubroot on canola using the biofungicide Serenade plus cultivar resistance. In Proceedings of the 13th International Rapeseed Congress; 2011 Jun 5; Prague, Czech Republic. 1134–1137. Peng, G., Lahlali, R., Hynes, R.H., Gossen, B.D., Falk, K.C., Yu, F. et al. 2013. Assessment of crop rotation, cultivar resistance and Bacillus subtilis biofungicide for control of clubroot on canola. Proc. VIth IS on Brassicas and XVIIIth Crucifer Genetics Workshop. Eds.: Branca, F. and Tribulato, A. Acta Hort. 1005, ISHS. 591–598. Peng, G., Lahlali, R., Hwang, S-F., Pageau, D., Hynes, R.K., McDonald, M.R. et al. 2014. Crop rotation, cultivar resistance, and fungicides/biofungicides for managing clubroot (Plasmodiophora brassicae) on canola. Can. J. Plant Pathol. 36: sup1. 99–112. Peng, G., Pageau, D., Strelkov, S.E., Gossen, B.D., Hwang, S.F., Lahlali, R.L. 2015. A >2-year break from susceptible canola reduces Plasmodiophora brassicae resting spores in heavily infested soil and maximizes canola yield. Eur. J. Agron. 70: 78–84. Piao, Z., Ramchiary, N. and Lim, Y.P. 2009. Genetics of clubroot resistance in Brassica species. J. Plant Growth Regul. 28: 252–264. Porter, I.J., Merrimana, P.R. and Keaneb, P.J. 1991. Soil solarisation combined with low rates of soil fumigants controls clubroot of cauliflowers, caused by Plasmodiophora brassicae Woron. Austr. J. Exp. Agr. 31: 843–851. Purrington, C.B. 2000. Costs of resistance. Curr. Opin. Plant Biol. 3: 305–308. Rahman, H., Shakir, A. and Hasan, M.J. 2011. Breeding for clubroot resistant spring canola (Brassica napus L.) for the Canadian prairies: Can the European winter canola cv. Mendel be used as a source of resistance? Can. J. Plant Sci. 91(3): 447–458.

Rahman, H., Peng, G., Yu, F., Falk, K.C., Kulkarni, M. and Selvaraj, G. 2014. Genetics and breeding for clubroot resistance in Canadian spring canola (Brassica napus L.). Can. J. Plant Pathol. 36(sup1): 122-134. Rakow, G. 2004. Species origin and economic importance of Brassica. Biotechnology in Agriculture and Forestry: Brassica. Pua, E.C. and Douglas, J. (Eds.) Germany: Springer-Verlag. 3–7.  

Page 118: Effect of Plasmodiophora brassicae ... - University of Guelph

  111  Ren, J.P., M.H. Dickson, and E.D. Earle. 2000. Improved resistance to bacterial soft rot by protoplast fusion between Brassica rapa and B. oleracea. Theor. Appl. Genet. 100(5): 810–819.    Rennie, D.C, Holtz, M.D., Turkington, T.K., LeBoldus, J.M., Hwang, S-F., Howard, R.J. and Strelkov, S.E. 2015. Movement of Plasmodiophora brassicae resting spores in windblown dust. Can. J. Plant Pathol. 37(2): 188–196.

Reyes, A.A, Davidson, T.R. and Marks, C.F. 1974. Races, pathogenicity and chemical control of Plasmodiophora brassicae in Ontario. Phytopathol. 64: 173–177. Richards, F.J. 1959. A flexible growth function for empirical use. J. Exp. Botany. 10(2): 290–301. Richardson, L.T. and Munnecke, D.E. 1964. Effective fungicide dosage in relation to inoculum concentration in soil. Can. J. Bot. 42(3): 301–306. Rimmer, S.R., Kutcher, H.R. and Morrall, R.A.A. 2003. Diseases of canola and mustard. In Diseases of field crops in Canada. Bailey, K.L., Gossen, B.D., Gugel, R.K. and Morrall, R.A.A. (Eds.) Saskatoon, SK: Canadian Phytopathol. Soc. 129–146. Saude, C., McKeown, A., Gossen, B.D., McDonald, M.R. 2012. Effect of host resistance and fungicide application on clubroot pathotype 6 in green cabbage and napa cabbage. HortTechnology. 22: 311–319.    Scott, J., Gordon, T., Kirkpatrick, S., Koike, S., Matheron, M., Ochoa, O., et al. 2012. Crop rotation and genetic resistance reduce risk of damage from Fusarium wilt in lettuce. California Agriculture. 66(1): 20–24. Scott, J.C., McRoberts, D.N. and Gordon, T.R. 2014. Colonization of lettuce cultivars and rotation crops by Fusarium oxysporum f. sp. lactucae, the cause of fusarium wilt of lettuce. Plant Pathol. 63(3): 548–553. Sevilla, F., Jimenez, A. and Lazarro, J.J. 2015. What do the plant mitochondrial antioxidant and redox systems have to say under salinity, drought and extreme temperature? In Reactive oxygen spiecies and oxidative damage in plants under stress. Gupta, D.K., Corpas, F.J. and Palma, J.M. (Eds.). Switzerland: Springer. 23–56. Sharma, K., Gossen, B.D. and McDonald, M.R. 2011a. Effect of temperature on primary infection by Plasmodiophora brassicae and initiation of clubroot symptoms. Plant Pathol. 60: 830–838.    Sharma, K., Gossen, B.D. and McDonald, M.R. 2011b. Effect of temperature on cortical infection by Plasmodiophora brassicae and clubroot severity. Phytopathol. 101: 1424–1432.  

Page 119: Effect of Plasmodiophora brassicae ... - University of Guelph

  112  Short, D.P.G., Sandoya, G., Vallad, G.E., Koike, S.T., Xiao, C.-L., Wu, B.-M., et al. 2015. Dynamics of verticillium species microsclerotia in field soils in response to fumigation, cropping patterns, and flooding. Phytopathol. 105(5): 638–45.

Siemens, J., Nagel, M., Ludwig-Müller, J., Sacristan, M.D. 2002. The interaction of Plasmodiophora brassicae and Arabidopsis thaliana: parameters for disease quantification and screening of mutant lines. J. Phytopathol. 150:592–605.    Siemens,  J.,  Keller,  I.,  Sarx,  J.,  Kunz,  S.,  Schuller,  A.,  Nagel,  W.  et  al.  2006.  Transcriptome  analysis  of  Arabidopsis  clubroots  indicate  a  key  role  for  cytokinins  in  disease  development.  Mol.  Plant  Microbe  In.  19:480–494.   Simko, I. and Piepho, H.P. 2012. The area under the disease progress stairs: calculation, advantage, and application. Phytopathol. 102(4): 381–389. Singh, J., Upadhyay, A.K., Prasad, K., Bahadur, A. and Rai, M. 2007. Variability of carotenes, vitamin C, E and phenolics in Brassica vegetables. J. Food Comp. Anal. 20: 106–112.    Snowdon, R., Luhs, W. and Friedt, W. 2007. Oilseed rape. In Genome Mapping and Molecular Breeding in Plants, Vol. 2: Oilseeds. Kole, C. (Ed.). Germany: Springer. 55–114.   Soengas, P., Sotelo, T., Velasco, P. and Cartea, M.E. 2011. Antioxidant Properties of Brassica Vegetables. Funct. Plant Sci. Biotechnol. 5 (Special Issue 2): 43–5.    Statistics Canada. 2013. Field Crop Reporting Series. Statistics Canada. Catalogue No. 22–002-x.   Strelkov, S.E. and Hwang, S.-F. 2014. Clubroot in the Canadian canola crop: 10 years into the outbreak. Can. J. Plant Pathol. 36(sup1): 27–36. Strelkov, S.E, Tewari, J.P., Hartman, M. and Orchard, D. 2005. Clubroot on Canola in Alberta in 2003 and 2004. Can. Plant. Dist. Surv. 85: 72–73.   Strelkov, S.E., Tewari, J.P. and Smith-Degenhardt, E. 2006. Characterization of Plasmodiophora brassicae populations from Alberta, Canada.    Strelkov, S.E., Hwang, S.F., Howard, R.J., Hartman, M. and Turkington, T.K. 2011. Progress towards the sustainable management of clubroot [Plasmodiophora brassicae] of canola on the Canadian prairies. Prairie Soils Crops. 4: 114–121. Sturhan, D. 1985. Species, subspecies, race and pathotype problems in nematodes. EPPO Bulletin. 15(2): 139–144.   Suzuki, K., Matumiya, E., Ueno, Y. and Mizutani, J. 1992. Some properties of germination-stimulating factor from plants for resting spores of Plasmodiophora brassicae. Ann Phytopathol Soc Jpn. 58:699–705.  

Page 120: Effect of Plasmodiophora brassicae ... - University of Guelph

  113    Tanaka, S., Kochi, S., Kunita, H., Ito, S. and Kameya-Iwaki, M. 1999. Biological mode of action of the fungicide flusalfamide against Plasmodiophora brassicae (clubroot). Eur. J. Plant Pathol. 105: 577–584.    Tanina, K., Tojo, M., Date, H., Nasu, H. and Kasuyama, S. 2004. Pythium rot of chinensai (Brassica campestris L. chinensis group) caused by Pythium ultimum var. ultimum and P. aphanidermatum. J. General Pl. Pathol. 70: 188–191.    Tewari, J.P., Strelkov, S.E., Orchard, D., Hartman, M., Lange, R.M. and  Turkington, T.K. 2005. Identification of clubroot of crucifers on canola (Brassica napus) in Alberta. Can. J. Plant Pathol. 27: 143–144.    Thuma, B.A., Rowe, R.C. and Madden, L.V. 1983. Relationships of soil temperature and moisture to clubroot (Plasmodiophora brassicae) severity on radish in organic soil. Plant Dis. 67: 758–762.  

Tian, D., Traw, M., Chen, J., Kreitman, M. and Bergelson, J. 2003. Fitness costs of R-gene-mediated resistance in Arabidopsis thaliana. Nature. 423(May): 74–77.

Tommerup, I. C. and Ingram, D. S. 1971. The life-cycle of Plasmodiophora brassicae Woron. in Brassica tissue cultures and in intact roots. New Phytol. 70: 327–332.    Traka, M.H., Spinks, C.A., Doleman, J.F., Melchini, A., Ball, R.Y., Mills, R.D. and Mithen, R.F. 2010. The dietary isothiocyanate sulforaphane modulates gene expression and alternative gene splicing in a PTEN null preclinical murine model of prostate cancer.  Molecular Cancer. 9: 189.    Tsunoda, S. 1980. Eco-physiology of wild and cultivated forms in Brassica and allied genera. In Tsunoda, S., Hinata, K. and Gomez-Campo, C. (Eds). Brassica crops and their wild allies, biology and breeding. Tokyo: Japan Scientific Societies Press. 109–120.    U, N. 1935. Genome analysis in Brassica with special reference to the experimental formation of B. napus and peculiar mode of fertilization. Jpn. J. Bot 7: 389–452.    Ueno, H., Matsumoto, E., Aruga, D., Kitagawa, S., Matsumara, H. and Havashida, N. 2012. Molecular characterization of the CRa gene conferring clubroot resistance in Brassica rapa. Plant Mol. Biol. 80(6): 621–629.    Van der Plank, J.E. 1963. Plant diseases: Epidemics and control. New York and London: Academic Press. Van Loon, L.C., Bakker, P.A. and Pieterse, C.M. 1998. Systemic resistance induced by rhizosphere bacteria. Annu. Rev. Phytopathol. 36: 453-483.  

Page 121: Effect of Plasmodiophora brassicae ... - University of Guelph

  114  Vogel, J. P., Raab, T. K., Schiff, C. and Somerville, S. C. 2002. PMR6, a pectate lyase-like gene required for powdery mildew susceptibility in Arabidopsis. The Plant Cell. 14(9): 2095–2106.

Volossiouk, T., Robb, E.J., and Nazar, R.N. 1995. Direct DNA extraction for PCR-mediated assays of soil organisms. Appl. Environ. Microbiol. 61(11): 3972.    Voorrips, R.E., Jongerius, M.C. and Kanne, H.J. 2003. Quantitative Trait Loci for Clubroot Resistance in Brassica oleracea. Biotechnology in Agriculture and Forestry, Vol 52. Brassicas and Legumes. Nagata, T. and Tabata, S. (Eds.) Germany: Springer-Verlag.    Vos, J., van Loon, C.D. and Bollen, G.J. 2012. Effects of crop rotation on potato production in the temperate zones: proceedings of the international conference on effects of crop rotation on potato production in the temperate zones. (40). The Netherlands: Springer Science & Business Media. Wallenhammar, A.C. 1996. Prevalence of Plasmodiophora brassicae in a spring oilseed rape growing area in central Sweden and factors influencing soil infestation levels. Plant Pathol. 45: 710–719.    Wallenhammar, A.C., Johnsson, L. and Gerhardson, B. 1999. Clubroot resistance and yield loss in spring oilseed turnip rape and spring oilseed rape. In Wratten, N. and Salisbury, P.A. New Horizons of an Old Crop. Proceedings of the 10th International Rapeseed Congress. Canberra, Australia.    Wang, K., Kang, L., Anandi, A., Lazarovits, G. and Mysore, K.S. 2007. Monitoring in planta bacterial infection at both cellular and whole-plant levels using the green fluorescent protein variant GFPuv. New Phytol. 174: 212–223. Wally, O., Jayaraj, J. and Punja, Z.K. 2009. Broad-spectrum disease resistance to necrotrophic and biotrophic pathogens in transgenic carrots (Daucus carota L.) expressing an Arabidopsis NPR1 gene. Planta. 231(1): 131–141.    Walters, D.R., Ratsep, J. and Havis, N.D. 2013. Controlling crop diseases using induced resistance: challenges for the future. J. Exp. Bot. 64 (5): 1263–1280.    Webster, M.A. and Dixon, G.R. 1991a. Calcium, pH and inoculum concentration influencing colonization by Plasmodiophora brassicae. Mycol. Res. 95: 64–73.    Webster, M.A., and Dixon, G.R. 1991b. Boron, pH and inoculum concentration influencing colonization by Plasmodiophora brassicae. Mycol. Res. 95: 74–79.  White, J.G. and Buczacki, S.T. 1977. The control of clubroot by soil partial sterilization: a review. Ann. Appl. Biol. 85(2): 287–300.

Page 122: Effect of Plasmodiophora brassicae ... - University of Guelph

  115  Williams, P.H. and McNabola, S.S. 1967. Fine structure of Plasmodiophora brassicae in sporogenesis. Can. J. Bot. 45(9): 1665–1669.    Williams, P.H. 1966. A system for the determination of races of Plasmodiophora  brassicae that infect cabbage and rutabaga. Phytopathol. 56: 624–626.    Williams, P.H. 1980. Black Rot: A Continuing Threat to World Crucifers. Plant Dis. 64(8): 736–742.    Xue, S., Cao, T., Howard, R.J., Hwang, S-F. and Strelkov, S.E. 2008. Isolation and variation in virulence of single-spore isolates of Plasmodiophora brassicae from Canada. Plant Dis. 92: 456–462.   Yamagishi, H., Yoshikawa, H., Ashizawa, M., Hida, K.I. and Yui, S. 1986. Effects of resistant plants as a catch crop on the reduction of resting spores of clubroot (Plasmodiophora brassicae Worn.) in soil. J. Japan. Soc. Hort. Sci. 54(4): 460–466.

Yang, D.-L., Yao, J., Mei, C.-S., Tong, X.-H., Zeng, L.-J., Li, Q. et al. 2012. PNAS Plus: Plant hormone jasmonate prioritizes defense over growth by interfering with gibberellin signaling cascade. Proceedings of the National Academy of Sciences, 109(19): E1192–E1200.

Zhao, J. and Meng, J. 2003. Detection of loci controlling seed glucosinolate content and their association with Sclerotinia resistance in Brassica napus. 122(1): 19–23.  

Page 123: Effect of Plasmodiophora brassicae ... - University of Guelph

  116  

APPENDIX 1: SUPPLEMENTARY TABLES FOR CHAPTER TWO

Table A1.1. Controlled environment study – canola: CI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 0 - - Repetition (R) 1 0 - - Error 3 0 Table A1.2. Controlled environment study – canola: DSI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 141 1.80 0.32 Repetition (R) 1 1147 14.64 0.03 Error 3 78 Table A1.3. Controlled environment study – canola: plant height at 6 WAI Source df Mean Square F value Pr>F Block (B) 3 16 2.13 0.12 Inoculation (I) 5 24 3.21 0.02 Repetition (R) 1 29 3.80 0.06 I x R 5 12 1.60 0.19 I linear (1) 40 5.22 0.03 I quadratic (1) 69 9.07 0.005 Error 33 8 Table A1.4. Controlled environment study – canola: Biomass (dry shoot weight) Source df Mean Square F value Pr>F Block (B) 3 2 2.16 0.11 Inoculation (I) 5 2 2.07 0.09 Repetition (R) 1 120 105.42 <0.0001 I x R 5 0 0.41 0.84 I linear (1) 4 3.68 0.06 I quadratic (1) 1 0.62 0.43 Error 33 1 Table A1.5. Controlled environment study – canola, repetition 1: Biomass (dry shoot weight) Source df Mean Square F value Pr>F Block (B) 3 6 10.01 0.0007 Inoculation (I) 5 1 2.41 0.09 I linear (1) 3 5.60 0.03 I quadratic (1) 0 0.05 0.82 Error 15 1

Page 124: Effect of Plasmodiophora brassicae ... - University of Guelph

  117  Table A1.6. Controlled environment study – canola, repetition 2: Biomass (dry shoot weight) Source df Mean Square F value Pr>F Block (B) 3 3 3.84 0.03 Inoculation (I) 5 1 2.03 0.13 I linear (1) 1 1.67 0.22 I quadratic (1) 1 1.49 0.24 Error 15 1 Table A1.7. Controlled environment study – canola, repetition 1: Maturation at 8 WAI Source df Mean Square F value Pr>F Inoculation (I) 5 968 6.95 0.0009 I linear (1) 3177 22.80 0.0002 I quadratic (1) 1314 9.43 0.007 I cubic (1) 12 0.08 0.78 Error 18 139 Table A1.8. Large pot studies – outdoor: CI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 60 1.69 0.27 Repetition (R) 2 7086 200.72 <0.0001 Error 6 35 Table A1.9. Large pot studies – outdoor: DSI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 145 1.18 0.39 Repetition (R) 2 5378 43.77 0.0003 Error 6 123 Table A1.10. Large pot studies – outdoor: plant height 6 WAI Source df Mean Square F value Pr>F Block 3 97 3.48 0.03 Repetition (R) 1 5146 185.71 <0.0001 Inoculation (I) 5 331 1.19 0.33 I linear (1) 27 0.96 0.34 I quadratic (1) 60 2.15 0.15 Error 33 28

Table A1.11. Large pot studies – outdoor, 2015 repetition 1: plant height 6 WAI Source df Mean Square F value Pr>F Block 3 18 1.30 0.31 Inoculation (I) 5 10 0.74 0.60 I linear (1) 32 2.36 0.15 I quadratic (1) 0 0.00 0.95 Error 15 14

Page 125: Effect of Plasmodiophora brassicae ... - University of Guelph

  118  Table A1.12. Large pot studies – outdoor, 2015 repetition 2: plant height 6 WAI Source df Mean Square F value Pr>F Block 3 150 4.52 0.02 Inoculation (I) 5 49 1.49 0.25 I linear (1) 3 0.08 0.78 I quadratic (1) 114 3.43 0.08 Error 15 33

Table A1.13. Large pot studies – outdoor, 2015: plant height 7 WAI Source df Mean Square F value Pr>F Block 3 40 2.14 0.11 Repetition 1 3977 214.39 <0.0001 Inoculation (I) 5 38 2.03 0.10 I linear (1) 0 0.03 0.88 I quadratic (1) 13 6.07 0.02 Error 33 19 Table A1.14. Large pot studies – outdoor, 2015 repetition 1: plant height 7 WAI Source df Mean Square F value Pr>F Block 3 6 0.43 0.73 Inoculation (I) 5 6 0.38 0.85 I linear (1) 8 0.55 0.47 I quadratic (1) 1 0.05 0.83 Error 15 14

Page 126: Effect of Plasmodiophora brassicae ... - University of Guelph

  119  Table A1.15. Large pot studies – outdoor, 2015 repetition 2: plant height 7 WAI Source df Mean Square F value Pr>F Block 3 47 2.00 0.16 Inoculation (I) 5 58 2.45 0.08 I linear (1) 3 0.14 0.71 I quadratic (1) 202 8.50 0.01 Error 15 24 Table A1.16. Controlled environment study – canola, napa cabbage, cabbage: CI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 168 1.18 0.37 Repetition (R) 1 2030 14.25 0.004 Inoculation (I) 1 6392 44.87 <0.0001 R x I 1 2030 14.25 0.004 Error 9 142 Table A1.17. Controlled environment study – canola, napa cabbage, cabbage: DSI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 33 1.94 0.19 Repetition (R) 1 693 40.74 0.0001 Inoculation (I) 1 1442 84.77 <0.0001 R x I 1 693 40.74 0.0001 Error 9 17 Table A1.18. Controlled environment study – canola, napa cabbage, cabbage: CI in susceptible napa cabbage Source df Mean Square F value Pr>F Block (B) 3 196 0.95 0.46 Repetition (R) 1 79 0.39 0.55 Inoculation (I) 1 4747 23.08 0.00 R x I 1 79 0.39 0.55 Error 9 206 Table A1.19. Controlled environment study – canola, napa cabbage, cabbage: DSI in susceptible napa cabbage Source df Mean Square F value Pr>F Block (B) 3 46 0.67 0.59 Repetition (R) 1 1 0.01 0.91 Inoculation (I) 1 1652 24.13 0.0008 R x I 1 1 0.01 0.91 Error 9 68

Page 127: Effect of Plasmodiophora brassicae ... - University of Guelph

  120  Table A1.20. Controlled environment study – canola, napa cabbage, cabbage: CI in susceptible cabbage Source df Mean Square F value Pr>F Block (B) 3 9 0.07 0.97 Repetition (R) 1 768 5.67 0.05 Inoculation (I) 1 4400 32.46 0.0007 R x I 1 751 5.54 0.05 Error 7 136 Table A1.21. Controlled environment study – canola, napa cabbage, cabbage: DSI in susceptible cabbage Source df Mean Square F value Pr>F Block (B) 3 6 0.12 0.94 Repetition (R) 1 95 2.11 0.19 Inoculation (I) 1 729 16.23 0.005 R x I 1 88 1.97 0.20 Error 7 45 Table A1.22. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: plant height of resistant canola at 2 WAI Random Effects Estimate Standard Error Z Pr > Z Block 0.0 0.02 0.19 0.42 Residual 0.2 0.05 3.24 0.0006 Fixed Effects Num df Den df F Pr > F Cultivar 3 21 1.37 0.28 Inoculation 1 21 3.24 0.09 Cultivar*Inoculation 3 21 1.64 0.21 Contrast 1 21 5.40 0.03

Table A1.23. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: plant height of resistant canola at 3 WAI Random Effects Estimate Standard Error Z Pr > Z Block 0.0 - - - Residual 0.2 0.05 3.46 0.0003 Fixed Effects Num df Den df F Pr > F Cultivar 3 21 0.90 0.46 Inoculation 1 21 1.68 0.21 Cultivar*Inoculation 3 21 0.98 0.42 Contrast 1 21 3.05 0.10 Table A1.24. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: plant height of resistant canola at 4 WAI Random Effects Estimate Standard Error Z Pr > Z Block 0 - - - Residual 0.2 0.05 3.46 0.0003

Page 128: Effect of Plasmodiophora brassicae ... - University of Guelph

  121  Fixed Effects Num df Den df F Pr > F Cultivar 3 21 4.76 0.01 Inoculation 1 21 1.09 0.31 Cultivar*Inoculation 3 21 2.12 0.13 Contrast 1 21 3.54 0.07

Table A1.25. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: plant height of resistant canola at 5 WAI Random Effects Estimate Standard Error Z Pr > Z Block 0.0 - - - Residual 0.5 0.14 3.46 0.0003 Fixed Effects Num df Den df F Pr > F Cultivar 3 21 11.15 0.0001 Inoculation 1 21 1.46 0.24 Cultivar*Inoculation 3 21 0.81 0.50 Contrast 1 21 3.27 0.08

Table A1.26. Controlled environment study – canola, napa cabbage, cabbage repetition 1: plant height at 6 WAI of resistant canola Random Effects Estimate Standard Error Z Pr > Z Block 0.0 - - - Residual 0.1 0.33 3.46 0.0003 Fixed Effects Num df Den df F Pr > F Cultivar 3 21 13.09 <0.0001 Inoculation 1 21 2.11 0.16 Cultivar*Inoculation 3 21 0.61 0.61 Contrast 1 21 3.30 0.08

Table A1.27. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: Area under growth stairs (AUGS) resistant canola 2-6 WAI Random Effects Estimate Standard Error Z Pr > Z Block 0.0 - - - Residual 375.9 108.52 3.46 0.0003 Fixed Effects Num df Den df F Pr > F Cultivar 3 21 7.81 0.001 Inoculation 1 21 3.50 0.08 Cultivar*Inoculation 3 21 0.25 0.86 Contrast 1 21 2.49 0.13

Page 129: Effect of Plasmodiophora brassicae ... - University of Guelph

  122  Table A1.28. Controlled environment study – canola, napa cabbage, cabbage, repetition 1: Maturity Random Effects Estimate Standard Error Z Pr > Z Block 0.0 - - - Residual 264.9 76.47 3.46 0.0003 Fixed Effects Num df Den df F Pr > F Inoculation 1 21 0.00 1.00 Maturity 3 21 19.86 <0.0001 Inoculation*Maturity 3 21 0.38 0.77 Contrast (veg) Contrast (bud) Contrast (flower) Contrast (pod)

1 1 1 1

21 21 21 21

30.38 0.10 4.64 9.25

<0.0001 0.75 0.04 0.006

Table A1.29. Controlled environment study – canola, napa cabbage, cabbage: leaf length at 5 WAI of resistant napa cabbage Random Effects Estimate Standard Error Z Pr > Z Repetition 6.9 9.81 0.71 0.24 Block 0.2 0.16 1.01 0.16 Residual 0.6 0.11 5.10 <0.0001 Fixed Effects Num df Den df F Pr > F Cultivar 3 52 3.57 0.02 Inoculation 1 52 6.92 0.01 Cultivar*Inoculation 3 52 0.99 0.41 Contrast 1 52 7.76 0.007

Page 130: Effect of Plasmodiophora brassicae ... - University of Guelph

  123  Table A1.30. Controlled environment study – canola, napa cabbage, cabbage: biomass (dry shoot weight) of resistant napa cabbage Random Effects Estimate Standard Error Z Pr > Z Repetition 6.9 9.81 0.71 0.24 Block 0.2 0.16 1.01 0.16 Residual 0.6 0.11 5.10 <0.0001 Fixed Effects Num df Den df F Pr > F Cultivar 3 52 3.57 0.02 Inoculation 1 52 6.92 0.01 Cultivar*Inoculation 3 52 0.99 0.41 Contrast 1 52 7.76 0.007

Table A1.31. Controlled environment study – canola, napa cabbage, cabbage: leaf length at 6 WAI of resistant cabbage Random Effects Estimate Standard Error Z Pr > Z Repetition 3.2 4.62 0.70 0.24 Block 0.0 0.04 0.17 0.43 Residual 0.5 0.10 4.67 <0.0001 Fixed Effects Num df Den df F Pr > F Cultivar 3 44 5.50 0.00 Inoculation 1 44 2.85 0.09 Cultivar*Inoculation 3 44 0.85 0.47 Contrast 1 44 1.49 0.23

Table A1.32. Controlled environment study – canola, napa cabbage, cabbage: biomass (dry shoot weight, g) of resistant cabbage Random Effects Estimate Standard Error Z Pr > Z Repetition 7.1 11.8 0.61 0.27 Block 51.0 44.6 1.14 0.13 Residual 28.6 6.1 4.70 <0.0001 Fixed Effects Num df Den df F Pr > F Cultivar 3 44 0.71 0.55 Inoculation 1 44 0.00 0.95 Cultivar*Inoculation 3 44 0.48 0.70 Contrast 1 44 0.32 0.57

Page 131: Effect of Plasmodiophora brassicae ... - University of Guelph

  124  Table A1.33. Field trials – CI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 273 1.00 0.43 Repetition (R) 1 0 0.00 <0.0001 Site (S) 2 4840 17.71 0.0003 R x S 1 0 0.00 <0.0001 Error 12 273 Table A1.34. Field trials – DSI in susceptible canola Source df Mean Square F value Pr>F Block (B) 3 126 1.46 0.28 Repetition (R) 1 4225 48.73 <0.0001 Site (S) 2 6684 77.10 <0.0001 R x S 1 25 0.29 0.60 Error 12 87 Table A1.35. Field trials – Levene’s test, homogeneity of variance: plant height of resistant canola Source df Mean Square F value Pr>F Site x Repetition 4 11206 1.74 0.16 Error 55 6454 Table A1.36. Field trial, 2014 – Levene’s test, homogeneity of variance: plant height of resistant canola Source df Mean Square F value Pr>F Site (S) 1 13640 1.92 0.18 Error 22 7093 Table A1.37. Field trial, 2015 – Levene’s test, homogeneity of variance: plant height of resistant canola Source df Mean Square F value Pr>F Site (S) 2 15743 2.54 0.09 Error 33 5431 Table A1.38. Field trials – plant height at 8 WAI of resistant canola Source df Mean Square F value Pr>F Site (S) 3 27225 12.15 <0.0001 Repetition (R) 1 876 377.41 <0.0001 S x R 0 - - - Error 55 72 Table A1.39. Field trial, 2014 – plant height at 8 WAI of resistant canola Source df Mean Square F value Pr>F Site (S) 1 2330 30.55 <0.0001 Error 22 76

Page 132: Effect of Plasmodiophora brassicae ... - University of Guelph

  125  Table A1.40. Field trial, 2015 – plant height at 8 WAI of resistant canola Source df Mean Square F value Pr>F Site (S) 2 149 2.15 0.13 Error 33 69 Table A1.41. Field trials – Levene’s test, homogeneity of variance: biomass of resistant canola Source df Mean Square F value Pr>F S x R 4 15307 1.69 0.17 Error 55 6028 Table A1.42. Field trial, 2014 – Levene’s test, homogeneity of variance: biomass of resistant canola Source df Mean Square F value Pr>F Site (S) 2 4 x 109 2.30 0.14 Error 22 2 x 109 Table A1.43. Field trial, 2015 – Levene’s test, homogeneity of variance: biomass of resistant canola Source df Mean Square F value Pr>F Site (S) 2 15307 2.54 0.09 Error 33 6028 Table A1.44. Field trials – biomass of resistant canola Source df Mean Square F value Pr>F Site (S) 2 49447 1.66 0.20 Repetition (R) 1 48186 1.62 0.20 S x R 1 166218 5.59 0.02 Error 75 29754 Table A1.45. Field trial, 2014 – biomass Source df Mean Square F value Pr>F Site (S) 2 297131 9.50 0.006 Error 22 31286 Table A1.46. Field trial, 2015 – biomass Source df Mean Square F value Pr>F Site (S) 2 990 0.05 0.95 Error 33 18023 Table A1.47. Field trials – Levene’s test, homogeneity of variance: maturity of resistant canola Source df Mean Square F value Pr>F S x R 4 289000 4.77 0.0022 Error 55 60606

Page 133: Effect of Plasmodiophora brassicae ... - University of Guelph

  126  Table A1.48. Field trial, 2014 – Levene’s test, homogeneity of variance: maturity of resistant canola Source df Mean Square F value Pr>F Site (S) 1 33750 0.22 0.64 Error 22 151515 Table A1.49. Field trial, 2015 – Levene’s test, homogeneity of variance: maturity of resistant canola Source df Mean Square F value Pr>F Site (S) 2 0 - - Error 33 0 Table A1.50. Field trials – maturity of resistant canola Source df Mean Square F value Pr>F Site (S) 2 7725 63.41 <0.0001 Repetition (R) 1 46875 384.79 <0.0001 S x R 1 6075 49.87 <0.0001 Error 55 122 Table A1.51. Field trial, 2014 – maturity Source df Mean Square F value Pr>F Site (S) 1 12150 39.90 <0.0001 Error 22 305 Table A1.52. Field trial, 2015 – maturity Source df Mean Square F value Pr>F Site (S) 2 0 - - Error 33 0 Table A1.53. Field trials – CI in ‘susceptible’ napa cabbage Source df Mean Square F value Pr>F Block (B) 3 72 0.28 0.84 Repetition (R) 1 15188 59.56 <0.0001 Site (S) 2 2278 8.93 0.004 R x S 1 4556 17.87 0.001 Error 12 255 Table A1.54. Field trials – DSI in ‘susceptible’ napa cabbage Source df Mean Square F value Pr>F Block (B) 3 12 0.12 0.95 Repetition (R) 1 2756 36.67 0.0001 Site (S) 2 479 4.68 0.03 R x S 1 958 9.35 0.001 Error 12 102

Page 134: Effect of Plasmodiophora brassicae ... - University of Guelph

  127  Table A1.55. Field trial, 2014 – CI in susceptible napa cabbage Source df Mean Square F value Pr>F Site (S) 1 6171 11.87 0.02 Error 5 520 Table A1.56. Field trial, 2014 – DSI in susceptible napa cabbage Source df Mean Square F value Pr>F Site (S) 1 1230 5.82 0.06 Error 5 211 Table A1.57. Field trials – leaf length of resistant napa cabbage Source df Mean Square F value Pr>F Site (S) 2 1882 52.96 <0.0001 Repetition (R) 1 272 36.35 <0.0001 S x R 1 9 1.74 0.19 Error 67 5 Table A1.58. Field trials – biomass of resistant napa cabbage Source df Mean Square F value Pr>F Site (S) 2 1543411 5.68 0.005 Repetition (R) 1 1004429 3.70 0.06 S x R 1 41007 0.15 0.70 Error 67 271514 Table A1.59. Raw data for controlled environment study – canola: CI & DSI

Repetition Spores ml-1 Cultivar Block CI DSI 1 1x107 ACSN39 1 100.0 76.9 1 1x107 ACSN39 2 100.0 50.0 1 1x107 ACSN39 3 100.0 77.8 1 1x107 ACSN39 4 100.0 81.0 1 0 45H29 1 0.0 0.0 1 0 45H29 2 0.0 0.0 1 0 45H29 3 0.0 0.0 1 0 45H29 4 0.0 0.0 1 1x104 45H29 1 0.0 0.0 1 1x104 45H29 2 0.0 0.0 1 1x104 45H29 3 0.0 0.0 1 1x104 45H29 4 0.0 0.0 1 1x105 45H29 1 0.0 0.0 1 1x105 45H29 2 0.0 0.0 1 1x105 45H29 3 0.0 0.0 1 1x105 45H29 4 0.0 0.0 1 1x106 45H29 1 0.0 0.0

Page 135: Effect of Plasmodiophora brassicae ... - University of Guelph

  128  1 1x106 45H29 2 0.0 0.0 1 1x106 45H29 3 0.0 0.0 1 1x106 45H29 4 0.0 0.0 1 1x107 45H29 1 0.0 0.0 1 1x107 45H29 2 0.0 0.0 1 1x107 45H29 3 0.0 0.0 1 1x107 45H29 4 0.0 0.0 1 1x108 45H29 1 0.0 0.0 1 1x108 45H29 2 0.0 0.0 1 1x108 45H29 3 0.0 0.0 1 1x108 45H29 4 0.0 0.0 2 1x107 ACSN39 1 100.0 92.3 2 1x107 ACSN39 2 100.0 92.6 2 1x107 ACSN39 3 100.0 88.9 2 1x107 ACSN39 4 100.0 87.5 2 0 45H29 1 0.0 0.0 2 0 45H29 2 0.0 0.0 2 0 45H29 3 0.0 0.0 2 0 45H29 4 0.0 0.0 2 1x104 45H29 1 0.0 0.0 2 1x104 45H29 2 0.0 0.0 2 1x104 45H29 3 0.0 0.0 2 1x104 45H29 4 0.0 0.0 2 1x105 45H29 1 0.0 0.0 2 1x105 45H29 2 0.0 0.0 2 1x105 45H29 3 0.0 0.0 2 1x105 45H29 4 0.0 0.0 2 1x106 45H29 1 0.0 0.0 2 1x106 45H29 2 0.0 0.0 2 1x106 45H29 3 0.0 0.0 2 1x106 45H29 4 0.0 0.0 2 1x107 45H29 1 0.0 0.0 2 1x107 45H29 2 0.0 0.0 2 1x107 45H29 3 0.0 0.0 2 1x107 45H29 4 0.0 0.0 2 1x108 45H29 1 0.0 0.0 2 1x108 45H29 2 0.0 0.0 2 1x108 45H29 3 0.0 0.0 2 1x108 45H29 4 0.0 0.0 3 1x108 ACSN39 1 100.0 92.6 3 1x108 ACSN39 2 100.0 92.6 3 1x108 ACSN39 3 100.0 96.3 3 1x108 ACSN39 4 100.0 100.0

Page 136: Effect of Plasmodiophora brassicae ... - University of Guelph

  129  3 0 45H29 1 0.0 0.0 3 0 45H29 2 0.0 0.0 3 0 45H29 3 0.0 0.0 3 0 45H29 4 0.0 0.0 3 1x104 45H29 1 0.0 0.0 3 1x104 45H29 2 0.0 0.0 3 1x104 45H29 3 0.0 0.0 3 1x104 45H29 4 0.0 0.0 3 1x105 45H29 1 0.0 0.0 3 1x105 45H29 2 0.0 0.0 3 1x105 45H29 3 0.0 0.0 3 1x105 45H29 4 0.0 0.0 3 1x106 45H29 1 0.0 0.0 3 1x106 45H29 2 0.0 0.0 3 1x106 45H29 3 0.0 0.0 3 1x106 45H29 4 0.0 0.0 3 1x107 45H29 1 0.0 0.0 3 1x107 45H29 2 0.0 0.0 3 1x107 45H29 3 0.0 0.0 3 1x107 45H29 4 0.0 0.0 3 1x108 45H29 1 0.0 0.0 3 1x108 45H29 2 0.0 0.0 3 1x108 45H29 3 0.0 0.0 3 1x108 45H29 4 0.0 0.0

Table A1.60. Raw data for controlled environment study – canola: plant height

Repetition Spores g-1 soil Cultivar Block 2 WAI 3 WAI 4 WAI 5 WAI 6 WAI

1 0 45H29 1 3.99 7.82 17.77 25.67 30.74 1 0 45H29 2 3.29 6.82 16.70 24.07 27.38 1 0 45H29 3 3.25 7.25 18.20 24.55 28.93 1 0 45H29 4 3.50 6.88 14.16 27.44 36.56 1 1x104 45H29 1 3.50 7.11 14.55 25.45 31.53 1 1x104 45H29 2 3.44 5.84 14.26 24.91 28.98 1 1x104 45H29 3 3.04 6.54 13.06 21.65 28.77 1 1x104 45H29 4 3.38 6.16 15.52 26.56 33.41 1 1x105 45H29 1 3.78 6.20 15.08 24.58 29.28 1 1x105 45H29 2 3.96 8.70 23.57 34.45 36.21 1 1x105 45H29 3 3.63 6.90 17.82 28.20 31.43 1 1x105 45H29 4 3.87 6.67 15.16 26.05 28.86 1 1x106 45H29 1 3.49 5.79 12.98 23.47 28.10

Page 137: Effect of Plasmodiophora brassicae ... - University of Guelph

  130  1 1x106 45H29 2 3.43 5.73 12.71 23.06 27.29 1 1x106 45H29 3 3.46 6.35 17.11 29.38 32.23 1 1x106 45H29 4 3.47 6.26 13.44 22.73 27.16 1 1x107 45H29 1 3.93 5.51 10.33 17.33 18.27 1 1x107 45H29 2 4.10 6.60 15.32 25.41 28.58 1 1x107 45H29 3 3.49 6.99 16.81 23.43 26.72 1 1x107 45H29 4 2.95 4.92 11.41 20.24 23.65 1 1x108 45H29 1 3.87 5.84 11.50 19.75 22.01 1 1x108 45H29 2 2.73 4.88 12.44 20.71 26.96 1 1x108 45H29 3 3.37 6.26 13.75 24.40 33.50 1 1x108 45H29 4 3.10 5.11 13.57 21.04 25.36 2 0 45H29 1 3.06 4.57 6.43 17.77 26.36 2 0 45H29 2 4.26 8.06 14.19 32.20 34.08 2 0 45H29 3 4.98 9.33 14.02 26.70 33.88 2 0 45H29 4 4.73 9.40 17.54 29.24 30.56 2 1x104 45H29 1 3.03 3.63 4.50 7.82 13.39 2 1x104 45H29 2 3.08 4.63 6.96 15.82 24.88 2 1x104 45H29 3 5.11 9.08 14.60 34.86 34.81 2 1x104 45H29 4 3.93 6.54 8.28 14.75 19.63 2 1x105 45H29 1 2.93 3.63 4.68 8.38 14.23 2 1x105 45H29 2 3.90 7.24 10.66 21.73 27.06 2 1x105 45H29 3 4.27 7.01 10.62 20.87 27.27 2 1x105 45H29 4 5.04 8.32 14.38 25.91 29.43 2 1x106 45H29 1 2.64 3.30 4.38 8.42 14.46 2 1x106 45H29 2 3.55 6.01 8.76 18.35 28.44 2 1x106 45H29 3 4.90 9.36 16.29 33.14 34.92 2 1x106 45H29 4 4.03 8.55 12.86 25.59 28.13 2 1x107 45H29 1 2.89 3.68 5.29 11.08 18.68 2 1x107 45H29 2 3.55 5.36 7.62 15.29 22.51 2 1x107 45H29 3 4.34 7.32 12.28 25.79 26.55 2 1x107 45H29 4 4.00 6.28 8.99 19.50 25.59 2 1x108 45H29 1 2.85 3.78 4.90 9.00 17.61 2 1x108 45H29 2 3.50 4.88 7.24 15.18 22.11 2 1x108 45H29 3 5.32 9.19 14.35 30.58 33.75 2 1x108 45H29 4 4.17 6.29 11.32 25.17 29.29 3 0 45H29 1 4.09 7.44 15.84 . 34.42 3 0 45H29 2 3.88 7.45 21.49 . 32.81 3 0 45H29 3 3.57 5.80 12.68 . 31.69 3 0 45H29 4 3.41 6.23 16.39 . 30.55 3 1x104 45H29 1 3.71 6.79 16.47 . 26.56 3 1x104 45H29 2 3.76 6.20 14.15 . 32.71 3 1x104 45H29 3 3.23 6.37 16.08 . 30.21 3 1x104 45H29 4 3.88 8.04 21.50 . 32.04

Page 138: Effect of Plasmodiophora brassicae ... - University of Guelph

  131  3 1x105 45H29 1 3.97 7.03 13.26 . 27.89 3 1x105 45H29 2 3.84 6.55 15.94 . 31.59 3 1x105 45H29 3 3.15 6.35 16.56 . 30.17 3 1x105 45H29 4 4.05 7.93 17.53 . 29.98 3 1x106 45H29 1 4.04 6.97 16.40 . 30.50 3 1x106 45H29 2 3.45 7.58 19.11 . 32.80 3 1x106 45H29 3 3.13 5.25 10.33 . 28.72 3 1x106 45H29 4 3.54 6.38 13.31 . 28.56 3 1x107 45H29 1 4.43 7.70 15.69 . 27.28 3 1x107 45H29 2 3.70 6.88 14.82 . 32.10 3 1x107 45H29 3 3.36 6.82 15.73 . 31.28 3 1x107 45H29 4 3.80 7.13 16.94 . 30.27 3 1x108 45H29 1 3.28 5.96 12.68 . 26.84 3 1x108 45H29 2 3.95 7.10 16.49 . 30.01 3 1x108 45H29 3 3.28 6.03 14.78 . 28.71 3 1x108 45H29 4 4.03 8.16 16.82 . 29.33

Table A1.61. Raw data for controlled environment study – canola: biomass (dry shoot weight) of resistant canola

Repetition Spores ml-1 Block Mean dry shoot weight of 10 plants (g)

1 0 1 6.13 1 0 2 5.78 1 0 3 5.07 1 0 4 8.71 1 1x104 1 5.08 1 1x104 2 5.69 1 1x104 3 5.75 1 1x104 4 6.56 1 1x105 1 4.78 1 1x105 2 6.98 1 1x105 3 7.96 1 1x105 4 7.83 1 1x106 1 4.30 1 1x106 2 4.91 1 1x106 3 6.78 1 1x106 4 6.74 1 1x107 1 5.64 1 1x107 2 4.86 1 1x107 3 6.16 1 1x107 4 7.55 1 1x108 1 4.24 1 1x108 2 4.48

Page 139: Effect of Plasmodiophora brassicae ... - University of Guelph

  132  1 1x108 3 5.55 1 1x108 4 6.43 2 0 1 . 2 0 2 5.40 2 0 3 4.30 2 0 4 2.20 2 1x104 1 3.10 2 1x104 2 3.10 2 1x104 3 2.90 2 1x104 4 3.40 2 1x105 1 2.20 2 1x105 2 4.10 2 1x105 3 4.20 2 1x105 4 3.20 2 1x106 1 2.40 2 1x106 2 3.50 2 1x106 3 5.10 2 1x106 4 2.80 2 1x107 1 2.50 2 1x107 2 3.30 2 1x107 3 2.30 2 1x107 4 3.40 2 1x108 1 2.60 2 1x108 2 2.90 2 1x108 3 2.60 2 1x108 4 3.40 3 0 1 11.33 3 0 2 11.56 3 0 3 9.56 3 0 4 8.56 3 1x104 1 7.89 3 1x104 2 10.67 3 1x104 3 8.89 3 1x104 4 8.70 3 1x105 1 9.13 3 1x105 2 10.90 3 1x105 3 8.89 3 1x105 4 8.73 3 1x106 1 7.90 3 1x106 2 10.20 3 1x106 3 8.50 3 1x106 4 8.40 3 1x107 1 8.20 3 1x107 2 9.20

Page 140: Effect of Plasmodiophora brassicae ... - University of Guelph

  133  3 1x107 3 8.30 3 1x107 4 9.60 3 1x108 1 9.00 3 1x108 2 8.40 3 1x108 3 9.10 3 1x108 4 8.40 Table A1.62. Raw data for controlled environment study – canola: maturity (percentage of plants at seedpod stage) of resistant canola

Repetition Resting spores g-1 soil Block Percentage of plants at pod development stage

1 0 1 72.7 1 0 2 90.9 1 0 3 60.0 1 0 4 91.7 1 1x104 1 75.0 1 1x104 2 91.7 1 1x104 3 81.8 1 1x104 4 83.3 1 1x105 1 83.3 1 1x105 2 100.0 1 1x105 3 90.9 1 1x105 4 90.0 1 1x106 1 75.0 1 1x106 2 81.8 1 1x106 3 100.0 1 1x106 4 58.3 1 1x107 1 54.5 1 1x107 2 58.3 1 1x107 3 72.7 1 1x107 4 54.5 1 1x108 1 54.5 1 1x108 2 40.0 1 1x108 3 63.6 1 1x108 4 40.0

Page 141: Effect of Plasmodiophora brassicae ... - University of Guelph

  134  Table A1.63. Raw data for large pot studies – outdoor, 2014 and 2015: CI & DSI

Year Repetition Spores g-1 soil Cultivar CI DI

2014 1 1x107 ACSN39 12.5 8.3 2014 1 1x107 ACSN39 42.9 19.0 2014 1 1x107 ACSN39 57.1 28.6 2014 1 1x107 ACSN39 28.6 9.5 2014 1 0 45H29 0.0 0.0 2014 1 0 45H29 0.0 0.0 2014 1 0 45H29 0.0 0.0 2014 1 0 45H29 0.0 0.0 2014 1 1x103 45H29 0.0 0.0 2014 1 1x103 45H29 0.0 0.0 2014 1 1x103 45H29 0.0 0.0 2014 1 1x103 45H29 0.0 0.0 2014 1 1x104 45H29 0.0 0.0 2014 1 1x104 45H29 0.0 0.0 2014 1 1x104 45H29 0.0 0.0 2014 1 1x104 45H29 0.0 0.0 2014 1 1x105 45H29 0.0 0.0 2014 1 1x105 45H29 0.0 0.0 2014 1 1x105 45H29 0.0 0.0 2014 1 1x105 45H29 0.0 0.0 2014 1 1x106 45H29 0.0 0.0 2014 1 1x106 45H29 0.0 0.0 2014 1 1x106 45H29 0.0 0.0 2014 1 1x106 45H29 0.0 0.0 2014 1 1x107 45H29 0.0 0.0 2014 1 1x107 45H29 0.0 0.0 2014 1 1x107 45H29 0.0 0.0 2014 1 1x107 45H29 0.0 0.0 2015 1 1x107 ACSN39 100.0 83.3 2015 1 1x107 ACSN39 100.0 80.0 2015 1 1x107 ACSN39 100.0 80.0 2015 1 1x107 ACSN39 90.0 70.0 2015 1 0 45H29 0.0 0.0 2015 1 0 45H29 0.0 0.0 2015 1 0 45H29 0.0 0.0 2015 1 0 45H29 0.0 0.0 2015 1 1x103 45H29 0.0 0.0 2015 1 1x103 45H29 0.0 0.0 2015 1 1x103 45H29 0.0 0.0 2015 1 1x103 45H29 0.0 0.0

Page 142: Effect of Plasmodiophora brassicae ... - University of Guelph

  135  2015 1 1x104 45H29 0.0 0.0 2015 1 1x104 45H29 0.0 0.0 2015 1 1x104 45H29 0.0 0.0 2015 1 1x104 45H29 0.0 0.0 2015 1 1x105 45H29 0.0 0.0 2015 1 1x105 45H29 0.0 0.0 2015 1 1x105 45H29 0.0 0.0 2015 1 1x105 45H29 0.0 0.0 2015 1 1x106 45H29 0.0 0.0 2015 1 1x106 45H29 0.0 0.0 2015 1 1x106 45H29 0.0 0.0 2015 1 1x106 45H29 0.0 0.0 2015 1 1x107 45H29 10.0 10.0 2015 1 1x107 45H29 20.0 13.3 2015 1 1x107 45H29 0.0 0.0 2015 1 1x107 45H29 0.0 0.0 2015 2 1x107 ACSN39 100.0 96.7 2015 2 1x107 ACSN39 100.0 93.3 2015 2 1x107 ACSN39 100.0 100.0 2015 2 1x107 ACSN39 100.0 96.7 2015 2 0 45H29 0.0 0.0 2015 2 0 45H29 0.0 0.0 2015 2 0 45H29 0.0 0.0 2015 2 0 45H29 0.0 0.0 2015 2 1x103 45H29 0.0 0.0 2015 2 1x103 45H29 0.0 0.0 2015 2 1x103 45H29 0.0 0.0 2015 2 1x103 45H29 10.0 3.3 2015 2 1x104 45H29 10.0 6.7 2015 2 1x104 45H29 0.0 0.0 2015 2 1x104 45H29 0.0 0.0 2015 2 1x104 45H29 0.0 0.0 2015 2 1x105 45H29 0.0 0.0 2015 2 1x105 45H29 10.0 6.7 2015 2 1x105 45H29 0.0 0.0 2015 2 1x105 45H29 0.0 0.0 2015 2 1x106 45H29 0.0 0.0 2015 2 1x106 45H29 0.0 0.0 2015 2 1x106 45H29 0.0 0.0 2015 2 1x106 45H29 20.0 10.0 2015 2 1x107 45H29 10.0 3.3 2015 2 1x107 45H29 10.0 10.0 2015 2 1x107 45H29 20.0 20.0

Page 143: Effect of Plasmodiophora brassicae ... - University of Guelph

  136  2015 2 1x107 45H29 20.0 13.3

Table A1.64. Raw data for large pot studies – outdoor, 2014 and 2015: plant height of resistant canola

Year Repetition Spores g-1 soil Cultivar

2 WAI

3 WAI

4 WAI

5 WAI

6 WAI

7 WAI

2014 1 0 45H29 1.65 2.53 4.09 7.24 10.08 . 2014 1 0 45H29 1.91 3.00 4.38 8.60 11.69 . 2014 1 0 45H29 1.72 2.35 3.05 5.77 7.40 . 2014 1 0 45H29 1.28 2.32 3.08 4.66 5.80 . 2014 1 1x103 45H29 1.45 2.22 3.61 6.81 8.45 . 2014 1 1x103 45H29 1.87 3.05 4.67 10.72 14.06 . 2014 1 1x103 45H29 1.88 2.79 3.74 6.80 9.12 . 2014 1 1x103 45H29 0.78 1.66 2.82 4.78 6.90 . 2014 1 1x104 45H29 1.32 2.03 2.84 6.36 8.10 . 2014 1 1x104 45H29 1.63 2.77 4.28 8.13 10.31 . 2014 1 1x104 45H29 1.89 2.58 3.47 7.03 9.73 . 2014 1 1x104 45H29 0.98 2.26 2.96 5.02 6.48 . 2014 1 1x105 45H29 2.10 3.15 5.69 12.74 14.76 . 2014 1 1x105 45H29 1.29 2.22 3.30 7.36 10.14 . 2014 1 1x105 45H29 1.50 2.31 3.59 8.09 11.58 . 2014 1 1x105 45H29 1.36 2.12 3.40 5.74 7.04 . 2014 1 1x106 45H29 1.39 2.30 4.01 9.93 13.76 . 2014 1 1x106 45H29 1.81 2.74 4.44 8.87 16.05 . 2014 1 1x106 45H29 1.40 2.10 2.74 4.99 6.36 . 2014 1 1x106 45H29 1.12 1.82 2.54 4.54 6.14 . 2014 1 1x107 45H29 1.48 2.16 3.38 7.08 9.25 . 2014 1 1x107 45H29 1.89 2.76 3.67 8.15 10.75 . 2014 1 1x107 45H29 1.09 1.61 2.00 3.61 4.55 . 2014 1 1x107 45H29 1.22 2.38 2.98 4.80 6.10 . 2015 1 0 45H29 5.40 11.36 25.37 49.89 57.06 60.60 2015 1 0 45H29 7.19 17.63 38.70 50.17 50.71 57.15 2015 1 0 45H29 8.75 23.41 37.71 50.77 50.22 53.85 2015 1 0 45H29 6.06 17.90 30.11 47.65 49.51 52.90 2015 1 1x103 45H29 10.86 15.81 33.97 48.86 50.96 55.90 2015 1 1x103 45H29 6.62 18.98 36.07 50.56 52.21 57.64 2015 1 1x103 45H29 8.87 23.14 33.15 38.38 46.55 47.87 2015 1 1x103 45H29 6.14 21.14 35.13 48.27 48.40 52.24 2015 1 1x104 45H29 6.31 10.40 32.01 43.70 47.00 53.86 2015 1 1x104 45H29 6.08 14.53 32.15 43.98 45.44 53.99 2015 1 1x104 45H29 5.49 14.38 38.00 55.59 55.91 59.83

Page 144: Effect of Plasmodiophora brassicae ... - University of Guelph

  137  2015 1 1x104 45H29 7.79 22.94 33.20 44.21 47.97 55.34 2015 1 1x105 45H29 5.97 17.92 38.33 44.08 51.67 55.22 2015 1 1x105 45H29 6.47 16.48 33.39 52.72 51.37 58.87 2015 1 1x105 45H29 8.95 24.56 39.15 48.44 48.00 52.14 2015 1 1x105 45H29 5.13 13.58 29.94 48.63 50.65 56.86 2015 1 1x106 45H29 5.98 14.74 29.20 42.21 48.17 50.47 2015 1 1x106 45H29 8.72 24.62 33.84 46.04 46.41 55.87 2015 1 1x106 45H29 6.13 13.30 34.80 49.46 56.83 57.52 2015 1 1x106 45H29 6.27 16.79 28.70 47.22 47.68 54.48 2015 1 1x107 45H29 4.98 9.34 29.44 38.90 45.39 48.30 2015 1 1x107 45H29 5.35 9.94 23.26 44.62 44.94 53.71 2015 1 1x107 45H29 5.73 16.77 32.50 51.84 53.65 59.70 2015 1 1x107 45H29 4.53 15.34 34.52 43.83 44.27 52.86 2015 2 0 45H29 2.85 7.89 18.40 34.78 66.24 69.74 2015 2 0 45H29 2.82 12.19 18.17 52.13 77.23 78.14 2015 2 0 45H29 2.39 12.91 21.71 57.88 80.43 78.42 2015 2 0 45H29 2.88 12.27 20.29 46.14 70.07 76.03 2015 2 1x103 45H29 3.42 11.68 22.81 44.44 67.69 74.02 2015 2 1x103 45H29 2.68 5.37 13.17 35.57 62.31 70.77 2015 2 1x103 45H29 3.13 14.20 22.05 52.95 72.11 71.93 2015 2 1x103 45H29 2.79 12.70 20.33 51.46 69.23 72.03 2015 2 1x104 45H29 3.48 11.49 20.81 50.59 72.27 77.91 2015 2 1x104 45H29 3.17 9.29 20.39 45.18 68.54 67.97 2015 2 1x104 45H29 3.30 11.83 23.64 63.42 81.54 83.79 2015 2 1x104 45H29 2.68 11.29 23.02 57.30 78.48 81.29 2015 2 1x105 45H29 3.20 11.77 21.21 39.72 63.12 65.57 2015 2 1x105 45H29 2.22 10.89 20.49 49.48 72.79 75.58 2015 2 1x105 45H29 2.29 11.30 23.31 54.26 76.51 76.87 2015 2 1x105 45H29 2.50 10.57 22.90 49.64 66.28 68.69 2015 2 1x106 45H29 2.71 9.67 20.02 36.21 54.20 59.94 2015 2 1x106 45H29 3.51 9.67 19.84 50.46 78.14 76.04 2015 2 1x106 45H29 2.13 7.59 13.70 39.53 65.97 66.14 2015 2 1x106 45H29 3.02 10.95 18.86 47.21 64.54 65.00 2015 2 1x107 45H29 3.57 11.68 20.49 41.05 59.60 69.34 2015 2 1x107 45H29 3.08 10.19 21.87 49.73 81.31 80.07 2015 2 1x107 45H29 3.38 17.68 26.52 58.59 73.63 77.27 2015 2 1x107 45H29 2.33 9.66 20.71 47.38 65.74 71.54

Table A1.65. Raw data for large pot studies – outdoor, 2014 and 2015: biomass (dry shoot weight) of resistant canola Year Repetition Resting spores g^1 soil Block Mean dry shoot

Page 145: Effect of Plasmodiophora brassicae ... - University of Guelph

  138  weight of 10 plants (g)

2014 1 0 1 275.71 2014 1 0 2 316.00 2014 1 0 3 175.50 2014 1 0 4 186.50 2014 1 1x103 1 152.89 2014 1 1x103 2 248.11 2014 1 1x103 3 146.50 2014 1 1x103 4 28.50 2014 1 1x104 1 250.67 2014 1 1x104 2 108.67 2014 1 1x104 3 119.50 2014 1 1x104 4 277.40 2014 1 1x105 1 210.20 2014 1 1x105 2 210.50 2014 1 1x105 3 162.71 2014 1 1x105 4 199.60 2014 1 1x106 1 88.60 2014 1 1x106 2 80.67 2014 1 1x106 3 251.40 2014 1 1x106 4 123.25 2014 1 1x107 1 104.75 2014 1 1x107 2 182.67 2014 1 1x107 3 101.67 2014 1 1x107 4 42.00 2015 1 0 1 654.00 2015 1 0 2 702.20 2015 1 0 3 672.20 2015 1 0 4 771.80 2015 1 1x103 1 768.20 2015 1 1x103 2 694.40 2015 1 1x103 3 816.80 2015 1 1x103 4 660.00 2015 1 1x104 1 683.20 2015 1 1x104 2 717.60 2015 1 1x104 3 715.40 2015 1 1x104 4 764.20 2015 1 1x105 1 734.60 2015 1 1x105 2 755.20 2015 1 1x105 3 622.60 2015 1 1x105 4 665.60 2015 1 1x106 1 644.20 2015 1 1x106 2 727.80

Page 146: Effect of Plasmodiophora brassicae ... - University of Guelph

  139  2015 1 1x106 3 687.20 2015 1 1x106 4 824.80 2015 1 1x107 1 1023.80 2015 1 1x107 2 738.60 2015 1 1x107 3 701.00 2015 1 1x107 4 619.40 2015 2 0 1 286.20 2015 2 0 2 193.60 2015 2 0 3 225.60 2015 2 0 4 215.60 2015 2 1x103 1 215.80 2015 2 1x103 2 128.00 2015 2 1x103 3 305.20 2015 2 1x103 4 262.20 2015 2 1x104 1 242.00 2015 2 1x104 2 293.20 2015 2 1x104 3 380.75 2015 2 1x104 4 338.80 2015 2 1x105 1 302.20 2015 2 1x105 2 261.00 2015 2 1x105 3 251.20 2015 2 1x105 4 354.40 2015 2 1x106 1 263.20 2015 2 1x106 2 200.00 2015 2 1x106 3 245.20 2015 2 1x106 4 257.40 2015 2 1x107 1 251.50 2015 2 1x107 2 229.50 2015 2 1x107 3 418.67 2015 2 1x107 4 218.20 Table A1.66. Raw data for large pot studies – outdoor, 2014 and 2015: maturity (% plants at flowering and seedpod development stages) of resistant canola

Year Repetition Resting spores g-1 soil Block % plants at flowering stage

% plants at pod stage

2014 1 0 1 50.0 0.0 2014 1 0 2 57.1 0.0 2014 1 0 3 0.0 0.0 2014 1 0 4 25.0 0.0

Page 147: Effect of Plasmodiophora brassicae ... - University of Guelph

  140  2014 1 1x103 1 30.0 0.0 2014 1 1x103 2 44.4 0.0 2014 1 1x103 3 25.0 0.0 2014 1 1x103 4 0.0 0.0 2014 1 1x104 1 14.3 0.0 2014 1 1x104 2 44.4 0.0 2014 1 1x104 3 25.0 0.0 2014 1 1x104 4 25.0 0.0 2014 1 1x105 1 88.9 0.0 2014 1 1x105 2 42.9 0.0 2014 1 1x105 3 42.9 0.0 2014 1 1x105 4 40.0 0.0 2014 1 1x106 1 60.0 0.0 2014 1 1x106 2 66.7 0.0 2014 1 1x106 3 20.0 0.0 2014 1 1x106 4 20.0 0.0 2014 1 1x107 1 25.0 0.0 2014 1 1x107 2 80.0 0.0 2014 1 1x107 3 16.7 0.0 2014 1 1x107 4 0.0 0.0 2015 1 0 1 0.0 100.0 2015 1 0 2 0.0 100.0 2015 1 0 3 0.0 100.0 2015 1 0 4 0.0 100.0 2015 1 1x103 1 0.0 100.0 2015 1 1x103 2 0.0 100.0 2015 1 1x103 3 0.0 100.0 2015 1 1x103 4 0.0 100.0 2015 1 1x104 1 0.0 100.0 2015 1 1x104 2 0.0 100.0 2015 1 1x104 3 0.0 100.0 2015 1 1x104 4 0.0 100.0 2015 1 1x105 1 0.0 100.0 2015 1 1x105 2 0.0 100.0 2015 1 1x105 3 0.0 100.0 2015 1 1x105 4 0.0 100.0 2015 1 1x106 1 0.0 100.0 2015 1 1x106 2 0.0 100.0 2015 1 1x106 3 0.0 100.0 2015 1 1x106 4 0.0 100.0 2015 1 1x107 1 0.0 100.0 2015 1 1x107 2 0.0 100.0 2015 1 1x107 3 0.0 100.0

Page 148: Effect of Plasmodiophora brassicae ... - University of Guelph

  141  2015 1 1x107 4 0.0 100.0 2015 2 0 1 0.0 100.0 2015 2 0 2 0.0 100.0 2015 2 0 3 0.0 100.0 2015 2 0 4 0.0 100.0 2015 2 1x103 1 0.0 100.0 2015 2 1x103 2 0.0 100.0 2015 2 1x103 3 0.0 100.0 2015 2 1x103 4 0.0 100.0 2015 2 1x104 1 0.0 100.0 2015 2 1x104 2 0.0 100.0 2015 2 1x104 3 0.0 100.0 2015 2 1x104 4 0.0 100.0 2015 2 1x105 1 0.0 100.0 2015 2 1x105 2 0.0 100.0 2015 2 1x105 3 0.0 100.0 2015 2 1x105 4 0.0 100.0 2015 2 1x106 1 0.0 100.0 2015 2 1x106 2 0.0 100.0 2015 2 1x106 3 0.0 100.0 2015 2 1x106 4 0.0 100.0 2015 2 1x107 1 0.0 100.0 2015 2 1x107 2 0.0 100.0 2015 2 1x107 3 0.0 100.0 2015 2 1x107 4 0.0 100.0

Table A1.67. Raw data for controlled environment study – canola: CI & DSI

Repetition Spores ml-1 Cultivar Block CI DSI

1 1x106 46A76 1 33.3 11.1 1 1x106 46A76 2 0.0 0.0 1 1x106 46A76 3 22.2 7.4 1 1x106 46A76 4 14.3 4.8 1 0 46A76 1 0.0 0.0 1 0 46A76 2 0.0 0.0 1 0 46A76 3 0.0 0.0 1 0 46A76 4 0.0 0.0 1 1x106 45H29 1 0.0 0.0 1 1x106 45H29 2 0.0 0.0 1 1x106 45H29 3 0.0 0.0 1 1x106 45H29 4 0.0 0.0 1 0 45H29 1 0.0 0.0 1 0 45H29 2 0.0 0.0

Page 149: Effect of Plasmodiophora brassicae ... - University of Guelph

  142  1 0 45H29 3 0.0 0.0 1 0 45H29 4 0.0 0.0 1 1x106 7367 1 0.0 0.0 1 1x106 7367 2 0.0 0.0 1 1x106 7367 3 0.0 0.0 1 1x106 7367 4 0.0 0.0 1 0 7367 1 0.0 0.0 1 0 7367 2 0.0 0.0 1 0 7367 3 0.0 0.0 1 0 7367 4 0.0 0.0 1 1x106 7377 1 0.0 0.0 1 1x106 7377 2 0.0 0.0 1 1x106 7377 3 0.0 0.0 1 1x106 7377 4 0.0 0.0 1 0 7377 1 0.0 0.0 1 0 7377 2 0.0 0.0 1 0 7377 3 0.0 0.0 1 0 7377 4 0.0 0.0 2 1x106 ACSN39 1 66.7 40.7 2 1x106 ACSN39 2 50.0 25.0 2 1x106 ACSN39 3 44.4 25.9 2 1x106 ACSN39 4 88.9 37.0 2 0 ACSN39 1 0.0 0.0 2 0 ACSN39 2 0.0 0.0 2 0 ACSN39 3 0.0 0.0 2 0 ACSN39 4 0.0 0.0 2 1x106 45H29 1 0.0 0.0 2 1x106 45H29 2 0.0 0.0 2 1x106 45H29 3 0.0 0.0 2 1x106 45H29 4 0.0 0.0 2 0 45H29 1 0.0 0.0 2 0 45H29 2 0.0 0.0 2 0 45H29 3 0.0 0.0 2 0 45H29 4 0.0 0.0 2 1x106 7367 1 0.0 0.0 2 1x106 7367 2 0.0 0.0 2 1x106 7367 3 0.0 0.0 2 1x106 7367 4 0.0 0.0 2 0 7367 1 0.0 0.0 2 0 7367 2 0.0 0.0 2 0 7367 3 0.0 0.0 2 0 7367 4 0.0 0.0 2 1x106 7377 1 0.0 0.0

Page 150: Effect of Plasmodiophora brassicae ... - University of Guelph

  143  2 1x106 7377 2 0.0 0.0 2 1x106 7377 3 0.0 0.0 2 1x106 7377 4 0.0 0.0 2 0 7377 1 0.0 0.0 2 0 7377 2 0.0 0.0 2 0 7377 3 0.0 0.0 2 0 7377 4 0.0 0.0

Page 151: Effect of Plasmodiophora brassicae ... - University of Guelph

  144   Table A1.68 Raw data for controlled environment study – canola: plant height  Repetition Spores ml-1 Cultivar Block 2 WAI 3 WAI 4 WAI 5 WAI 6 WAI 1 1x106 46A76 1 2.80 3.54 3.97 4.72 5.45 1 1x106 46A76 2 3.33 . 3.52 4.29 4.59 1 1x106 46A76 3 2.79 3.53 3.85 4.98 6.28 1 1x106 46A76 4 3.12 . 3.70 4.14 4.50 1 1x106 45H29 1 2.38 3.18 3.58 4.63 5.93 1 1x106 45H29 2 2.77 3.19 3.93 4.93 6.16 1 1x106 45H29 3 3.57 3.87 5.17 6.66 9.04 1 1x106 45H29 4 2.76 3.03 3.70 5.00 6.16 1 1x106 7377 1 2.35 2.99 3.05 3.61 4.25 1 1x106 7377 2 2.88 3.29 3.41 3.99 4.92 1 1x106 7377 3 1.56 1.78 2.60 3.20 3.60 1 1x106 7377 4 2.44 3.45 3.83 4.55 4.80 1 1x106 7367 1 2.75 3.32 3.47 4.29 5.26 1 1x106 7367 2 2.37 2.59 3.06 3.23 3.58 1 1x106 7367 3 2.41 3.07 3.04 3.53 4.08 1 1x106 7367 4 2.31 2.76 2.71 3.01 4.05 1 0 46A76 1 3.00 3.30 3.40 3.60 4.10 1 0 46A76 2 2.93 3.47 3.45 3.90 4.42 1 0 46A76 3 3.34 3.69 3.77 4.68 5.53 1 0 46A76 4 2.30 2.88 3.05 4.55 6.05 1 0 45H29 1 2.68 3.28 3.89 5.33 7.16 1 0 45H29 2 3.43 4.00 4.34 7.20 9.89 1 0 45H29 3 2.67 2.71 4.03 6.38 9.55 1 0 45H29 4 2.92 3.36 3.86 4.56 5.68 1 0 7377 1 3.26 3.74 3.90 5.04 5.63 1 0 7377 2 3.28 3.36 3.68 4.16 4.62 1 0 7377 3 3.08 3.38 3.99 4.44 5.20 1 0 7377 4 2.34 2.96 3.45 4.00 4.83 1 0 7367 1 2.48 2.96 3.10 3.40 4.00 1 0 7367 2 3.02 3.41 3.80 3.92 4.80 1 0 7367 3 2.75 3.30 3.20 3.61 4.19 1 0 7367 4 3.08 3.64 4.13 4.65 5.37 2 1x106 ACSN39 1 2.68 3.12 4.10 9.42 14.58 2 1x106 ACSN39 2 2.74 3.14 4.54 13.28 16.81 2 1x106 ACSN39 3 2.70 3.34 4.71 12.61 19.20 2 1x106 ACSN39 4 3.29 3.73 5.46 15.26 21.41 2 1x106 45H29 1 3.12 4.57 8.55 18.78 22.17 2 1x106 45H29 2 2.82 3.64 5.74 15.13 21.84 2 1x106 45H29 3 3.17 4.59 8.90 23.63 26.72 2 1x106 45H29 4 3.89 5.21 10.68 20.15 28.96

Page 152: Effect of Plasmodiophora brassicae ... - University of Guelph

  145  2 1x106 7377 1 2.86 3.49 4.96 13.21 20.33 2 1x106 7377 2 3.04 3.39 5.15 14.57 21.82 2 1x106 7377 3 2.74 3.16 4.00 10.13 14.53 2 1x106 7377 4 3.73 4.26 5.73 11.77 16.89 2 1x106 7367 1 3.03 3.38 4.21 8.06 11.03 2 1x106 7367 2 3.13 3.69 5.02 14.21 22.05 2 1x106 7367 3 3.35 3.68 4.90 11.73 15.94 2 1x106 7367 4 4.18 5.61 8.81 17.78 21.44 2 0 ACSN39 1 2.64 3.51 5.44 14.14 21.79 2 0 ACSN39 2 2.96 3.26 4.10 9.69 15.41 2 0 ACSN39 3 2.90 3.18 4.76 14.20 19.23 2 0 ACSN39 4 3.60 4.10 5.63 13.45 21.48 2 0 45H29 1 3.13 4.58 7.59 20.62 25.22 2 0 45H29 2 2.80 4.01 9.60 25.02 27.96 2 0 45H29 3 3.24 4.26 7.01 15.51 20.94 2 0 45H29 4 3.91 5.51 10.26 27.09 30.57 2 0 7377 1 2.84 3.54 4.92 12.02 17.41 2 0 7377 2 2.84 3.42 5.28 13.76 19.09 2 0 7377 3 2.98 3.56 5.34 13.83 17.19 2 0 7377 4 2.91 3.53 4.28 9.03 12.39 2 0 7367 1 3.54 4.10 6.26 16.47 21.75 2 0 7367 2 3.09 3.79 5.05 12.05 18.31 2 0 7367 3 3.60 4.07 5.77 14.08 20.37 2 0 7367 4 4.25 5.01 6.78 14.33 18.24

Table A1.69. Raw data for controlled environment study – canola: biomass (dry shoot weight)

Repetition Cultivar Resting spores ml-1 Block Mean dry shoot weight of 10 plants (g)

1 46A76 1x106 1 3.17 1 46A76 1x106 2 3.53 1 46A76 1x106 3 3.36 1 46A76 1x106 4 2.91 1 46A76 0 1 2.52 1 46A76 0 2 3.28 1 46A76 0 3 3.13 1 46A76 0 4 3.15 1 45H29 1x106 1 2.69 1 45H29 1x106 2 3.11 1 45H29 1x106 3 2.76 1 45H29 1x106 4 0.75 1 45H29 0 1 2.50 1 45H29 0 2 2.80

Page 153: Effect of Plasmodiophora brassicae ... - University of Guelph

  146  1 45H29 0 3 2.94 1 45H29 0 4 3.61 1 7367 1x106 1 3.07 1 7367 1x106 2 2.83 1 7367 1x106 3 2.10 1 7367 1x106 4 2.55 1 7367 0 1 3.58 1 7367 0 2 2.21 1 7367 0 3 2.47 1 7367 0 4 2.53 1 7377 1x106 1 2.61 1 7377 1x106 2 3.21 1 7377 1x106 3 2.75 1 7377 1x106 4 2.78 1 7377 0 1 2.40 1 7377 0 2 2.08 1 7377 0 3 2.31 1 7377 0 4 2.55 2 ACSN39 1x106 1 15.54 2 ACSN39 1x106 2 10.89 2 ACSN39 1x106 3 11.91 2 ACSN39 1x106 4 12.78 2 ACSN39 0 1 15.18 2 ACSN39 0 2 13.26 2 ACSN39 0 3 13.39 2 ACSN39 0 4 12.95 2 45H29 1x106 1 14.82 2 45H29 1x106 2 14.40 2 45H29 1x106 3 13.97 2 45H29 1x106 4 13.88 2 45H29 0 1 15.70 2 45H29 0 2 13.46 2 45H29 0 3 10.36 2 45H29 0 4 15.16 2 7367 1x106 1 14.16 2 7367 1x106 2 12.28 2 7367 1x106 3 11.13 2 7367 1x106 4 14.39 2 7367 0 4 9.70 2 7367 0 1 11.36 2 7367 0 2 12.80 2 7367 0 3 11.81 2 7377 1x106 1 15.56

Page 154: Effect of Plasmodiophora brassicae ... - University of Guelph

  147  2 7377 1x106 2 13.15 2 7377 1x106 3 11.88 2 7377 1x106 4 11.83 2 7377 0 1 16.12 2 7377 0 2 14.27 2 7377 0 3 11.23 2 7377 0 4 10.98

Table A1.70. Raw data for controlled environment study – canola: maturity: % plants at flowering and seedpod development stages

Repetition Cultivar Resting spores g^1 soil Block

% plants at flowering stage

% plants at pod stage

1 45H29 1x106 1 25.0 0.0 1 45H29 1x106 2 0.0 0.0 1 45H29 1x106 3 14.3 14.3 1 45H29 1x106 4 0.0 0.0 1 45H29 0 1 0.0 0.0 1 45H29 0 2 12.5 25.0 1 45H29 0 3 0.0 25.0 1 45H29 0 4 0.0 0.0 2 ACSN39 1x106 1 100.0 0.0 2 ACSN39 1x106 2 50.0 37.5 2 ACSN39 1x106 3 90.0 0.0 2 ACSN39 1x106 4 77.8 22.2 2 ACSN39 0 1 70.0 30.0 2 ACSN39 0 2 80.0 10.0 2 ACSN39 0 3 70.0 30.0 2 ACSN39 0 4 37.5 62.5 2 45H29 1x106 1 50.0 50.0 2 45H29 1x106 2 50.0 50.0 2 45H29 1x106 3 30.0 70.0 2 45H29 1x106 4 22.2 77.8 2 45H29 0 1 30.0 70.0 2 45H29 0 2 30.0 70.0 2 45H29 0 3 77.8 22.2 2 45H29 0 4 40.0 60.0 2 7367 1x106 1 90.0 0.0 2 7367 1x106 2 90.0 10.0 2 7367 1x106 3 80.0 10.0 2 7367 1x106 4 30.0 70.0 2 7367 0 1 80.0 20.0 2 7367 0 2 90.0 0.0 2 7367 0 3 90.0 0.0

Page 155: Effect of Plasmodiophora brassicae ... - University of Guelph

  148  2 7367 0 4 40.0 60.0 2 7377 1x106 1 90.0 10.0 2 7377 1x106 2 80.0 0.0 2 7377 1x106 3 80.0 0.0 2 7377 1x106 4 100.0 0.0 2 7377 0 1 100.0 0.0 2 7377 0 2 88.9 11.1 2 7377 0 3 90.0 0.0 2 7377 0 4 66.7 0.0

Table A1.71. Raw data for controlled environment study – napa cabbage: leaf length

Repetition Cultivar Spores ml-1 Block

2 WAI

3 WAI

4 WAI

5 WAI

6 WAI

1 ChinaGold 1x106 1 2.49 2.68 2.75 3.28 3.7 1 ChinaGold 1x106 2 2.30 2.47 2.83 3.10 3.78 1 ChinaGold 1x106 3 2.83 2.82 2.83 3.29 3.77 1 ChinaGold 1x106 4 2.51 2.89 2.76 3.00 3.43 1 ChinaGold 0 1 2.65 2.93 2.92 3.70 3.97 1 ChinaGold 0 2 2.67 2.76 3.02 3.20 3.84 1 ChinaGold 0 3 2.78 2.91 2.78 3.48 3.94 1 ChinaGold 0 4 2.43 2.54 2.95 3.27 3.50 1 Emiko 1x106 1 2.33 2.40 2.62 3.03 3.39 1 Emiko 1x106 2 2.53 2.56 2.91 2.73 3.02 1 Emiko 1x106 3 2.41 2.82 2.82 3.22 3.26 1 Emiko 1x106 4 2.36 2.63 2.80 2.90 3.13 1 Emiko 0 1 2.40 2.50 2.85 3.49 3.77 1 Emiko 0 2 2.39 2.51 2.68 2.51 2.83 1 Emiko 0 3 2.23 2.63 2.91 3.26 3.43 1 Emiko 0 4 2.44 3.54 2.70 2.99 3.05 1 Mirako 1x106 1 2.77 2.89 2.92 3.65 4.39 1 Mirako 1x106 2 2.80 2.91 3.16 3.38 3.77 1 Mirako 1x106 3 2.63 2.98 3.14 3.44 3.65 1 Mirako 1x106 4 2.53 2.94 2.97 3.27 3.51 1 Mirako 0 1 2.79 2.85 3.15 3.59 3.96 1 Mirako 0 2 2.72 2.93 3.21 3.41 3.81 1 Mirako 0 3 2.70 3.14 3.37 3.58 3.95 1 Mirako 0 4 2.71 3.17 3.13 3.71 3.78 1 Yuki 1x106 1 2.06 2.09 2.60 2.73 3.57 1 Yuki 1x106 2 2.08 2.41 2.68 3.03 3.53 1 Yuki 1x106 3 1.90 2.33 2.73 3.04 3.40 1 Yuki 1x106 4 2.30 2.67 2.79 2.99 3.27 1 Yuki 0 1 2.05 2.36 2.48 3.07 3.37 1 Yuki 0 2 2.36 2.65 2.97 3.32 3.71

Page 156: Effect of Plasmodiophora brassicae ... - University of Guelph

  149  1 Yuki 0 3 2.12 2.62 2.88 2.15 3.19 1 Yuki 0 4 2.13 2.70 2.83 3.18 3.42 2 ChinaGold 1x106 1 5.25 5.55 7.45 7.75 7.76 2 ChinaGold 1x106 2 5.16 5.30 6.01 6.92 6.01 2 ChinaGold 1x106 3 4.43 4.60 6.64 7.48 7.18 2 ChinaGold 1x106 4 4.53 5.26 6.06 5.96 5.75 2 ChinaGold 0 1 5.80 5.85 8.05 10.23 9.05 2 ChinaGold 0 2 4.94 5.74 6.82 7.71 7.37 2 ChinaGold 0 3 4.66 4.92 6.99 7.22 6.12 2 ChinaGold 0 4 4.17 4.26 6.43 7.52 6.55 2 Emiko 1x106 1 4.40 4.76 5.76 6.59 6.61 2 Emiko 1x106 2 4.03 4.31 5.73 6.09 6.08 2 Emiko 1x106 3 4.53 4.76 6.46 6.07 6.69 2 Emiko 1x106 4 3.61 3.87 4.64 5.07 5.18 2 Emiko 0 1 4.92 5.01 5.95 11.22 6.83 2 Emiko 0 2 4.22 4.66 5.73 6.97 6.49 2 Emiko 0 3 4.55 3.63 4.85 6.81 6.20 2 Emiko 0 4 3.34 3.63 5.08 5.78 5.81 2 Mirako 1x106 1 4.22 4.54 5.89 7.22 6.97 2 Mirako 1x106 2 4.77 4.84 5.93 6.58 5.70 2 Mirako 1x106 3 4.04 4.09 6.35 7.23 7.42 2 Mirako 1x106 4 4.10 4.09 5.35 6.23 5.70 2 Mirako 0 1 5.47 5.53 6.31 7.28 7.13 2 Mirako 0 2 4.94 5.20 7.17 7.69 7.13 2 Mirako 0 3 4.89 4.71 6.50 6.59 6.88 2 Mirako 0 4 5.02 4.71 5.56 6.46 6.37 2 Yuki 1x106 1 3.13 4.71 6.64 7.02 5.90 2 Yuki 1x106 2 3.80 4.03 10.31 6.23 5.90 2 Yuki 1x106 3 3.49 5.26 6.03 6.01 5.38 2 Yuki 1x106 4 3.45 4.22 5.16 5.19 5.05 2 Yuki 0 1 3.97 5.07 6.70 6.92 6.45 2 Yuki 0 2 4.55 5.21 7.00 7.06 6.45 2 Yuki 0 3 4.03 3.64 6.12 5.74 4.73 2 Yuki 0 4 3.93 4.33 6.30 6.37 5.70

Table A1.72. Raw data for controlled environment study – napa cabbage: biomass (dry shoot weight)

Repetition Cultivar Spores g-1 soil Block

Mean dry shoot weight (g)

1 ChinaGold 1x106 1 4.54 1 ChinaGold 1x106 2 2.88 1 ChinaGold 1x106 3 3.94 1 ChinaGold 1x106 4 2.49

Page 157: Effect of Plasmodiophora brassicae ... - University of Guelph

  150  1 ChinaGold 0 1 4.10 1 ChinaGold 0 2 2.90 1 ChinaGold 0 3 2.99 1 ChinaGold 0 4 3.10 1 Emiko 1x106 1 3.31 1 Emiko 1x106 2 4.11 1 Emiko 1x106 3 2.31 1 Emiko 1x106 4 1.99 1 Emiko 0 1 5.39 1 Emiko 0 2 3.32 1 Emiko 0 3 2.96 1 Emiko 0 4 2.16 1 Mirako 1x106 1 5.17 1 Mirako 1x106 2 4.46 1 Mirako 1x106 3 3.04 1 Mirako 1x106 4 2.67 1 Mirako 0 1 5.19 1 Mirako 0 2 5.42 1 Mirako 0 3 3.79 1 Mirako 0 4 3.08 1 Yuki 1x106 1 5.51 1 Yuki 1x106 2 5.05 1 Yuki 1x106 3 4.81 1 Yuki 1x106 4 4.47 1 Yuki 0 1 5.40 1 Yuki 0 2 4.08 1 Yuki 0 3 3.93 1 Yuki 0 4 3.96 2 ChinaGold 1x106 1 12.26 2 ChinaGold 1x106 2 10.13 2 ChinaGold 1x106 3 10.79 2 ChinaGold 1x106 4 9.17 2 ChinaGold 0 1 11.29 2 ChinaGold 0 2 11.53 2 ChinaGold 0 3 9.09 2 ChinaGold 0 4 10.83 2 Emiko 1x106 1 19.42 2 Emiko 1x106 2 17.18 2 Emiko 1x106 3 20.38 2 Emiko 1x106 4 11.35 2 Emiko 0 1 32.33 2 Emiko 0 2 17.38 2 Emiko 0 3 18.28

Page 158: Effect of Plasmodiophora brassicae ... - University of Guelph

  151  2 Emiko 0 4 17.57 2 Mirako 1x106 1 10.17 2 Mirako 1x106 2 9.35 2 Mirako 1x106 3 9.97 2 Mirako 1x106 4 9.64 2 Mirako 0 1 14.21 2 Mirako 0 2 18.79 2 Mirako 0 3 11.69 2 Mirako 0 4 13.36 2 Yuki 1x106 1 24.60 2 Yuki 1x106 2 18.72 2 Yuki 1x106 3 17.78 2 Yuki 1x106 4 14.93 2 Yuki 0 1 17.79 2 Yuki 0 2 17.70 2 Yuki 0 3 27.95 2 Yuki 0 4 27.45

Page 159: Effect of Plasmodiophora brassicae ... - University of Guelph

  152  Table A1.73. Raw data for field trials – cabbage: CI & DSI

Repetition Spores g-1 soil Cultivar Block CI DSI

1 1x106 Bronco 1 37.5 12.5 1 1x106 Bronco 2 62.5 29.2 1 1x106 Bronco 3 57.1 19.0 1 0 Bronco 1 0.0 0.0 1 0 Bronco 2 0.0 0.0 1 0 Bronco 3 0.0 0.0 1 1x106 Kilaxy 1 0.0 0.0 1 1x106 Kilaxy 2 0.0 0.0 1 1x106 Kilaxy 3 0.0 0.0 1 0 Kilaxy 1 0.0 0.0 1 0 Kilaxy 2 0.0 0.0 1 0 Kilaxy 3 0.0 0.0 1 1x106 Kilaton 1 0.0 0.0 1 1x106 Kilaton 2 0.0 0.0 1 1x106 Kilaton 3 0.0 0.0 1 0 Kilaton 1 0.0 0.0 1 0 Kilaton 2 0.0 0.0 1 0 Kilaton 3 0.0 0.0 1 1x106 Tekila 1 0.0 0.0 1 1x106 Tekila 2 0.0 0.0 1 1x106 Tekila 3 0.0 0.0 1 0 Tekila 1 0.0 0.0 1 0 Tekila 2 0.0 0.0 1 0 Tekila 3 0.0 0.0 2 1x106 Bronco 1 40.0 20.0 2 1x106 Bronco 2 10.0 3.3 2 1x106 Bronco 3 11.1 3.7 2 1x106 Bronco 4 30.0 13.3 2 0 Bronco 1 0.0 0.0 2 0 Bronco 2 0.0 0.0 2 0 Bronco 3 0.0 0.0 2 0 Bronco 4 0.0 0.0 2 1x106 Kilaherb 1 0.0 0.0 2 1x106 Kilaherb 2 0.0 0.0 2 1x106 Kilaherb 3 0.0 0.0 2 1x106 Kilaherb 4 0.0 0.0 2 0 Kilaherb 1 0.0 0.0 2 0 Kilaherb 2 0.0 0.0 2 0 Kilaherb 3 0.0 0.0 2 0 Kilaherb 4 0.0 0.0

Page 160: Effect of Plasmodiophora brassicae ... - University of Guelph

  153  2 1x106 Kilaton 1 0.0 0.0 2 1x106 Kilaton 2 0.0 0.0 2 1x106 Kilaton 3 0.0 0.0 2 1x106 Kilaton 4 0.0 0.0 2 0 Kilaton 1 0.0 0.0 2 0 Kilaton 2 0.0 0.0 2 0 Kilaton 3 0.0 0.0 2 0 Kilaton 4 0.0 0.0 2 1x106 Tekila 1 0.0 0.0 2 1x106 Tekila 2 0.0 0.0 2 1x106 Tekila 3 0.0 0.0 2 1x106 Tekila 4 0.0 0.0 2 0 Tekila 1 0.0 0.0 2 0 Tekila 2 0.0 0.0 2 0 Tekila 3 0.0 0.0 2 0 Tekila 4 0.0 0.0

Table A1.74. Raw data for controlled environment study – cabbage: leaf length

Run Cultivar Spores ml-1 Block 2 WAI 3 WAI 4 WAI 5 WAI 6 WAI

1 Bronco 1x106 1 2.82 3.09 3.30 3.43 3.49 1 Bronco 1x106 2 2.93 2.79 . 3.31 3.56 1 Bronco 1x106 3 2.98 2.82 3.23 3.31 3.34 1 Bronco 0 1 2.65 2.67 2.96 3.18 3.32 1 Bronco 0 2 2.98 2.97 3.21 3.47 3.63 1 Bronco 0 3 3.18 3.13 3.17 3.43 3.77 1 Kilaxy 1x106 1 2.47 2.65 3.13 3.65 3.04 1 Kilaxy 1x106 2 2.79 2.80 3.09 3.88 4.02 1 Kilaxy 1x106 3 2.94 2.93 2.91 3.10 3.36 1 Kilaxy 0 1 2.78 2.92 3.23 3.72 3.71 1 Kilaxy 0 2 2.76 3.00 . 3.50 3.57 1 Kilaxy 0 3 2.81 2.91 3.19 3.22 3.52 1 Kilaton 1x106 1 1.93 1.09 1.24 3.16 3.35 1 Kilaton 1x106 2 1.98 2.10 . 2.81 3.26 1 Kilaton 1x106 3 2.26 2.55 2.81 3.06 3.19 1 Kilaton 0 1 1.82 1.57 1.50 2.98 3.38 1 Kilaton 0 2 1.98 2.17 2.55 2.76 3.33 1 Kilaton 0 3 2.07 2.33 2.52 2.85 3.01 1 Tekila 1x106 1 2.50 2.57 2.84 3.11 3.13 1 Tekila 1x106 2 2.43 2.56 . 3.03 3.15 1 Tekila 1x106 3 2.62 2.90 3.09 3.62 3.70 1 Tekila 0 1 2.61 2.55 2.76 3.26 3.46

Page 161: Effect of Plasmodiophora brassicae ... - University of Guelph

  154  1 Tekila 0 2 2.63 2.84 3.84 3.13 3.17 1 Tekila 0 3 2.74 2.74 3.03 3.07 3.2 2 Bronco 1x106 1 4.37 4.55 5.06 5.49 5.68 2 Bronco 1x106 2 5.20 5.26 5.40 5.50 5.43 2 Bronco 1x106 3 4.72 4.37 4.35 4.16 4.26 2 Bronco 1x106 4 5.91 6.47 6.61 6.41 6.87 2 Bronco 0 1 6.48 6.35 6.74 7.04 7.24 2 Bronco 0 2 6.15 6.16 6.53 6.50 6.57 2 Bronco 0 3 6.58 6.82 7.19 7.24 6.92 2 Bronco 0 4 4.39 4.45 4.47 4.46 4.40 2 Kilaherb 1x106 1 4.93 5.76 6.55 7.07 7.03 2 Kilaherb 1x106 2 4.43 5.48 6.04 6.20 5.69 2 Kilaherb 1x106 3 4.54 5.13 5.49 5.42 5.06 2 Kilaherb 1x106 4 5.30 7.01 7.84 8.39 7.54 2 Kilaherb 0 1 5.28 6.69 7.02 7.54 7.77 2 Kilaherb 0 2 6.17 7.15 7.46 7.70 7.35 2 Kilaherb 0 3 5.00 6.48 7.02 7.62 7.48 2 Kilaherb 0 4 5.31 6.27 7.13 7.52 7.16 2 Kilaton 1x106 1 4.45 5.15 4.85 5.98 5.72 2 Kilaton 1x106 2 4.23 4.31 4.47 4.70 4.96 2 Kilaton 1x106 3 3.50 4.18 4.38 4.50 4.63 2 Kilaton 1x106 4 . 5.40 6.50 5.90 5.55 2 Kilaton 0 1 4.21 4.64 5.17 5.36 5.26 2 Kilaton 0 2 3.85 5.37 6.57 6.03 5.54 2 Kilaton 0 3 3.68 4.30 4.45 4.67 4.47 2 Kilaton 0 4 4.47 5.73 6.68 6.90 6.75 2 Tekila 1x106 1 4.84 5.37 5.67 6.04 5.86 2 Tekila 1x106 2 5.52 5.58 5.85 5.74 5.60 2 Tekila 1x106 3 5.55 5.74 6.11 6.16 5.85 2 Tekila 1x106 4 5.21 5.39 5.98 5.63 5.69 2 Tekila 0 1 4.34 4.89 5.43 5.21 5.41 2 Tekila 0 2 4.12 4.53 4.73 4.92 5.11 2 Tekila 0 3 5.43 5.74 6.16 6.14 5.92 2 Tekila 0 4 6.19 6.14 6.51 6.30 6.19

Table A1.75. Raw data for controlled environment study – cabbage: biomass (dry shoot weight of 10 plants, g)

Repetition Cultivar Spores ml-1 Block

Dry shoot weight (g)

1 Bronco 1x106 1 8.73 1 Bronco 1x106 2 8.16 1 Bronco 1x106 3 6.91 1 Bronco 0 1 6.88 1 Bronco 0 2 6.96 1 Bronco 0 3 6.74

Page 162: Effect of Plasmodiophora brassicae ... - University of Guelph

  155  1 Kilaxy 1x106 1 5.55 1 Kilaxy 1x106 2 4.80 1 Kilaxy 1x106 3 6.13 1 Kilaxy 0 1 5.80 1 Kilaxy 0 2 4.05 1 Kilaxy 0 3 5.12 1 Kilaton 1x106 1 7.70 1 Kilaton 1x106 2 7.03 1 Kilaton 1x106 3 6.20 1 Kilaton 0 1 7.90 1 Kilaton 0 2 8.04 1 Kilaton 0 3 7.57 1 Tekila 1x106 1 7.57 1 Tekila 1x106 2 8.14 1 Tekila 1x106 3 8.50 1 Tekila 0 1 8.09 1 Tekila 0 2 9.58 1 Tekila 0 3 8.37 2 Bronco 1x106 1 6.45 2 Bronco 1x106 2 10.39 2 Bronco 1x106 3 7.24 2 Bronco 1x106 4 27.94 2 Bronco 0 1 7.94 2 Bronco 0 2 9.59 2 Bronco 0 3 3.97 2 Bronco 0 4 16.78 2 Kilaherb 1x106 1 13.41 2 Kilaherb 1x106 2 13.96 2 Kilaherb 1x106 3 30.99 2 Kilaherb 1x106 4 9.18 2 Kilaherb 0 1 11.39 2 Kilaherb 0 2 16.48 2 Kilaherb 0 3 13.54 2 Kilaherb 0 4 28.68 2 Kilaton 1x106 1 7.10 2 Kilaton 1x106 2 10.93 2 Kilaton 1x106 3 5.71 2 Kilaton 1x106 4 40.00 2 Kilaton 0 1 9.25 2 Kilaton 0 2 10.79 2 Kilaton 0 3 16.42 2 Kilaton 0 4 26.53 2 Tekila 1x106 1 8.51 2 Tekila 1x106 2 9.48 2 Tekila 1x106 3 5.49 2 Tekila 1x106 4 25.10

Page 163: Effect of Plasmodiophora brassicae ... - University of Guelph

  156  2 Tekila 0 1 8.48 2 Tekila 0 2 6.29 2 Tekila 0 3 18.83 2 Tekila 0 4 30.01

Table A1.76. Raw data for field trials – canola: CI & DSI

Year Site ID Cultivar Block CI DSI

2014 Low ACSN39 1 100.0 100.0 2014 Low ACSN39 2 100.0 100.0 2014 Low ACSN39 3 100.0 100.0 2014 Low ACSN39 4 100.0 100.0 2014 Low 45H29 1 0.0 0.0 2014 Low 45H29 2 0.0 0.0 2014 Low 45H29 3 0.0 0.0 2014 Low 45H29 4 0.0 0.0 2014 Low 7377 1 0.0 0.0 2014 Low 7377 2 0.0 0.0 2014 Low 7377 3 0.0 0.0 2014 Low 7377 4 0.0 0.0 2014 Low 7367 1 0.0 0.0 2014 Low 7367 2 0.0 0.0 2014 Low 7367 3 0.0 0.0 2014 Low 7367 4 0.0 0.0 2014 High ACSN39 1 100.0 100.0 2014 High ACSN39 2 100.0 100.0 2014 High ACSN39 3 100.0 100.0 2014 High ACSN39 4 100.0 100.0 2014 High 45H29 1 0.0 0.0 2014 High 45H29 2 0.0 0.0 2014 High 45H29 3 0.0 0.0 2014 High 45H29 4 0.0 0.0 2014 High 7377 1 0.0 0.0 2014 High 7377 2 0.0 0.0 2014 High 7377 3 0.0 0.0 2014 High 7377 4 0.0 0.0 2014 High 7367 1 0.0 0.0 2014 High 7367 2 0.0 0.0 2014 High 7367 3 0.0 0.0 2014 High 7367 4 0.0 0.0 2015 BDL ACSN39 1 30.0 16.7

Page 164: Effect of Plasmodiophora brassicae ... - University of Guelph

  157  2015 BDL ACSN39 2 100.0 43.3 2015 BDL ACSN39 3 20.0 6.7 2015 BDL ACSN39 4 30.0 10.0 2015 BDL 45H29 1 0.0 0.0 2015 BDL 45H29 2 0.0 0.0 2015 BDL 45H29 3 10.0 3.3 2015 BDL 45H29 4 0.0 0.0 2015 BDL 7377 1 0.0 0.0 2015 BDL 7377 2 0.0 0.0 2015 BDL 7377 3 10.0 3.3 2015 BDL 7377 4 10.0 3.3 2015 BDL 7367 1 0.0 0.0 2015 BDL 7367 2 0.0 0.0 2015 BDL 7367 3 0.0 0.0 2015 BDL 7367 4 0.0 0.0 2015 Low ACSN39 1 100.0 63.3 2015 Low ACSN39 2 100.0 76.7 2015 Low ACSN39 3 100.0 56.7 2015 Low ACSN39 4 100.0 63.3 2015 Low 45H29 1 0.0 0.0 2015 Low 45H29 2 0.0 0.0 2015 Low 45H29 3 10.0 3.3 2015 Low 45H29 4 0.0 0.0 2015 Low 7377 1 0.0 0.0 2015 Low 7377 2 10.0 3.3 2015 Low 7377 3 0.0 0.0 2015 Low 7377 4 0.0 0.0 2015 Low 7367 1 0.0 0.0 2015 Low 7367 2 10.0 3.3 2015 Low 7367 3 10.0 3.3 2015 Low 7367 4 0.0 0.0 2015 High ACSN39 1 100.0 86.7 2015 High ACSN39 2 100.0 63.3 2015 High ACSN39 3 100.0 66.7 2015 High ACSN39 4 100.0 63.3 2015 High 45H29 1 0.0 0.0 2015 High 45H29 2 10.0 6.7 2015 High 45H29 3 0.0 0.0 2015 High 45H29 4 0.0 0.0 2015 High 7377 1 20.0 6.7 2015 High 7377 2 0.0 0.0 2015 High 7377 3 0.0 0.0 2015 High 7377 4 0.0 0.0

Page 165: Effect of Plasmodiophora brassicae ... - University of Guelph

  158  2015 High 7367 1 0.0 0.0 2015 High 7367 2 0.0 0.0 2015 High 7367 3 40.0 26.7 2015 High 7367 4 0.0 0.0

Table A1.77. Raw data for field trials – canola: plant height

Year Site ID Cultivar Block 4 WAS 5 WAS 6 WAS 7 WAS 8 WAS 9 WAS

2014 Low ACSN39 1 3.11 7.52 12.52 22.14 27.95 2014 Low ACSN39 2 2.98 6.80 13.97 22.63 23.06 2014 Low ACSN39 3 2.85 6.15 14.72 22.24 26.94 2014 Low ACSN39 4 3.52 11.00 24.33 32.20 37.93 2014 Low 45H29 1 2.54 7.14 26.07 78.63 73.04 2014 Low 45H29 2 2.72 8.11 33.96 65.73 76.30 2014 Low 45H29 3 2.58 5.18 15.83 46.38 68.96 2014 Low 45H29 4 2.79 6.25 27.78 63.99 72.73 2014 Low 7377 1 2.72 5.08 15.97 44.50 64.51 2014 Low 7377 2 2.76 5.00 12.90 29.88 45.38 2014 Low 7377 3 2.53 5.83 17.73 35.97 51.60 2014 Low 7377 4 2.58 6.70 22.03 31.22 51.50 2014 Low 7367 1 3.97 6.91 19.96 61.82 71.44 2014 Low 7367 2 3.82 8.48 32.25 55.22 72.77 2014 Low 7367 3 3.18 5.56 17.63 38.88 62.50 2014 Low 7367 4 2.95 6.97 19.55 49.74 66.60 2014 High ACSN39 1 2.99 4.25 8.66 21.71 39.13 2014 High ACSN39 2 1.92 4.27 9.81 26.85 34.49 2014 High ACSN39 3 3.11 5.17 10.45 19.55 18.89 2014 High ACSN39 4 2.82 6.05 11.37 23.16 30.83 2014 High 45H29 1 2.36 3.40 9.22 25.08 36.23 2014 High 45H29 2 2.36 3.98 7.91 20.09 39.65 2014 High 45H29 3 2.68 4.25 7.23 22.99 43.67 2014 High 45H29 4 2.69 4.42 10.89 32.50 47.41 2014 High 7377 1 3.10 9.72 8.25 20.16 41.45 2014 High 7377 2 2.73 4.71 9.78 27.20 46.03 2014 High 7377 3 2.83 4.19 8.63 29.71 54.82 2014 High 7377 4 3.19 5.06 12.63 33.16 54.49 2014 High 7367 1 3.08 4.77 9.03 19.62 34.06 2014 High 7367 2 3.98 6.15 9.64 29.60 44.82 2014 High 7367 3 4.23 5.21 10.07 26.98 42.57 2014 High 7367 4 3.83 5.11 7.89 27.65 55.67 2015 BDL ACSN39 1 15.22 29.70 55.22 83.58 103.02 112.51

Page 166: Effect of Plasmodiophora brassicae ... - University of Guelph

  159  2015 BDL ACSN39 2 10.69 27.07 63.55 87.98 110.60 112.20 2015 BDL ACSN39 3 13.91 25.51 59.49 91.96 101.82 101.77 2015 BDL ACSN39 4 9.90 19.42 43.60 75.18 98.74 106.47 2015 BDL 45H29 1 6.88 18.03 52.91 78.34 103.32 111.35 2015 BDL 45H29 2 7.33 18.20 63.59 97.28 110.69 113.65 2015 BDL 45H29 3 7.53 19.03 53.92 90.42 105.38 97.98 2015 BDL 45H29 4 5.50 12.14 55.67 80.91 100.50 107.53 2015 BDL 7377 1 4.48 9.74 44.12 68.50 104.28 96.44 2015 BDL 7377 2 4.09 8.04 35.03 64.44 110.95 110.99 2015 BDL 7377 3 5.15 9.40 40.16 72.52 99.40 109.24 2015 BDL 7377 4 4.43 8.69 45.23 82.48 100.34 131.24 2015 BDL 7367 1 7.61 14.30 40.21 59.98 95.65 94.59 2015 BDL 7367 2 8.00 15.53 40.45 62.48 98.47 106.73 2015 BDL 7367 3 7.07 12.84 53.59 64.98 103.84 105.68 2015 BDL 7367 4 7.25 11.13 41.78 72.00 95.82 113.83 2015 Low ACSN39 1 17.67 32.22 66.70 83.05 89.75 95.96 2015 Low ACSN39 2 13.85 28.29 56.85 99.12 101.01 110.28 2015 Low ACSN39 3 14.01 24.65 56.12 83.89 102.24 103.57 2015 Low ACSN39 4 13.14 22.68 49.73 83.23 103.24 98.96 2015 Low 45H29 1 7.70 16.67 52.90 75.41 94.12 95.99 2015 Low 45H29 2 6.43 15.32 47.75 65.91 97.96 105.84

Page 167: Effect of Plasmodiophora brassicae ... - University of Guelph

  160  2015 Low 45H29 3 6.49 18.14 65.70 92.81 111.51 103.20 2015 Low 45H29 4 4.81 13.06 50.89 77.69 99.00 107.44 2015 Low 7377 1 6.07 13.47 47.13 68.82 105.83 99.92 2015 Low 7377 2 5.65 12.98 54.49 86.51 104.06 105.15 2015 Low 7377 3 5.73 13.48 52.83 81.21 97.51 106.30 2015 Low 7377 4 5.39 10.81 44.31 67.40 88.25 100.54 2015 Low 7367 1 12.55 23.07 63.84 81.87 78.06 88.76 2015 Low 7367 2 8.03 14.00 49.36 84.90 102.67 99.37 2015 Low 7367 3 6.72 14.80 47.20 73.20 99.74 107.56 2015 Low 7367 4 7.25 14.79 43.76 67.00 87.29 107.90 2015 High ACSN39 1 9.48 21.05 66.74 86.51 96.05 99.19 2015 High ACSN39 2 13.20 25.44 61.17 81.22 99.41 104.79 2015 High ACSN39 3 16.29 28.90 63.53 71.43 81.10 105.76 2015 High ACSN39 4 14.21 26.98 72.82 92.97 100.07 102.55 2015 High 45H29 1 6.50 15.80 62.23 82.51 100.23 97.63 2015 High 45H29 2 5.34 14.88 57.67 83.06 103.83 102.11 2015 High 45H29 3 6.68 13.70 58.05 89.55 109.35 105.86 2015 High 45H29 4 6.55 14.16 57.32 83.52 105.01 104.62 2015 High 7377 1 6.31 8.44 31.40 51.35 82.05 84.43 2015 High 7377 2 5.61 11.64 39.66 59.14 92.21 97.31 2015 High 7377 3 5.40 10.37 33.18 63.41 100.50 97.83 2015 High 7377 4 9.31 15.29 58.80 64.63 105.74 85.39 2015 High 7367 1 10.47 13.72 44.50 60.68 88.93 83.80 2015 High 7367 2 10.80 16.73 61.58 71.67 95.69 91.59 2015 High 7367 3 12.04 21.86 69.82 75.78 82.06 98.41 2015 High 7367 4 6.94 8.34 30.48 53.96 82.39 75.87

WAS = Weeks after seeding 2014 Low = 7 x 105 spores g-1 soil

2014 High = 7 x 106 spores g-1 soil

2015 BDL = Below detection limit (<1000 spores g-1 soil) 2015 Low = 3 x 106 spores g-1 soil 2015 High = 1 x 107 spores g-1 soil

Table A1.78. Raw data for field trials – canola: biomass (dry shoot weight)

Year Site ID Cultivar Block Dry shoot weight (g)

2014 Low ACSN39 1 15.50 2014 Low ACSN39 2 22.60 2014 Low ACSN39 3 17.80 2014 Low ACSN39 4 374.57 2014 Low 7377 1 889.82 2014 Low 7377 2 581.00 2014 Low 7377 3 238.29

Page 168: Effect of Plasmodiophora brassicae ... - University of Guelph

  161  2014 Low 7377 4 482.20 2014 Low 7367 1 362.70 2014 Low 7367 2 680.40 2014 Low 7367 3 335.83 2014 Low 7367 4 451.20 2014 Low 45H29 1 304.60 2014 Low 45H29 2 501.80 2014 Low 45H29 3 885.43 2014 Low 45H29 4 443.80 2014 High ACSN39 1 126.20 2014 High ACSN39 2 117.40 2014 High ACSN39 3 31.40 2014 High ACSN39 4 112.50 2014 High 7377 1 372.60 2014 High 7377 2 601.50 2014 High 7377 3 178.70 2014 High 7377 4 122.75 2014 High 7367 1 302.80 2014 High 7367 2 281.70 2014 High 7367 3 330.20 2014 High 7367 4 183.70 2014 High 45H29 1 247.10 2014 High 45H29 2 161.30 2014 High 45H29 3 285.40 2014 High 45H29 4 418.90 2015 BDL ACSN39 1 422.40 2015 BDL ACSN39 2 733.20 2015 BDL ACSN39 3 291.40 2015 BDL ACSN39 4 647.20 2015 BDL 45H29 1 398.80 2015 BDL 45H29 2 529.40 2015 BDL 45H29 3 394.20 2015 BDL 45H29 4 665.20 2015 BDL 7377 1 265.20 2015 BDL 7377 2 326.00 2015 BDL 7377 3 535.60 2015 BDL 7377 4 260.00 2015 BDL 7367 1 375.60 2015 BDL 7367 2 324.20 2015 BDL 7367 3 453.40 2015 BDL 7367 4 271.60 2015 Low ACSN39 1 241.40 2015 Low ACSN39 2 430.80

Page 169: Effect of Plasmodiophora brassicae ... - University of Guelph

  162  2015 Low ACSN39 3 219.60 2015 Low ACSN39 4 351.00 2015 Low 45H29 1 433.20 2015 Low 45H29 2 572.00 2015 Low 45H29 3 106.80 2015 Low 45H29 4 609.20 2015 Low 7377 1 362.40 2015 Low 7377 2 397.20 2015 Low 7377 3 407.00 2015 Low 7377 4 235.20 2015 Low 7367 1 207.00 2015 Low 7367 2 443.00 2015 Low 7367 3 512.80 2015 Low 7367 4 306.20 2015 High ACSN39 1 353.80 2015 High ACSN39 2 494.00 2015 High ACSN39 3 299.20 2015 High ACSN39 4 481.80 2015 High 45H29 1 379.40 2015 High 45H29 2 327.00 2015 High 45H29 3 223.40 2015 High 45H29 4 544.60 2015 High 7377 1 481.20 2015 High 7377 2 384.00 2015 High 7377 3 398.40 2015 High 7377 4 337.60 2015 High 7367 1 175.20 2015 High 7367 2 505.20 2015 High 7367 3 610.00 2015 High 7367 4 388.20

2014 Low = 7 x 105 spores g-1 soil

2014 High = 7 x 106 spores g-1 soil

2015 BDL = Below detection limit (<1000 spores g-1 soil) 2015 Low = 3 x 106 spores g-1 soil 2015 High = 1 x 107 spores g-1 soil

Table A1.79. Raw data for field trials – canola: maturity (% plants at flowering and seepod

development stage)

Year Site ID Cultivar Block

% Flowering

% Seedpod

2014 Low ACSN39 1 0 80 2014 Low ACSN39 2 20 20

Page 170: Effect of Plasmodiophora brassicae ... - University of Guelph

  163  2014 Low ACSN39 3 50 20 2014 Low ACSN39 4 30 60 2014 Low 7377 1 60 40 2014 Low 7377 2 40 60 2014 Low 7377 3 40 60 2014 Low 7377 4 20 80 2014 Low 7367 1 30 70 2014 Low 7367 2 40 60 2014 Low 7367 3 80 20 2014 Low 7367 4 50 40 2014 Low 45H29 1 20 80 2014 Low 45H29 2 30 70 2014 Low 45H29 3 40 60 2014 Low 45H29 4 20 80 2014 High ACSN39 1 80 10 2014 High ACSN39 2 60 0 2014 High ACSN39 3 20 20 2014 High ACSN39 4 50 10 2014 High 7377 1 90 10 2014 High 7377 2 50 30 2014 High 7377 3 80 10 2014 High 7377 4 100 0 2014 High 7367 1 60 0 2014 High 7367 2 90 10 2014 High 7367 3 80 0 2014 High 7367 4 90 10 2014 High 45H29 1 40 50 2014 High 45H29 2 70 30 2014 High 45H29 3 90 0 2014 High 45H29 4 60 30 2015 BDL ACSN39 1 100 100 2015 BDL ACSN39 2 100 100 2015 BDL ACSN39 3 100 100 2015 BDL ACSN39 4 100 100 2015 BDL 7377 1 100 100 2015 BDL 7377 2 100 100 2015 BDL 7377 3 100 100 2015 BDL 7377 4 100 100 2015 BDL 7367 1 100 100 2015 BDL 7367 2 100 100 2015 BDL 7367 3 100 100 2015 BDL 7367 4 100 100 2015 BDL 45H29 1 100 100

Page 171: Effect of Plasmodiophora brassicae ... - University of Guelph

  164  2015 BDL 45H29 2 100 100 2015 BDL 45H29 3 100 100 2015 BDL 45H29 4 100 100 2015 Low ACSN39 1 100 100 2015 Low ACSN39 2 100 100 2015 Low ACSN39 3 100 100 2015 Low ACSN39 4 100 100 2015 Low 7377 1 100 100 2015 Low 7377 2 100 100 2015 Low 7377 3 100 100 2015 Low 7377 4 100 100 2015 Low 7367 1 100 100 2015 Low 7367 2 100 100 2015 Low 7367 3 100 100 2015 Low 7367 4 100 100 2015 Low 45H29 1 100 100 2015 Low 45H29 2 100 100 2015 Low 45H29 3 100 100 2015 Low 45H29 4 100 100 2015 High ACSN39 1 100 100 2015 High ACSN39 2 100 100 2015 High ACSN39 3 100 100 2015 High ACSN39 4 100 100 2015 High 7377 1 100 100 2015 High 7377 2 100 100 2015 High 7377 3 100 100 2015 High 7377 4 100 100 2015 High 7367 1 100 100 2015 High 7367 2 100 100 2015 High 7367 3 100 100 2015 High 7367 4 100 100 2015 High 45H29 1 100 100 2015 High 45H29 2 100 100 2015 High 45H29 3 100 100 2015 High 45H29 4 100 100

Page 172: Effect of Plasmodiophora brassicae ... - University of Guelph

  165  2014 Low = 7 x 105 spores g-1 soil

2014 High = 7 x 106 spores g-1 soil

2015 BDL = Below detection limit (<1000 spores g-1 soil) 2015 Low = 3 x 106 spores g-1 soil 2015 High = 1 x 107 spores g-1 soil

Table A1.80. Raw data for field trials – napa cabbage: CI & DSI

Year Site ID Cultivar Block CI DSI

2014 Low Chinagold 1 0.0 0.0 2014 Low Chinagold 2 0.0 0.0 2014 Low Chinagold 3 0.0 0.0 2014 Low Chinagold 4 0.0 0.0 2014 Low Emiko 1 0.0 0.0 2014 Low Emiko 2 0.0 0.0 2014 Low Emiko 3 0.0 0.0 2014 Low Emiko 4 0.0 0.0 2014 Low Mirako 1 0.0 0.0 2014 Low Mirako 2 50.0 33.3 2014 Low Mirako 3 40.0 16.7 2014 Low Mirako 4 0.0 0.0 2014 Low Yuki 1 0.0 0.0 2014 Low Yuki 2 0.0 0.0 2014 Low Yuki 3 0.0 0.0 2014 Low Yuki 4 0.0 0.0 2014 High Chinagold 1 0.0 0.0 2014 High Chinagold 2 0.0 0.0 2014 High Chinagold 3 0.0 0.0 2014 High Chinagold 4 0.0 0.0 2014 High Emiko 1 0.0 0.0 2014 High Emiko 2 0.0 0.0 2014 High Emiko 3 0.0 0.0 2014 High Emiko 4 0.0 0.0 2014 High Mirako 1 100.0 58.3 2014 High Mirako 2 60.0 26.7 2014 High Mirako 3 100.0 44.4 2014 High Mirako 4 100.0 44.4 2014 High Yuki 1 0.0 0.0 2014 High Yuki 2 0.0 0.0 2014 High Yuki 3 0.0 0.0 2014 High Yuki 4 0.0 0.0 2015 BDL Chinagold 1 0.0 0.0 2015 BDL Chinagold 2 0.0 0.0

Page 173: Effect of Plasmodiophora brassicae ... - University of Guelph

  166  2015 BDL Chinagold 3 0.0 0.0 2015 BDL Chinagold 4 0.0 0.0 2015 BDL Emiko 1 0.0 0.0 2015 BDL Emiko 2 0.0 0.0 2015 BDL Emiko 3 0.0 0.0 2015 BDL Emiko 4 0.0 0.0 2015 BDL Suzuko 1 0.0 0.0 2015 BDL Suzuko 2 0.0 0.0 2015 BDL Suzuko 3 0.0 0.0 2015 BDL Suzuko 4 0.0 0.0 2015 BDL Yuki 1 0.0 0.0 2015 BDL Yuki 2 0.0 0.0 2015 BDL Yuki 3 0.0 0.0 2015 BDL Yuki 4 0.0 0.0 2015 Low Chinagold 1 0.0 0.0 2015 Low Chinagold 2 0.0 0.0 2015 Low Chinagold 3 0.0 0.0 2015 Low Chinagold 4 0.0 0.0 2015 Low Emiko 1 0.0 0.0 2015 Low Emiko 2 0.0 0.0 2015 Low Emiko 3 0.0 0.0 2015 Low Emiko 4 0.0 0.0 2015 Low Suzuko 1 0.0 0.0 2015 Low Suzuko 2 0.0 0.0 2015 Low Suzuko 3 0.0 0.0 2015 Low Suzuko 4 0.0 0.0 2015 Low Yuki 1 0.0 0.0 2015 Low Yuki 2 0.0 0.0 2015 Low Yuki 3 0.0 0.0 2015 Low Yuki 4 0.0 0.0 2015 High Chinagold 1 0.0 0.0 2015 High Chinagold 2 0.0 0.0 2015 High Chinagold 3 0.0 0.0 2015 High Chinagold 4 0.0 0.0 2015 High Emiko 1 0.0 0.0 2015 High Emiko 2 0.0 0.0 2015 High Emiko 3 0.0 0.0 2015 High Emiko 4 0.0 0.0 2015 High Suzuko 1 0.0 0.0 2015 High Suzuko 2 0.0 0.0 2015 High Suzuko 3 0.0 0.0 2015 High Suzuko 4 0.0 0.0 2015 High Yuki 1 0.0 0.0

Page 174: Effect of Plasmodiophora brassicae ... - University of Guelph

  167  2015 High Yuki 2 0.0 0.0 2015 High Yuki 3 0.0 0.0 2015 High Yuki 4 0.0 0.0

Table A1.81. Raw data for field trials – napa cabbage: leaf length

Year Site ID Cultivar Block 4 WAS 5 WAS 6 WAS 7 WAS 8 WAS

2014 Low ChinaGold 1 7.52 10.82 13.46 15.46 14.65 2014 Low ChinaGold 2 7.45 11.69 14.47 15.18 16.91 2014 Low ChinaGold 3 6.41 10.75 13.70 15.30 17.10 2014 Low ChinaGold 4 9.51 12.26 14.53 18.94 19.41 2014 Low Emiko 1 9.65 12.75 14.22 16.96 18.23 2014 Low Emiko 2 9.05 14.10 15.17 19.05 18.82 2014 Low Emiko 3 10.04 12.30 14.25 16.80 17.17 2014 Low Emiko 4 11.51 14.05 16.17 20.74 18.71 2014 Low Yuki 1 4.64 9.06 12.39 13.51 12.78 2014 Low Yuki 2 8.13 9.40 12.24 12.64 13.70 2014 Low Yuki 3 8.08 9.80 11.78 11.97 12.13 2014 Low Yuki 4 5.65 10.56 12.87 12.04 10.53 2014 Low Mirako 1 9.09 12.98 13.45 14.44 14.18 2014 Low Mirako 2 9.22 13.25 14.43 16.15 15.36 2014 Low Mirako 3 8.92 12.80 14.40 17.91 15.83 2014 Low Mirako 4 9.16 11.65 14.69 15.72 15.34 2014 High ChinaGold 1 8.63 12.43 14.61 12.88 11.44 2014 High ChinaGold 2 7.67 11.60 11.80 12.74 12.34 2014 High ChinaGold 3 6.27 11.29 12.21 11.41 10.21 2014 High ChinaGold 4 6.57 9.39 9.73 11.67 11.73 2014 High Emiko 1 7.16 10.70 13.80 12.89 9.88 2014 High Emiko 2 7.48 10.90 12.54 12.88 12.10 2014 High Emiko 3 8.27 11.32 12.73 12.38 11.93 2014 High Emiko 4 7.95 10.75 11.52 11.82 11.18 2014 High Yuki 1 4.21 8.13 10.56 10.39 10.25 2014 High Yuki 2 5.79 10.12 11.24 10.66 10.35 2014 High Yuki 3 2.88 6.77 9.90 10.46 9.40 2014 High Yuki 4 2.92 5.31 8.99 10.06 9.59 2014 High Mirako 1 5.67 9.21 11.46 12.25 10.73 2014 High Mirako 2 7.48 10.65 12.82 11.04 11.84 2014 High Mirako 3 6.71 10.03 11.80 12.82 10.84 2014 High Mirako 4 7.42 11.70 11.80 12.38 12.07 2015 BDL Chinagold 1 9.45 7.18 29.55 32.95 28.15 2015 BDL Chinagold 2 10.26 9.84 24.60 30.40 27.55 2015 BDL Chinagold 3 10.57 9.26 23.40 27.57 27.40

Page 175: Effect of Plasmodiophora brassicae ... - University of Guelph

  168  2015 BDL Chinagold 4 9.05 9.48 21.56 22.92 21.83 2015 BDL Emiko 1 9.86 8.86 23.57 28.11 28.00 2015 BDL Emiko 2 8.73 7.23 25.06 30.59 30.45 2015 BDL Emiko 3 9.26 8.75 24.65 27.92 28.32 2015 BDL Emiko 4 8.39 8.42 23.70 24.37 26.45 2015 BDL Suzuko 1 9.92 8.77 30.91 33.61 25.47 2015 BDL Suzuko 2 8.54 8.96 22.06 24.61 26.33 2015 BDL Suzuko 3 8.44 7.20 22.75 22.04 28.69 2015 BDL Suzuko 4 7.87 7.54 18.13 17.63 22.43 2015 BDL Yuki 1 8.17 7.51 21.88 26.01 24.09 2015 BDL Yuki 2 9.25 7.69 23.51 24.80 22.22 2015 BDL Yuki 3 8.88 6.29 24.27 21.79 23.83 2015 BDL Yuki 4 9.48 6.63 23.01 19.43 20.61 2015 Low Chinagold 1 9.92 7.22 19.84 25.42 27.40 2015 Low Chinagold 2 10.02 7.21 17.44 25.55 25.89 2015 Low Chinagold 3 9.49 6.96 6.65 25.03 26.34 2015 Low Chinagold 4 8.28 6.91 19.01 23.07 27.37 2015 Low Emiko 1 8.41 7.60 20.31 28.91 29.57 2015 Low Emiko 2 8.96 6.73 18.92 25.49 29.98 2015 Low Emiko 3 8.13 6.48 19.09 23.54 25.08 2015 Low Emiko 4 8.66 6.73 15.66 20.52 24.14 2015 Low Suzuko 1 10.73 8.50 19.22 23.97 29.48 2015 Low Suzuko 2 9.47 8.49 14.75 25.37 25.75 2015 Low Suzuko 3 9.25 8.20 16.44 20.55 26.02 2015 Low Suzuko 4 8.75 8.26 13.10 20.11 26.75 2015 Low Yuki 1 9.30 6.03 14.49 25.31 26.32 2015 Low Yuki 2 8.39 6.05 16.31 23.67 26.64 2015 Low Yuki 3 8.28 5.89 6.59 20.65 24.86 2015 Low Yuki 4 7.20 6.58 16.21 19.85 26.08 2015 High Chinagold 1 9.54 9.58 19.08 20.05 19.56 2015 High Chinagold 2 9.30 10.14 18.40 24.06 19.95 2015 High Chinagold 3 10.87 9.11 18.38 21.02 20.65 2015 High Chinagold 4 8.37 8.84 16.16 22.49 22.65 2015 High Emiko 1 8.41 8.23 18.38 21.45 25.05 2015 High Emiko 2 9.46 8.16 18.02 21.61 21.58 2015 High Emiko 3 11.46 7.06 18.69 22.99 18.99 2015 High Emiko 4 10.35 7.02 19.86 22.93 20.00 2015 High Suzuko 1 10.25 10.15 17.59 16.82 19.84 2015 High Suzuko 2 10.69 8.89 14.57 19.00 22.26 2015 High Suzuko 3 9.89 9.26 17.96 20.58 20.01 2015 High Suzuko 4 10.48 8.86 16.86 20.52 20.63 2015 High Yuki 1 8.18 7.41 14.84 13.93 16.57 2015 High Yuki 2 9.61 6.82 17.14 19.08 18.39

Page 176: Effect of Plasmodiophora brassicae ... - University of Guelph

  169  2015 High Yuki 3 8.91 5.98 15.52 18.53 17.21 2015 High Yuki 4 8.81 6.17 17.05 19.74 18.85

2014 Low = 7 x 105 spores g-1 soil

2014 High = 7 x 106 spores g-1 soil

2015 BDL = Below detection limit (<1000 spores g-1 soil) 2015 Low = 3 x 106 spores g-1 soil 2015 High = 1 x 107 spores g-1 soil

Table A1.82. Raw data for field trials – napa cabbage: biomass (dry shoot weight)

Year Site ID Cultivar Block

Dry shoot weight (g)

2014 Low ChinaGold 1 664.11 2014 Low ChinaGold 2 857.90 2014 Low ChinaGold 3 900.70 2014 Low ChinaGold 4 861.50 2014 Low Emiko 1 826.30 2014 Low Emiko 2 750.40 2014 Low Emiko 3 917.33 2014 Low Emiko 4 1276.00 2014 Low Mirako 1 460.71 2014 Low Mirako 2 1045.63 2014 Low Mirako 3 1207.30 2014 Low Mirako 4 1175.67 2014 Low Yuki 1 191.63 2014 Low Yuki 2 771.00 2014 Low Yuki 3 614.56 2014 Low Yuki 4 979.86 2014 High ChinaGold 1 685.50 2014 High ChinaGold 2 912.00 2014 High ChinaGold 3 922.67 2014 High ChinaGold 4 609.63 2014 High Emiko 1 423.67 2014 High Emiko 2 437.50 2014 High Emiko 3 771.20 2014 High Emiko 4 247.50 2014 High Mirako 1 494.25 2014 High Mirako 2 995.75 2014 High Mirako 3 554.50 2014 High Mirako 4 1292.67 2014 High Yuki 1 124.56 2014 High Yuki 2 627.88 2014 High Yuki 3 278.00 2014 High Yuki 4 233.00

Page 177: Effect of Plasmodiophora brassicae ... - University of Guelph

  170  2015 BDL ChinaGold 1 1465.60 2015 BDL ChinaGold 2 2298.60 2015 BDL ChinaGold 3 885.60 2015 BDL ChinaGold 4 699.40 2015 BDL Emiko 1 720.00 2015 BDL Emiko 2 2117.80 2015 BDL Emiko 3 906.80 2015 BDL Emiko 4 2572.80 2015 BDL Suzuko 1 836.60 2015 BDL Suzuko 2 2066.40 2015 BDL Suzuko 3 652.20 2015 BDL Suzuko 4 1854.90 2015 BDL Yuki 1 650.50 2015 BDL Yuki 2 682.00 2015 BDL Yuki 3 1290.80 2015 BDL Yuki 4 442.20 2015 Low ChinaGold 1 741.60 2015 Low ChinaGold 2 1784.80 2015 Low ChinaGold 3 393.80 2015 Low ChinaGold 4 1022.80 2015 Low Emiko 1 208.40 2015 Low Emiko 2 541.60 2015 Low Emiko 3 1625.00 2015 Low Emiko 4 370.80 2015 Low Suzuko 1 499.00 2015 Low Suzuko 2 737.00 2015 Low Suzuko 3 1639.20 2015 Low Suzuko 4 1197.80 2015 Low Yuki 1 755.20 2015 Low Yuki 2 1132.40 2015 Low Yuki 3 2628.60 2015 Low Yuki 4 994.20 2015 High ChinaGold 1 356.00 2015 High ChinaGold 2 1900.80 2015 High ChinaGold 3 980.00 2015 High ChinaGold 4 976.60 2015 High Emiko 1 1280.90 2015 High Emiko 2 265.40 2015 High Emiko 3 867.80 2015 High Emiko 4 515.80 2015 High Suzuko 1 512.60 2015 High Suzuko 2 828.80 2015 High Suzuko 3 1703.80

Page 178: Effect of Plasmodiophora brassicae ... - University of Guelph

  171  2015 High Suzuko 4 658.60 2015 High Yuki 1 685.00 2015 High Yuki 2 761.40 2015 High Yuki 3 823.80 2015 High Yuki 4 441.80 2014 Low = 7 x 105 spores g-1 soil

2014 High = 7 x 106 spores g-1 soil

2015 BDL = Below detection limit (<1000 spores g-1 soil) 2015 Low = 3 x 106 spores g-1 soil 2015 High = 1 x 107 spores g-1 soil

 

     

Page 179: Effect of Plasmodiophora brassicae ... - University of Guelph

  172  APPENDIX 2: SUPPLEMENTARY TABLES FOR CHAPTER THREE

Table A2.1 Mean monthly air temperature and total monthly rainfall during the growing period of canola at Normandin, Quebec, from 2007-2014. Month Temperature (oC) Rainfall (mm)

LTA1 Actual LTA Actual 2007 May 9.7 10.0 49.5 12.6 June 15.9 14.8 71.2 52.4 July 17.1 16.3 102.0 90.4 Aug 15.9 14.0 63.2 97.8 Sept 13.1 11.2 82.5 95.9 2008 May 9.3 8.0 49.5 38.4 June 15.8 14.9 71.4 128.4 July 17.2 17.8 102.0 119.8 Aug 15.9 16.8 63.2 52.4 Sept 13.1 10.9 79.7 50.6 2009 May 8.7 7.0 52.3 70.6 June 14.7 15.0 68.4 39.0 July 17.2 16.3 97.3 105.2 Aug 16.0 16.2 63.6 106.8 Sept 12.8 11.5 74.1 49.2 2010 May 9.1 10.7 48.9 43.0 June 14.8 13.8 69.2 62.8 July 17.3 18.9 93.3 47.8 Aug 16.0 17.1 64.3 50.0 Sept 11.6 11.3 86.7 184.7 2011 May 8.8 9.3 55.2 96.3 June 14.8 15.1 68.7 64.9 July 17.4 18.1 90.8 94.8 Aug 16.0 17.0 81.1 202.4 Sept 11.7 12.4 79.4 53.7 2012 May 9.1 11.1 54.6 76.2 June 15.1 16.6 72.9 104.4 July 17.5 17.7 86.6 39.1 Aug 16.0 16.9 85.9 103.3 Sept 11.7 11.9 88.1 120.9 2013 May 9.2 10.6 62.0 99.3 June 14.9 13.7 78.2 94.9 July 17.6 18.0 85.2 50.7

Page 180: Effect of Plasmodiophora brassicae ... - University of Guelph

  173  Aug 16.0 16.6 83.9 20.9 Sept 11.5 11.6 93.0 132.0 2014 May 9.6 10.2 59.6 67.1 June 15.4 16.7 76.1 78.1 July 17.6 17.9 85.8 101.1 Aug 16.2 17.2 93.5 143.4 Sept 11.5 11.3 95.3 85.3

1Based on 10-year average

Table A2.2. Decline in resting spores, qPCR: Type I MS

Source df Sum of Squares Mean Square F value Pr>F Rotation 5 11.62 2.32 9.29 0.0008 Block(Rotation) 6 2.10 0.35 1.40 0.2913 Error 12 3.00 0.25 Total 23 16.72 2.92 Table A2.3. Decline in resting spores, qPCR: Type III MS block(rotation) as Error Term

Source df Sum of Squares Mean Square F value Pr>F Rotation 5 11.62 2.32 6.64 0.0196 Block(Rotation) 6 2.10 0.35 1.40 0.2913 Error 12 3.00 0.25 Total 23 16.72 2.92 Table A2.4. Decline in resting spores, log residuals: Type I MS

Source df Sum of Squares Mean Square F value Pr>F Rotation 5 11.62 2.32 9.29 0.0008 Block(Rotation) 6 2.10 0.35 1.40 0.2913 Error 12 3.00 0.25 Total 23 16.72 2.92 Table A2.5. Decline in resting spores, qPCR: Type III MS block(rotation) as Error Term

Source df Sum of Squares Mean Square F value Pr>F Rotation 5 11.62 2.32 6.64 0.0196 Block(Rotation) 6 2.10 0.35 1.40 0.2913 Error 12 3.00 0.25 Total 23 16.72 2.92

Page 181: Effect of Plasmodiophora brassicae ... - University of Guelph

  174  Table A2.6. Controlled environment study: crop rotation and spore concentration: plant height 2 WAI

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0 . . . Residual 0.2977 0.06077 4.90 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 0.50 0.4813 Soil (S) 1 45 15.42 0.0003 Rotation (R) 1 45 1.67 0.2029 Inoculation (I) 1 45 3.38 0.0728 S x R 1 45 2.98 0.0910 R x I 1 45 0.01 0.9093 I x S 1 45 0.68 0.4139 S x R x I 1 45 0.10 0.7568 Repetition S x R 1 45 0.02 0.8913 Repetition x R x I 1 45 0.01 0.9419 Repetition x I x S 1 45 1.01 0.3210 Repetition x S x R x I 1 45 2.47 0.1231 Table A2.7. Controlled environment study: crop rotation and spore concentration: plant height 3 WAI

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0 . . . Residual 0.3233 0.06599 4.90 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 0.56 0.4600 Soil (S) 1 45 5.50 0.0235 Rotation (R) 1 45 1.94 0.1709 Inoculation (I) 1 45 3.98 0.0522 S x R 1 45 1.92 0.1722 R x I 1 45 0.55 0.6607 I x S 1 45 0.62 0.4366 S x R x I 1 45 0.35 0.5543 Repetition S x R 1 45 0.12 0.7283 Repetition x R x I 1 45 0.26 0.6140 Repetition x I x S 1 45 0.37 0.5485 Repetition x S x R x I 1 45 1.59 0.2143

Page 182: Effect of Plasmodiophora brassicae ... - University of Guelph

  175  Table A2.8. Controlled environment study: crop rotation and spore concentration: plant height 4 WAI

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0 . . . Residual 0.4233 0.08641 4.90 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 2.97 0.0917 Soil (S) 1 45 0.23 0.6320 Rotation (R) 1 45 0.18 0.6704 Inoculation (I) 1 45 4.17 0.0470 S x R 1 45 2.05 0.1593 R x I 1 45 0.14 0.7069 I x S 1 45 1.35 0.2512 S x R x I 1 45 0.45 0.5035 Repetition S x R 1 45 0.64 0.4273 Repetition x R x I 1 45 1.31 0.2590 Repetition x I x S 1 45 0.05 0.8202 Repetition x S x R x I 1 45 0.25 0.6185 Table A2.9. Controlled environment study: crop rotation and spore concentration: plant height 5 WAI

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0.04533 0.06688 0.68 0.2490 Residual 0.5768 0.1216 4.74 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 0.27 0.6090 Soil (S) 1 45 7.64 0.0083 Rotation (R) 1 45 0.20 0.6577 Inoculation (I) 1 45 4.58 0.0377 S x R 1 45 1.63 0.2087 R x I 1 45 0.14 0.7080 I x S 1 45 0.44 0.5085 S x R x I 1 45 0.53 0.4717 Repetition S x R 1 45 0.59 0.4481 Repetition x R x I 1 45 0.91 0.3441 Repetition x I x S 1 45 0.37 0.5467 Repetition x S x R x I 1 45 0.13 0.7178

Page 183: Effect of Plasmodiophora brassicae ... - University of Guelph

  176  Table A2.10. Controlled environment study: crop rotation and spore concentration: plant height 6 WAI

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0.4210 0.4397 0.96 0.1692 Residual 1.8665 0.3935 4.74 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 6.22 0.0164 Soil (S) 1 45 34.24 <0.0001 Rotation (R) 1 45 2.22 0.1431 Inoculation (I) 1 45 4.35 0.0428 S x R 1 45 0.11 0.7468 R x I 1 45 0.05 0.8279 I x S 1 45 0.02 0.9008 S x R x I 1 45 0.07 0.7941 Repetition S x R 1 45 0.25 0.6217 Repetition x R x I 1 45 0.01 0.9427 Repetition x I x S 1 45 0.39 0.5340 Repetition x S x R x I 1 45 0.04 0.8350 Table A2.11. Controlled environment study: crop rotation and spore concentration: Area Under Growth Stairs

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0 . . . Residual 286.52 58.4860 4.90 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 0.93 0.3411 Soil (S) 1 45 0.19 0.6639 Rotation (R) 1 45 0.77 0.3862 Inoculation (I) 1 45 4.44 0.0407 S x R 1 45 13.36 0.0007 R x I 1 45 0.18 0.6761 I x S 1 45 0.80 0.3760 S x R x I 1 45 0.38 0.5383 Repetition S x R 1 45 1.26 0.2941 Repetition x R x I 1 45 0.37 0.6953 Repetition x I x S 1 45 0.59 0.5569 Repetition x S x R x I 1 45 0.60 0.5538

Page 184: Effect of Plasmodiophora brassicae ... - University of Guelph

  177  Table A2.12. Controlled environment study – crop rotation and spore concentration: biomass (dry shoot weight)

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0.06136 0.07768 0.79 0.2148 Residual 0.5340 0.1126 4.74 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 6.53 0.0141 Soil (S) 1 45 6.78 0.0125 Rotation (R) 1 45 1.31 0.2578 Inoculation (I) 1 45 9.21 0.0040 S x R 1 45 16.16 0.0002 R x I 1 45 0.48 0.4910 I x S 1 45 0.00 0.9864 S x R x I 1 45 0.12 0.7262 Repetition S x R 1 45 0.17 0.6859 Repetition x R x I 1 45 0.15 0.7009 Repetition x I x S 1 45 0.33 0.5660 Repetition x S x R x I 1 45 0.04 0.8410 Table A2.13. Controlled environment study – crop rotation and spore concentration: biomass (dry root weight)

Random Effects Estimate Standard Error Z value Pr>Z Block(Repetition) 0 . . . Residual 0.2304 0.04704 4.90 <0.0001 Fixed Effects Numerator df Denominator df F value Pr>F Repetition 1 45 0.31 0.5832 Soil (S) 1 45 8.52 0.0055 Rotation (R) 1 45 0.08 0.7843 Inoculation (I) 1 45 0.94 0.3371 S x R 1 45 4.29 0.0440 R x I 1 45 1.20 0.2797 I x S 1 45 0.11 0.7466 S x R x I 1 45 0.00 0.9707 Repetition S x R 1 45 10.05 0.0027 Repetition x R x I 1 45 0.00 0.9523 Repetition x I x S 1 45 0.60 0.4411 Repetition x S x R x I 1 45 1.63 0.2077

Page 185: Effect of Plasmodiophora brassicae ... - University of Guelph

  178  Table A2.14. Raw data for decline in resting spores (g -1 soil) following susceptible canola – Normandin, QC

Break interval (years) Block Part Initial Eff.% R2 Final 0 3 1 31842906.8 99.4 0.97 489754836.7 0 3 2 10736450.5 99.4 0.97 13566309.7 0 4 1 18367851.3 99.4 0.97 20425424.5 0 4 2 2127172.3 99.4 0.97 2127172.3 1 58 1 3660434.3 99.4 0.97 3660434.3 1 58 2 550095.4 99.4 0.97 577176.2 1 59 1 7147390.5 99.4 0.98 9865309.3 1 59 2 4469322.2 99.4 0.98 4486544.3 2 81 1 721554.4 99.4 0.98 1038424.3 2 81 2 217536.2 99.4 0.98 260023.4 2 85 1 902728.1 99.4 0.98 902728.1 2 85 2 1747163.9 99.4 0.98 1747163.9 3 98 1 651739.4 99.4 0.98 747297.1 3 98 2 898102.4 99.4 0.97 1260552.5 3 111 1 110304.8 91.8 1.00 474457.9 3 111 2 160636.5 91.8 1.00 397480.0 5 112 1 149810.9 91.8 1.00 367118.9 5 112 2 89076.2 91.8 1.00 234167.2 5 103 1 49711.5 91.8 1.00 163516.6 5 103 2 48105.1 91.8 1.00 129578.9 6 93 1 95385.8 96.7 0.97 95385.8 6 93 2 1134565.0 96.7 0.97 1134565.0 6 94 1 208450.0 96.7 0.97 208450.0 6 94 2 166969.3 96.7 0.97 218985.7

Page 186: Effect of Plasmodiophora brassicae ... - University of Guelph

  179  Table A2.15. Raw data for decline in resting spores (g -1 soil) before and after a resistant canola crop – Normandin, QC

Sample date

Break interval (years)

Block Part Initial Eff.% R2 Final

Spring 2014 0 62 1 1651544.5 95.0 0.99 2407924.0 Spring 2014 0 62 2 2432379.4 95.0 0.99 4085490.3 Spring 2014 0 64 1 751138.1 95.0 0.99 1390224.9 Spring 2014 0 64 2 778070.2 95.0 0.99 1008971.3 Spring 2014 0 57 1 24030.5 104.0 1.00 2433273.5 Spring 2014 0 57 2 74935.4 104.0 1.00 792413.2 Spring 2014 0 61 1 2878129.9 104.0 1.00 222577291.0 Spring 2014 0 61 2 4400837.8 104.0 1.00 38820982.7 Spring 2014 0 63 1 865120.5 104.0 1.00 6112557.0 Spring 2014 0 63 2 1044708 104.0 1.00 19444601.8 Spring 2014 0 66 1 4842109.8 102.7 0.99 4842109.8 Spring 2014 0 66 2 17842108.9 102.7 0.99 43852775.5 Spring 2014 0 67 1 30437267.3 102.7 0.99 3012874574.7 Spring 2014 0 67 2 1897319.5 102.7 0.99 3063775.6 Spring 2014 0 68 1 9490721.5 102.7 0.99 9490721.5 Spring 2014 0 68 2 5393416.2 102.7 0.99 6333519.5 Spring 2014 1 72 1 980652.4 90.9 0.97 2581562.3 Spring 2014 1 72 2 6438091 90.9 0.97 7502654.0 Spring 2014 1 74 1 4171803.9 90.9 0.97 7822848.2 Spring 2014 1 74 2 7893526.2 90.9 0.97 17115389.4 Spring 2014 1 76 1 3070363.6 95.0 0.99 11387450.6 Spring 2014 1 76 2 2948671.3 95.0 0.99 33757681.4 Spring 2014 1 75 1 11829802.9 102.7 1.00 11829802.8 Spring 2014 1 75 2 795743 102.7 1.00 2902677.6 Spring 2014 1 78 1 25239920.1 96.7 1.00 28063929.4 Spring 2014 1 78 2 81249517.2 96.7 1.00 81249517.2 Spring 2014 1 79 1 54963075.3 96.7 1.00 55828524.1 Spring 2014 1 79 2 5481857.7 96.7 1.00 6075074.2 Spring 2014 1 80 1 1120969.6 96.7 1.00 1469296.0 Spring 2014 1 80 2 2489989.4 96.7 1.00 4048575.5 Spring 2014 1 84 1 2560205.8 96.7 1.00 3377778.0 Spring 2014 1 84 2 1095075.2 102.7 0.99 1376149.4 Spring 2014 2 105 1 943795.7 90.9 0.97 2506020.5 Spring 2014 2 105 2 1142227.3 90.9 0.97 1909299.3 Spring 2014 2 106 1 616045.7 90.9 0.97 1142510.9 Spring 2014 2 106 2 61862.8 95.0 0.99 96755.0 Spring 2014 2 109 1 346908.8 98.2 0.99 636929.6

Page 187: Effect of Plasmodiophora brassicae ... - University of Guelph

  180  Spring 2014 2 109 2 638354.4 98.2 0.99 737502.0 Spring 2014 2 110 1 626307.4 98.2 0.99 1816699.4 Spring 2014 2 110 2 1722110 98.2 0.99 4532246.3 Fall 2014 0 57 1 4009275.2 100.4 0.99 12548967.5 Fall 2014 0 57 2 4389860.8 100.4 0.99 18374536.3 Fall 2014 0 62 1 33975334 100.4 0.99 286276162.8 Fall 2014 0 62 2 13121072.1 100.4 0.99 52611456.8 Fall 2014 0 64 1 54824620.6 100.4 0.99 125596197.0 Fall 2014 0 64 2 104555166.7 100.4 0.99 1028095440.1 Fall 2014 0 63 1 6236049.5 100.4 0.99 48642096.4 Fall 2014 0 63 2 36554772.8 93.6 0.98 36554772.8 Fall 2014 0 61 2 55606010.6 93.6 0.98 55606010.6 Fall 2014 0 61 1 20959422.5 93.6 0.98 20959422.5 Fall 2014 0 66 1 28964918.2 98.0 0.99 1010164290.6 Fall 2014 0 66 2 1381877.6 98.0 0.99 5589932.9 Fall 2014 0 67 1 3435511.9 98.0 0.99 34678241.9 Fall 2014 0 67 2 40368238.5 98.0 0.99 36748276082.1 Fall 2014 0 68 1 19896775 98.0 0.99 224088420.3 Fall 2014 0 68 2 8684497.9 98.0 0.99 448972582.9 Fall 2014 1 72 1 10363884.7 93.6 0.98 10363884.7 Fall 2014 1 72 2 3604292.6 93.6 0.98 3604292.6 Fall 2014 1 74 1 29772023.1 93.6 0.98 29772023.1 Fall 2014 1 74 2 87501751.1 93.6 0.98 87501751.1 Fall 2014 1 76 1 16723855.5 98.2 0.99 16723855.5 Fall 2014 1 76 2 4320259.9 98.2 0.99 4320259.9 Fall 2014 1 78 1 9794831.8 109.1 0.98 16315751.8 Fall 2014 1 78 2 1300472.6 109.1 0.98 3961106.4 Fall 2014 1 79 1 473213.3 109.1 0.98 608892.2 Fall 2014 1 79 2 4910536.7 109.1 0.98 5144730.0 Fall 2014 1 80 1 8499003.4 109.1 0.98 14202418.5 Fall 2014 1 80 2 5312898.9 109.1 0.98 5312898.9 Fall 2014 1 84 1 4804560.5 109.1 0.98 4842045.6 Fall 2014 1 84 2 3694633.3 98.0 0.99 13733426.3 Fall 2014 1 75 1 7444487.3 102.7 1.00 7925338.2 Fall 2014 1 75 2 4594269.9 102.7 1.00 5067467.5 Fall 2014 2 105 1 19272362.8 108.7 0.98 19272362.8 Fall 2014 2 105 2 3016737.3 108.7 0.98 10785712.6 Fall 2014 2 106 1 1231227.5 108.7 0.98 4519633.6 Fall 2014 2 106 2 798544.8 108.7 0.98 2578024.3 Fall 2014 2 109 1 8208424.4 108.7 0.98 23565906.8 Fall 2014 2 109 2 2646966 108.7 0.98 9507192.6 Fall 2014 2 110 1 6943413 108.7 0.98 22825209.2 Fall 2014 2 110 2 61862.8 98.2 0.99 9235407.1

Page 188: Effect of Plasmodiophora brassicae ... - University of Guelph

  181  Spring 2015 0 57 1 773872 103.2 1.00 15382540.8 Spring 2015 0 61 1 4252068.6 103.2 1.00 20787474.6 Spring 2015 0 62 1 33233427.1 103.2 1.00 464890407.6 Spring 2015 0 63 1 27519.6 93.2 1.00 65331202.4 Spring 2015 0 64 1 100906027.2 103.2 1.00 100906027.2 Spring 2015 0 66 1 3147604.5 93.2 1.00 4540324.0 Spring 2015 0 67 1 958731.2 93.2 1.00 958731.2 Spring 2015 0 68 1 660392.3 93.2 1.00 660392.3 Spring 2015 1 69 1 23561.9 93.2 1.00 101009.3 Spring 2015 1 70 1 123488.8 93.2 1.00 409156.9 Spring 2015 1 82 1 5249.8 93.2 1.00 8946.4 Spring 2015 2 95 1 17864 93.2 1.00 542752.6 Spring 2015 2 96 1 73407.1 93.2 1.00 2094928.3  Table A2.16. Raw data for controlled environmental study – crop rotation and spore concentration: CI & DSI

Repetition Soil Spores ml-1 Cultivar Block CI DSI 1 Scott 1x106 ACSN39 1 71.4 47.6 1 Scott 1x106 ACSN39 2 80.0 53.3 1 Scott 1x106 ACSN39 3 100.0 70.4 1 Scott 1x106 ACSN39 4 100.0 83.3 1 Scott 0 ACSN39 1 0.0 0.0 1 Scott 0 ACSN39 2 0.0 0.0 1 Scott 0 ACSN39 3 0.0 0.0 1 Scott 0 ACSN39 4 0.0 0.0 1 Scott 1x106 45H29 1 0.0 0.0 1 Scott 1x106 45H29 2 0.0 0.0 1 Scott 1x106 45H29 3 0.0 0.0 1 Scott 1x106 45H29 4 0.0 0.0 1 Scott 0 45H29 1 0.0 0.0 1 Scott 0 45H29 2 0.0 0.0 1 Scott 0 45H29 3 0.0 0.0 1 Scott 0 45H29 4 0.0 0.0 1 Scott 1x106 7367 1 0.0 0.0 1 Scott 1x106 7367 2 0.0 0.0 1 Scott 1x106 7367 3 0.0 0.0 1 Scott 1x106 7367 4 0.0 0.0 1 Scott 0 7367 1 0.0 0.0 1 Scott 0 7367 2 0.0 0.0 1 Scott 0 7367 3 0.0 0.0 1 Scott 0 7367 4 0.0 0.0 1 Scott 1x106 7377 1 0.0 0.0

Page 189: Effect of Plasmodiophora brassicae ... - University of Guelph

  182  1 Scott 1x106 7377 2 0.0 0.0 1 Scott 1x106 7377 3 0.0 0.0 1 Scott 1x106 7377 4 0.0 0.0 1 Scott 0 7377 1 0.0 0.0 1 Scott 0 7377 2 0.0 0.0 1 Scott 0 7377 3 0.0 0.0 1 Scott 0 7377 4 0.0 0.0 2 Scott 1x106 ACSN39 1 100.0 62.5 2 Scott 1x106 ACSN39 2 87.5 54.2 2 Scott 1x106 ACSN39 3 80.0 60.0 2 Scott 1x106 ACSN39 4 100.0 100.0 2 Scott 0 ACSN39 1 0.0 0.0 2 Scott 0 ACSN39 2 0.0 0.0 2 Scott 0 ACSN39 3 0.0 0.0 2 Scott 0 ACSN39 4 0.0 0.0 2 Scott 1x106 45H29 1 0.0 0.0 2 Scott 1x106 45H29 2 0.0 0.0 2 Scott 1x106 45H29 3 0.0 0.0 2 Scott 1x106 45H29 4 0.0 0.0 2 Scott 0 45H29 1 0.0 0.0 2 Scott 0 45H29 2 0.0 0.0 2 Scott 0 45H29 3 0.0 0.0 2 Scott 0 45H29 4 0.0 0.0 2 Scott 1x106 7367 1 0.0 0.0 2 Scott 1x106 7367 2 0.0 0.0 2 Scott 1x106 7367 3 0.0 0.0 2 Scott 1x106 7367 4 0.0 0.0 2 Scott 0 7367 1 0.0 0.0 2 Scott 0 7367 2 0.0 0.0 2 Scott 0 7367 3 0.0 0.0 2 Scott 0 7367 4 0.0 0.0 2 Scott 1x106 7377 1 0.0 0.0 2 Scott 1x106 7377 2 0.0 0.0 2 Scott 1x106 7377 3 0.0 0.0 2 Scott 1x106 7377 4 0.0 0.0 2 Scott 0 7377 1 0.0 0.0 2 Scott 0 7377 2 0.0 0.0 2 Scott 0 7377 3 0.0 0.0 2 Scott 0 7377 4 0.0 0.0 1 Elora 1x106 ACSN39 1 80.0 30.0 1 Elora 1x106 ACSN39 2 60.0 30.0 1 Elora 1x106 ACSN39 3 50.0 26.7 1 Elora 1x106 ACSN39 4 62.5 20.8

Page 190: Effect of Plasmodiophora brassicae ... - University of Guelph

  183  1 Elora 0 ACSN39 1 0.0 0.0 1 Elora 0 ACSN39 2 0.0 0.0 1 Elora 0 ACSN39 3 0.0 0.0 1 Elora 0 ACSN39 4 0.0 0.0 1 Elora 1x106 45H29 1 0.0 0.0 1 Elora 1x106 45H29 2 0.0 0.0 1 Elora 1x106 45H29 3 0.0 0.0 1 Elora 1x106 45H29 4 0.0 0.0 1 Elora 0 45H29 1 0.0 0.0 1 Elora 0 45H29 2 0.0 0.0 1 Elora 0 45H29 3 0.0 0.0 1 Elora 0 45H29 4 0.0 0.0 1 Elora 1x106 7367 1 0.0 0.0 1 Elora 1x106 7367 2 0.0 0.0 1 Elora 1x106 7367 3 0.0 0.0 1 Elora 1x106 7367 4 0.0 0.0 1 Elora 0 7367 1 0.0 0.0 1 Elora 0 7367 2 0.0 0.0 1 Elora 0 7367 3 0.0 0.0 1 Elora 0 7367 4 0.0 0.0 1 Elora 1x106 7377 1 0.0 0.0 1 Elora 1x106 7377 2 0.0 0.0 1 Elora 1x106 7377 3 0.0 0.0 1 Elora 1x106 7377 4 0.0 0.0 1 Elora 0 7377 1 0.0 0.0 1 Elora 0 7377 2 0.0 0.0 1 Elora 0 7377 3 0.0 0.0 1 Elora 0 7377 4 0.0 0.0 2 Elora 1x106 ACSN39 1 80.0 43.3 2 Elora 1x106 ACSN39 2 60.0 26.7 2 Elora 1x106 ACSN39 3 55.6 25.9 2 Elora 1x106 ACSN39 4 70.0 33.3 2 Elora 0 ACSN39 1 0.0 0.0 2 Elora 0 ACSN39 2 0.0 0.0 2 Elora 0 ACSN39 3 0.0 0.0 2 Elora 0 ACSN39 4 0.0 0.0 2 Elora 1x106 45H29 1 0.0 0.0 2 Elora 1x106 45H29 2 0.0 0.0 2 Elora 1x106 45H29 3 0.0 0.0 2 Elora 1x106 45H29 4 0.0 0.0 2 Elora 0 45H29 1 0.0 0.0 2 Elora 0 45H29 2 0.0 0.0 2 Elora 0 45H29 3 0.0 0.0

Page 191: Effect of Plasmodiophora brassicae ... - University of Guelph

  184  2 Elora 0 45H29 4 0.0 0.0 2 Elora 1x106 7367 1 0.0 0.0 2 Elora 1x106 7367 2 0.0 0.0 2 Elora 1x106 7367 3 0.0 0.0 2 Elora 1x106 7367 4 0.0 0.0 2 Elora 0 7367 1 0.0 0.0 2 Elora 0 7367 2 0.0 0.0 2 Elora 0 7367 3 0.0 0.0 2 Elora 0 7367 4 0.0 0.0 2 Elora 1x106 7377 1 0.0 0.0 2 Elora 1x106 7377 2 0.0 0.0 2 Elora 1x106 7377 3 0.0 0.0 2 Elora 1x106 7377 4 0.0 0.0 2 Elora 0 7377 1 0.0 0.0 2 Elora 0 7377 2 0.0 0.0 2 Elora 0 7377 3 0.0 0.0 2 Elora 0 7377 4 0.0 0.0

Table A2.17. Raw data for controlled environmental study – crop rotation and spore concentration: plant height

Repe-tition Soil

Rota-tion

Spores ml-1 Block 2 WAI 3 WAI 4 WAI 5 WAI 6 WAI

1 Scott 2 1x106 1 3.27 3.55 3.81 4.15 5.79 1 Scott 2 1x106 2 3.84 4.29 4.60 5.31 7.07 1 Scott 2 1x106 3 4.79 5.14 5.48 6.15 8.84 1 Scott 2 1x106 4 4.08 4.81 5.04 6.13 6.96 1 Scott 2 0 1 3.60 3.83 4.26 4.89 6.07 1 Scott 2 0 2 4.21 4.52 4.87 5.28 6.61 1 Scott 2 0 3 4.18 4.48 5.22 6.53 10.05 1 Scott 2 0 4 5.13 5.37 5.71 6.47 9.66 1 Scott 0 1x106 1 3.24 3.61 3.99 4.61 5.80 1 Scott 0 1x106 2 3.30 3.73 3.96 4.58 5.90 1 Scott 0 1x106 3 3.03 3.52 3.92 4.62 7.02 1 Scott 0 1x106 4 3.44 3.64 4.03 4.88 6.63 1 Scott 0 0 1 4.41 4.81 5.07 5.92 6.81 1 Scott 0 0 2 4.22 4.61 5.21 5.81 6.46 1 Scott 0 0 3 4.05 4.65 5.41 6.89 9.83 1 Scott 0 0 4 3.73 4.37 4.85 5.18 6.63 2 Scott 2 1x106 1 3.22 3.67 3.66 3.98 4.32 2 Scott 2 1x106 2 4.18 4.56 4.81 5.09 5.43 2 Scott 2 1x106 3 3.71 4.19 4.31 4.97 5.24 2 Scott 2 1x106 4 2.61 2.86 3.02 3.71 4.21 2 Scott 2 0 1 4.10 4.37 4.43 4.77 5.63

Page 192: Effect of Plasmodiophora brassicae ... - University of Guelph

  185  2 Scott 2 0 2 4.01 4.54 4.98 5.33 6.19 2 Scott 2 0 3 3.87 4.34 4.52 5.02 5.60 2 Scott 2 0 4 3.04 3.26 3.54 3.79 4.21 2 Scott 0 1x106 1 3.46 3.86 4.26 4.59 4.69 2 Scott 0 1x106 2 3.57 3.73 3.93 4.47 5.10 2 Scott 0 1x106 3 3.37 3.54 3.94 4.48 4.60 2 Scott 0 1x106 4 2.53 2.76 2.91 2.97 3.28 2 Scott 0 0 1 3.67 4.38 4.79 5.04 5.38 2 Scott 0 0 2 3.19 3.45 3.86 4.16 4.39 2 Scott 0 0 3 3.93 4.27 4.32 4.65 5.27 2 Scott 0 0 4 2.11 2.53 2.97 3.49 4.17 1 Elora 2 1x106 1 2.79 3.18 4.10 4.75 5.36 1 Elora 2 1x106 2 2.91 3.83 4.10 4.75 5.59 1 Elora 2 1x106 3 2.64 3.24 4.22 4.68 6.08 1 Elora 2 1x106 4 2.85 3.62 4.88 5.58 6.59 1 Elora 2 0 1 2.90 3.46 3.94 4.69 5.65 1 Elora 2 0 2 2.89 3.43 4.20 4.77 5.75 1 Elora 2 0 3 3.09 3.82 4.47 5.53 6.46 1 Elora 2 0 4 3.51 4.10 5.54 5.92 7.29 1 Elora 0 1x106 1 3.31 3.67 4.16 4.63 5.70 1 Elora 0 1x106 2 2.86 3.48 3.90 4.92 5.49 1 Elora 0 1x106 3 2.80 3.27 4.32 4.80 5.60 1 Elora 0 1x106 4 3.34 3.45 4.24 4.79 5.62 1 Elora 0 0 1 2.90 3.44 4.19 4.80 6.06 1 Elora 0 0 2 2.99 3.56 4.18 4.87 5.72 1 Elora 0 0 3 2.37 2.93 4.32 5.07 6.08 1 Elora 0 0 4 3.40 4.19 5.23 5.92 7.02 2 Elora 2 1x106 1 3.79 4.22 4.38 5.59 9.14 2 Elora 2 1x106 2 3.25 3.73 4.30 5.69 10.99 2 Elora 2 1x106 3 2.34 3.08 3.62 4.90 11.77 2 Elora 2 1x106 4 3.52 4.14 4.53 5.72 8.50 2 Elora 2 0 1 4.00 4.44 5.14 6.70 12.83 2 Elora 2 0 2 3.18 3.81 4.18 5.54 10.60 2 Elora 2 0 3 2.46 3.13 3.52 5.09 10.06 2 Elora 2 0 4 3.38 4.06 4.67 6.19 9.42 2 Elora 0 1x106 1 2.79 3.34 3.73 4.68 6.63 2 Elora 0 1x106 2 2.43 3.08 3.73 5.26 8.86 2 Elora 0 1x106 3 3.99 4.80 5.87 8.16 15.02 2 Elora 0 1x106 4 3.26 3.78 5.87 6.21 6.60 2 Elora 0 0 1 4.10 4.36 4.68 5.76 9.26 2 Elora 0 0 2 2.35 2.97 3.69 5.21 8.14 2 Elora 0 0 3 4.10 5.03 5.88 8.26 12.86 2 Elora 0 0 4 3.46 3.93 4.41 5.27 11.05

Page 193: Effect of Plasmodiophora brassicae ... - University of Guelph

  186  Table A2.18. Raw data for controlled environmental study – crop rotation and spore concentration: biomass (dry shoot weight)

Repetition Soil Rotation Spore ml-1 Block

Dry shoot weight (g)

1 Scott 2 1x106 1 3.30 1 Scott 2 1x106 2 3.83 1 Scott 2 1x106 3 3.13 1 Scott 2 1x106 4 2.75 1 Scott 2 0 1 3.50 1 Scott 2 0 2 4.00 1 Scott 2 0 3 5.00 1 Scott 2 0 4 3.67 1 Scott 0 1x106 1 2.11 1 Scott 0 1x106 2 2.00 1 Scott 0 1x106 3 3.00 1 Scott 0 1x106 4 2.50 1 Scott 0 0 1 3.11 1 Scott 0 0 2 2.11 1 Scott 0 0 3 3.25 1 Scott 0 0 4 2.50 2 Scott 2 1x106 1 1.90 2 Scott 2 1x106 2 2.44 2 Scott 2 1x106 3 2.11 2 Scott 2 1x106 4 2.00 2 Scott 2 0 1 2.89 2 Scott 2 0 2 3.22 2 Scott 2 0 3 2.60 2 Scott 2 0 4 2.57 2 Scott 0 1x106 1 2.00 2 Scott 0 1x106 2 1.71 2 Scott 0 1x106 3 2.67 2 Scott 0 1x106 4 2.80 2 Scott 0 0 1 2.25 2 Scott 0 0 2 1.75 2 Scott 0 0 3 2.00 2 Scott 0 0 4 4.75 1 Elora 2 1x106 1 2.00 1 Elora 2 1x106 2 1.80 1 Elora 2 1x106 3 1.83 1 Elora 2 1x106 4 2.20 1 Elora 2 0 1 2.00 1 Elora 2 0 2 2.50 1 Elora 2 0 3 2.00 1 Elora 2 0 4 3.40

Page 194: Effect of Plasmodiophora brassicae ... - University of Guelph

  187  1 Elora 0 1x106 1 2.80 1 Elora 0 1x106 2 3.10 1 Elora 0 1x106 3 2.00 1 Elora 0 1x106 4 2.78 1 Elora 0 0 1 2.90 1 Elora 0 0 2 2.60 1 Elora 0 0 3 2.33 1 Elora 0 0 4 3.56 2 Elora 2 1x106 1 2.10 2 Elora 2 1x106 2 3.20 2 Elora 2 1x106 3 3.70 2 Elora 2 1x106 4 3.10 2 Elora 2 0 1 3.80 2 Elora 2 0 2 2.00 2 Elora 2 0 3 3.70 2 Elora 2 0 4 5.44 2 Elora 0 1x106 1 3.20 2 Elora 0 1x106 2 4.70 2 Elora 0 1x106 3 5.50 2 Elora 0 1x106 4 3.90 2 Elora 0 0 1 4.40 2 Elora 0 0 2 4.00 2 Elora 0 0 3 5.20 2 Elora 0 0 4 6.90