:f. :2. «Cr . . lefl‘ .. .. “3”ngan yuan a. a 1-7 awe ru- «at! .1l.L 311...}. .43.; v5? 3;: up S"- urn-q nq . . 4.41.”.qu . -I(vull'b|t-[ it‘lfiizggts‘ LIBRARY Q Michigan State 0200] g f University This is to certify that the dissertation entitled THE ROLE OF ROOT GROWTH TRAITS IN RESISTANCE TO THE BIOTIC STRESS, FUSARIUM ROOT ROT AND THE ABlOTlC STRESS, LOW SOIL PHOSPHORUS IN COMMON BEAN (PHASEOLUS VULGARIS L.). presented by KAREN ANN CICHY has been accepted towards fulfillment of the requirements for the Doctoral degree in Plant Breeding and Genetics 7 4/ / Major Weéor’s Signature [22,; / Z 200 C Date MSU is an Affirmative Action/Equal Opportunity Institution —~—.—.- -.--.—.-..gJ---.—.-.-.---.-.--.-.-.-.-.-.-.-.—.-.-.-.-o-.-.---n--.-.-.-—.-.-.-.-o-.—-.— PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATEDUE DAIEDUE DAIEDUE 6/07 p:/CIRC/DateDue.indd-p.1 THE ROLE OF ROOT GROWTH TRAITS IN RESISTANCE TO THE BIOTIC STRESS, FUSARIUM ROOT ROT AND THE ABIOTIC STRESS, LOW SOIL PHOSPHORUS IN COMMON BEAN (PHASEOLUS VULGARIS L.). By Karen Ann Cichy A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of ' DOCTOR OF PHILOSOPHY Plant Breeding and Genetics Program 2006 ABSTRACT THE ROLE OF ROOT GROWTH TRAITS IN RESISTANCE TO THE BIOTIC STRESS, FUSARIUM ROOT ROT AND THE ABIOTIC STRESS, LOW SOIL PHOSPHORUS IN COMMON BEAN (PHASEOLUS VULGARIS L.). BY Karen Ann Cichy Genetic and environmental variability for root architecture has been identified in common bean. The overall objective of this research was to determine if root architecture traits are an adaptation against an abiotic and a biotic stress. The stresses compared in this study were the fungal disease Fusarium root rot caused by F usarium solam' f.sp. phaseolz' (Fsp) and phosphorus deficient soils, both significantly reduce yield in major bean production regions. Fusarium root rot is greatly influenced by environmental conditions that stress plant roots, which has hindered progress in breeding resistant bean cultivars. The first objective was to determine if Fusarium root rot resistance is conferred by genesexpressed in the root or shoot and if root architecture plays a role in resistance. Reciprocal grafting of a resistant (FR266) and susceptible (Montcalm) bean cultivar was used to study resistance to Fsp in two root grth environments, one without additional stress (Experiment 1) and one with the additional stress of a compacted soil layer (Experiment 2). Reciprocal grafting revealed that root rot incidence was controlled by genes expressed only in the roots in Experiment 1. Variability for root architecture was present, but it did not affect root rot incidence. In Experiment 2, root and shoot genotype both dictated root rot incidence and root traits including root length and root dry weight appeared to be related to root rot incidence. Root architecture traits have been shown to be related to P uptake and influence tolerance to low P soils. The second objective was to study the relationship between root architecture and P uptake in an Andean recombinant inbred line (RIL) population developed from a cross between a low P tolerant (G19833) and susceptible (AND696) bean line. The RILS also varied in plant growth habit and this was examined in relation to root architecture and yield. The population was field- grown under both high and low P levels and root architecture traits, P uptake and seed yield were measured. A linkage map was developed and a QTL study was conducted to determine which regions of the genome controlled these traits. Two QTL for root length density were identified that explained nearly 40% of the phenotypic variation, but root traits were not important for low P tolerance as measured by P uptake and seed yield. Growth habit influenced yield differently across soil treatments and indeterminate RILS had higher root length density than determinate RILS in the high P treatment. The third objective was to examine the relationship between tolerance to low P soil and seed P, Fe, and Zn levels and identify QTL controlling these traits in the population. Iron and Zn are important to human nutrition and interact with P stored as phytic acid. Variability for seed Fe and Zn levels related to seed P was detected and QTL for these traits co-localized to linkage group B01 near the region of the gene that controls determinate plant grth habit. ACKNOWLEDGEMENTS I would like to thank my advisor Dr. Sieg Snapp, for her support and interest in my research. I was fortunate to be part of her lab group, where I gained a holistic view of agriculture science. I would also like to thank my co-advisor, Dr. James Kelly, his impeccable knowledge of plant breeding and common bean improvement was inspiring to me and what I have learned from him has helped me become a better scientist. I am grateful to the members of my guidance committee: Dr. Mitch McGrath, for his support in terms of openness to new research ideas and generous sharing of research equipment, and Dr. Willie Kirk, who helped direct and focus my research through helpfiJl conversations. I am very grateful to Dr. Matthew Blair at the International Center for Tropical Agriculture. He allowed me to conduct research at CIAT and also provided genetic and laboratory materials that made it possible for me to address thoroughly 3 complex research question. He was also a great mentor. I would like to thank the members of Dr. Snapp’s lab group, including Kitty Oneil, Rich Price, and Bill Quackenbush who where a great help in experimental preparation and were also great to work with. Thanks also to the members of Dr. Kelly’s lab for their help, including Evan Wright, Halima Awalie, Belinda Roman, Veronica Vallejo, Jen Wagner and Karolyn Terpestra. I would like to thank the . many people at CIAT who helped with the field and laboratory work presented here, _ especially Gina Caldas and Carlos Galeano. iv I would like to thank Dr. George Hosfield for being an excellent mentor during my graduate studies. I would like to thank my parents and my brothers and sister for their support and understanding. I am also thankful to my husband, David Mota-Sanchez for his faith and support. I would also like to thank my daughters Sofia and Liliana for being wonderful. I would also like to thank the many fi'iends I made while at Michigan State University who created a wonderfiil environment in which to work and live. TABLE OF CONTENTS LIST OF TABLES ............................................................................ xii LIST OF FIGURES .......................................................... '. ................ xvii CHAPTER 1: LITERATURE REVIEW: BREEDING A BETTER ROOT SYSTEM ......................................................................................... I Introduction .................................................................................. 1 Abiotic Stress: Phosphorus Deficiency .................................................. 3 Plant adaptation to low P soil ..................................................... 5 Phosphorus acquisition efficiency ....................................... 6 Root system architecture ......................................... 7 Root hairs ........................................................... 9 Mycorrhizae ...................................................... 1 1 Root exudates ..................................................... 11 Phosphorus utilization efficiency ....................................... 13 summary ........................................................................... 15 Biotic Stress: Fusarium Root Rot ........................................................ 16 Nature of Disease .................................................................. 17 Environmental Factors ............................................................ 1 8 SoilTemperature..........................................................19 Drought .................................................................... 20 CompactiOn ............................................................... 20 Methods of Control ............................................................... 22 vi Genetics ............................................................................. 24 Sources of Resistance ................................................... 24 Inheritance of Resistance ................................................ 25 Root Architecture and Disease .................................................. 26 Root and Shoot Growth Habit Coordination ......................... 28 Summary ........................................................................... 29 References .......................................................................... 30 Figures .............................................................................. 40 CHAPTER 2: FUSARIUM ROOT ROT INCIDENCE AND ROOT SYSTEM ARCHITECTURE IN GRAFTED COMMON BEAN LINES ............................. 43 Abstract ............................................................................. 43 Introduction ........................................................................ 44 Materials and Methods ............................................................ 48 Plant Materials ............................................................ 48 Reciprocal Grafting ...................................................... 48 Experiments .............................. ' ................................. SO Experiment 1 Non-compacted soil ............................ 50 Experiment 2: Compacted soil layer .......................... 52 Plant Measurements ..................................................... 54 Data Analysis ............................................................. 55 Results .............................................................................. 55 Experiment 1 .............................................................. 55 Experiment 2 .............................................................. 57 vii Discussion .......................................................................... 58 Conclusions ........................................................................ 60 References........._. ................................................................ 61 Tables ................................................................................ 64 Figures .............................................................................. 72 CHAPTER 3: THE RELATIONSHIP AMONG ROOT ARCHITECTURE TRAITS, PLANT GROWTH HABIT AND TOLERANCE TO LOW SOIL PHOSPHORUS LEVELS IN AN ANDEAN BEAN POPULATION .............. 73 Abstract .............................................................................. 73 Introduction ................ ......... 74 Materials and Methods ............................................................ 77 Results ............................................................................... 80 P uptake .................................... ' ................................ 80 P use efficiency ........................................................... 80 Seed P content ............................................................. 81 Yield ........................................................................ 81 Correlations ................................................................ 81 Root traits and P uptake ................................................. 82 Differences by Growth Habit ........................... A ................ 83 Discussion .......................................................................... 85 Conclusions ........................................................................ 87 References. ................................ ......................................... 89 Tables ................................................................................ 92 viii Figures ........................................................................... 101 CHAPTER 4: IDENTIFICATION OF QTL RELATED TO ROOT ARCHITECTURE TRAITS AND LOW PHOSPHORUS TOLERANCE IN AN ANDEAN BEAN POPULATION ..................................................... 107 Abstract ........................................................................... 107 Introduction ..... 108 Materials and Methods ......................................................... 111 Plant Material ............................................................ 111 Field Trials ............................................................... 111 DNA Isolation and Molecular Marker Analysis .................... 112 SSR Markers .................................................... 1 13 RAPD Markers ................................................. 113 AF LP Markers ................................................. 114 Genotypic Data Analysis .............................................. 115 Results and Discussion ......................................................... 116 Linkage Map ............................................................. 1 l6 QTL Identification ................. .. .................................... 117 Root Architecture ............................................... 1 17 P uptake ......................................................... 118 Seed Yield ...................................... ' ................. 119 P use efficiency ................................................. 120 Seed P content ................................................. 121 Conclusions ...................................................................... 121 ix References ........................................................................ 123 Tables .............................................................................. 126 Figures .............................................................................. 127 CHAPTER 5: QTL ANALYSIS OF SEED NUTRIENT LEVELS IN AN ANDEAN BEAN POPULATION SEGREGATING FOR TOLERANCE TO LOW PHOSPHORUS SOILS .......................................................... 137 Abstract ........................................................................... 137 Introduction ...................................................................... l 3 8 Materials and Methods ..................................................... ....140 Results ............................................................................. 143 Variability ............................................................... 139 ' Seed Weight ..................................................... 139 Seed P and Phytic acid ........................................ 144 Seed Fe and Zn ................................................. 144 Correlations .............................................................. 145 Growth Habit ............................................................ 145 QTL Identification.......................... ........................... 146 Seed Weight ..................................................... 146 Seed P and Phytic acid ........................................ 147 Seed Fe and Zn ................................................. 147 Discussion ........................................................................ 148 Conclusions .................................................................... ' ...150 References ........................................................................ 151 Tables .............................................................................. 155 Figures ............................................................................. 162 APPENDIX ................................................................................ 170 xi LIST OF TABLES CHAPTER 2: Table 1. Selected plant and seed characteristics of the common bean cultivars FR266 and Montcalm ......................................................................... 64 Table 2. Mean response root rot score, shoot dry weight, and root dry weight of different graft combinations of common bean lines FR266 and Montcalm in the presence (+ Fus) or absence (-FuS) of F usarium solani f.sp phaseoli inoculum. Data analyzed across 2 runs of experiment 1 ..................................................... 65 Table 3: Mean response for total root length and average root diameter of ungrafted Montcalm and FR266 in the presence (+ Fus) or absence (—FuS) of F usarium solani f.sp phaseolz' inoculum. Data analyzed across 2 runs of experiment 1 .................. 66 Table 4: Mean response and analysis of variance for root traits of different graft combinations of common bean lines FR266 and Montcalm in the presence of F usarium solani f.sp phaseolz’ inoculum. Data analyzed across 2 runs of experiment 1 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response ...................... 67 Table 5: Mean response for root traits of different graft combinations of common bean lines FR266 and Montcalm in the presence of F usarium solam’ f.sp phaseolz' inoculum. Data analyzed across 2 runs of experiment 1 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response, means are averaged by root genotype and shoot genotype ............................................................................. 68 Table 6. Mean response of root rot score, shoot dry weight, and root dry weight of different graft combinations of common bean lines FR266 and Montcalm grown in containers with a layer of compacted soil in the presence (+ Fus) or absence ( Fus) of Fusarium solani f. sp phaseoli inoculum, Data analyzed across 1 run (Nov. 2005) of experiment 2 .................................................................................... 69 Table 7: Mean response for total root length and average root diameter of ungrafted Montcalm and FR266 grown in containers with a layer of compacted soil in the presence .(+ Fus) or absence (-Fus) of Fusarium solam‘ f.sp phaseoli inoculum, Data analyzed across 2 runs of experiment 2 .................................................... 70 Table 8. Mean response and analysis of variance for root traits of different graft combinations of common bean lines FR266 and Montcahn grown in containers with . xii a layer of compacted soil in the presence of F usarz'um solam' f.sp phaseoli inoculum. Data analyzed across 1 run (Nov. 2005) of experiment 2 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response ............................................... 71 CHAPTER 3: Table 1. Analysis of variance of traits related to phosphorus uptake and use, and seed yield in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien, Colombia in 2000 and 2005 under two treatments: high and low soil phosphorus ................................................................ 92 Table 2. Mean plant growth traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and G19833 and the means and ranges of 75 recombinant inbred lines (RILS) developed from the parents. Means are also included for the check variety, Carioca. The experiment was planted in 2000 and 2005 in Darien, Colombia. Mean values are of 3 replications. P value indicates level of significant genotypic differences among the RILS for each trait ............... 93 Table 3. Phenotypic correlations among P uptake, P use efficiency, and seed yield in a population of 75 recombinant inbred lines from a AND696/G19833 cross grown in high (HP) or low (LP) soil phosphorus in Darien, Colombia in 2005 .................. 94 Table 4. Analysis of variance of root traits in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien, Colombia in 2005 in two environments: high and low soil phosphorus ...................................... 95 Table 5. Mean root growth traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and G19833 and the means and ranges of 75 - recombinant inbred lines (RILS) developed from the parents. Means are also included for the check variety, Carioca. The traits are from the 2005 field experiment planted in Darien, Colombia. Mean values are of 3 replications. P value indicates level of significant genotypic differences among the RILS for each trait ............................................................................................... 96 Table 6. Phenotypic correlations between root traits and P uptake and seed yield in a population of 75 recombinant inbred lines developed from a AND696/619833 cross grown in high (HP) or low (LP) soil phosphorus in Darien, Colombia in 2005 ...... 97 Table 7. Means of root and shoot traits of 75 RILS developed from a AND696/G19833 cross and grown under high and low soil phosphorus treatments in Darien, Colombia in 2000 and 2005, grouped and averaged by plant growth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.) ....... 98 xiii Table 8. Phenotypic correlations between root traits and P uptake and seed yield in a population of 75 recombinant inbred lines (RILS) developed from a AND696/G19833 cross grown under high or low soil phosphorus treatments in Darien, Colombia in 2005. For the analysis, RILS were grouped according to plant grth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.) .............................................................................. 99 Table 9. Phenotypic correlations between P uptake and seed yield and P seed content and seed yield in a population of 75 recombinant inbred lines (RILS) from a AND696/G19833 cross grown under high or low soil phosphorus treatments in Darien, Colombia in 2000 and 2005. For the analysis, RILS were grouped according to plant grth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.) ............................................................................. 100 CHAPTER 4: Table 1. Putative QTL for seed traits identified from 75 recombinant inbred lines developed from AND696/G19833 cross grown under high (HP) and low (LP) soil phosphorus conditions in Darien, Colombia in 2000 and 2005 ......................... 126 CHAPTER 5: Table 1: Analysis of variance of seed traits in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien, Colombia in 2000 ' and 2005 in two environments: high and low soil . phosphorus ..................................................................................... 1 55 Table 2: Seed traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and G19833 and the means and ranges of 75 recombinant inbred lines (RILS) developed from the parents. The experiment was planted in 2000 and 2005 in Darien, Colombia. Mean values presented (n=3). P value indicates level of significant genotypic differences among the RILS for each traits ...................... 156 Table 3. Phenotypic correlations among seed traits in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien in 2005 in high and low soil phosphorus. The values above the diagonal line are under low soil P and the values below the diagonal line are under high soil P ......................... 157 Table 4. Phenotypic correlations among seed traits in a population of 75 recombinant inbred lines developed fromAND696/G19833 and field grown in Darien in 2000 in high and low soil phosphorus. The values above the diagonal line are under low soil P and the values below the diagonal line are under high soil P ........................ 158 xiv Table 5. Seed traits in a population of 75 recombinant inbred lines developed from an AND696/G19833 cross and grown under high and low soil phosphorus conditions in Darien, Colombia in 2000 and 2005, grouped and averaged by plant growth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.). Mean values presented (n=3) ............................................................... 159 Table 6. Putative QTL for seed traits identified from 75 recombinant inbred lines developed from an AND696/G19833 cross grown under high (HP) and low (LP) soil phosphorus conditions in Darien, Colombia in 2000 and 2005. . . . . . . . .......... 160 APPENDIX: Table A2-1. Mean response for root dry weight, root length, and average root diameter in the top soil layer of different grafi combinations of common bean lines FR266 and Montcalm grown in containers with a layer of compacted soil in the presence (+ Fus) or absence (-F us) of F usarium solam’ f.sp phaseoli inoculum, Data analyzed across 1 run (Nov. 2005) ofexperiment 2............. 171 Table A2-2. Mean response for root dry weight, root length, and average root diameter in the middle soil layer of different graft combinations of common bean lines FR266 and Montcalm grown in containers with a layer of compacted soil in the presence (+ Fus) or absence (-Fus) of F usarium solam' f.sp phaseoli inoculum, Data analyzed across 1 run (Nov. 2005) of experiment 2 ..................................... 172 Table A2-3. Mean response for root dry weight, root length, and average root diameter in the bottom soil layer of different graft combinations of common bean lines FR266 and Montcalm grown in containers with a layer of compacted soil in the presence (+ Fus) or absence (-Fus) ofFusarium solani f.sp phaseoli inoculum, Data analyzed across 1 run (Nov. 2005) of experiment 2 ..................................... 173 Table A3-1. Bartlett test for variance homogeneity based on growth habit, where there were two classes, determinate and indeterminate, and 1 degree of freedom. Tests were conducted for shoot and seed traits measured under low P (LP) and high P (HP) treatments in the recombinant inbred line population developed from a cross between common bean lines AND696 and G19833 .................................... 174 Table A3-2. Bartlett test for variance homogeneity based on grth habit, where there were two classes, determinate and indeterminate, and 1 degree of freedom. Tests were conducted for roots traits measured under low P (LP) and high P (HP) treatments in the recombinant inbred line population developed from a cross between common bean lines AND696 and 619833 ............................................... 175 Table A4-1. Amplified fragment length polymorphisms (AF LP) names and selective primers used in developing a linkage map of AND696/G19833 ....................... 183 -XV Table A4-2. Simple sequence repeat (SSR) primer sequence, linkage group and source information for markers polymorphic between AND696 and G19833 and used to develop a linkage map of AND696/G19833 .................................... 184 Table A4-3. Random Amplified Polymorphism DNA (RAPD) primer sequence for markers polymorphic between AND696 and G19833 and used to develop a linkage map of AND696/G19833 recombinant inbred line population ......................... 188 Table A4-4. Bartlett test for variance homogeneity based on growth habit, where there were two classes, determinate and indeterminate, and 1 degree of freedom. Tests were conducted for seed traits measured under low P (LP) and high P (HP) treatments in the recombinant inbred line population developed from a cross between common bean lines AND696 and G19833 ................................................ 189 xvi LIST OF FIGURES CHAPTER 1: Figure 1. Responses of Arabidopsis root systems to supply of nutrients P, N, and S. Figure reproduced from Lopes-Bucio et a1. (2003) ........................................ 40 Figure 2: The Phosphorus Cycle (Potash and Phosphate Institute, 2004). . . . . . . . ......41 Figure 3: Fusarium root rot infection in beans (Schwartz, 2006) .................... _. ..42 CHAPTER 2: Figure 1. Photos of containers used in Experiment 1 (a) and Experiment 2 (b). The bulk densities of soil are listed in g/cm3 for each of the 3 layers of containers used in experiment 2 .................................................................................... 72 CHAPTER 3: Figure 1. Frequency distribution of root length density and specific root length in 75 recombinant inbred lines (RILS) developed fiom a cross between AND696 and G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. The distribution of RILS in LP is shown on the graphs on the left and on graphs on the right show distribution under HP. Frequency distributions are separated by soil P level with LP represented by graphs on the left and HP by graphs on the right. Within each graph RILS are also separated by grth habit. Arrows represent mean values of determinate and indeterminate RILS within each treatment. Means are the average of 3 replications ...................................................................... 101 Figure 2. Frequency distribution of root surface area and average root diameter in 75 recombinant inbred lines (RILS) developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. The distribution of RILS in LP is shown on the graphs on the left and on graphs on the right show distribution under HP. Frequency distributions are separated by soil P level with LP represented by graphs on the left and HP by graphs on the right. Within each graph RILS are also separated by growth habit. Arrows represent mean values of determinate and indeterminate RILS within each treatment. Means are the average of 3 replications ...................................................................... 103 xvii Figure 3. Regression of seed yield under low P by seed yield under high P in the 2005 growing season of the population in 75 recombinant inbred lines developed from a cross between AND696 and G19833. Horizontal bar represents mean yield under low P and vertical bar represents mean yield under high P. Note: G19833 yield values are shown from 2000, not 2005, because seed of this parent was not planted in 2005 ............................................................................................ 105 Figure 4. Regression of seed yield under low P by seed yield under high P in the 2000 growing season of the population in 75 recombinant inbred lines developed from a cross between AND696 and G19833. Horizontal bar represents mean yield under low P and vertical bar represents mean yield under high P ...................... 106 CHAPTER 4: Figure 1. Common bean linkage map of AND696 by 619833 developed from 75 recombinant inbred lines of the F57 population. The map contains RAPD, AF LP, SSR molecular markers and one phenotypic marker for a total of 6330M on 12 linkage groups. Primer information for markers at each locus can be found in Tables A4-1, A4-2, and A4—3 of the Appendix ................................................... 127 Figure 2. Frequency distribution of root traits in 75 recombinant inbred lines developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005). Means are the average of 3 replications ................................................... 129 Figure 3: Frequency distribution of shoot phosphorus concentration and phosphorus ‘ uptake in 75 recombinant inbred lines developed from a cross between AND696 and _ G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005). Means are the average of 3 replications. . . . . ....131 Figure 4. Frequency distribution of seed phosphorus content and seed yield in 75 recombinant inbred lines developed from a cross between AND696 and G19833 ' planted in Darien, Colombia in 2000 and 2005 in high (HP) and low (LP) phosphorus soil. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005 and AND696 was not grown in 2000). Means are the average of 3 replications .......................................... 132 Figure 5. Frequency distribution of phosphorus use efficiency (defined in this study . as the amount of seed yield per unit of P taken up by the plant) in 75 recombinant inbred lines developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005). Means are the average of 3 replications ......................... 134 xviii Figure 6. AND696/G19833 linkage map with QTL locations for seed yield, seed P content, root length density (RLD), root surface area (surfarea), shoot P concentration, P uptake, and P use efficiency (PUB). QTL are fiirther identified by treatment, high P (HP) and low P (LP) and year, 2000 and 2005 ..................... 135 CHAPTER 5: Fig. 1. Frequency distributions for seed weight and seed P in 75 recombinant inbred lines (RILS) fi'om the population AND696/G19833 grown in Darien Colombia in 2000 and 2005 under high (HP) and low (LP) soil phosphorus levels. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005 and AND696 was not grown in 2000). Values are the average of 3 replications. Graphs on the left are for 2005 and those on the right are for 2000 ......... . ................................................................................... 162 Fig. 2. Frequency distributions for seed phytic acid in 75 recombinant inbred lines (RILS) from the population AND696/G19833 grown in Darien Colombia in 2005 under high (HP) and low (LP) soil phosphorus levels. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005). Values are the average of 3 replications ................................................... 164 Fig. 3. Frequency distributions for seed Fe and Zn in 75 recombinant inbred lines (RILS) from the population AND696/G19833 grown in Darien Colombia in 2000 and 2005 under high (HP) and low (LP) soil phosphorus levels. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005 and AND696 was not grown in 2000). Values are the average of 3 replications. Graphs on the left are for 2005 and those on the right are for 2000 ............................................................................................. 165 Figure 4. AND696/G19833 linkage map with QTL locations for seed P, Fe, and Zn concentration, percent phytic acid in the seed (PA), and seed weight (seedwt). QTL are further identified by treatment, high P (HP) and low P (LP) and year, 2000 and 2005 ..... g ....................................................................................... 167 APPENDIX: Figure A3-l. Frequency distribution of shoot P concentration, P uptake, and P use efficiency of 75 recombinant inbred lines (RILS) developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2005 in high (HP) and low (LP) phosphorus soil. The distribution of RILS in LP is shown on the graphs on the xix left and on graphs on the right show distribution under HP. Frequency distributions are separated by soil P level with LP represented by graphs on the left and HP by graphs on the right. Within each graph RILS are also separated by growth habit. Arrows represent mean values of determinate and indeterminate RILS within each treatment. Means are the average of 3 replications ..................................... 176 Figure A3-2. Frequency distribution of seed P content in 75 recombinant inbred lines (RILS) developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2000 and 2005 in high (HP) and low (LP) phosphorus soil. The distribution of RILS in LP is shown on the graphs on the left and on graphs on the right Show distribution under HP. Frequency distributions are separated by soil P level with LP represented by graphs on the left and HP by graphs on the right. Within each graph RILS are also separated by growth habit. Arrows represent mean values of determinate and indeterminate RILS within each treatment. Means are the average of 3 replications ..................................................................... 179 Figure A3-3. Frequency distribution of seed yield in 75 recombinant inbred lines (RILS) developed from a cross between AND696 and G19833 planted in Darien, Colombia in 2000 and 2005 in high (HP) and low (LP) phosphorus soil. The distribution of RILS in LP is shown on the graphs on the left and on graphs on the right Show distribution under I—IP. Frequency distributions are separated by soil P level with LP represented by graphs on the left and HP by graphs on the right. Within each graph RILS are also separated by growth habit. Arrows represent mean values of determinate and indeterminate RILS within each treatment. Means are the average of 3 replications ..................................................................... 181 Chapter 1: Literature Review: Breeding a Better Root System Introduction: To understand how root systems function as an organ, it is useful to classify roots based on type and distribution. Roots are often grouped into four broad classes, the taproot, basal roots, adventitious roots, and lateral roots (Zobel, 1991). The taproot or the primary root is Of embryonic origin, basal roots originate from the root/shoot transition zone, adventitious roots originate from non—root tissue, and lateral roots originate from existing roots. Lateral roots can also be further classified into secondary and tertiary roots depending branching from the taproot (Fitter, 2002). There are some species specific specialized root types such as the proteiod root Clusters, which are clusters of densely positioned lateral roots, found in the Proteaceae family. Additionally, some species have enlarged, starchy roots that are used as food crops, including carrot (Daucus carota), beet (Beta vulgaris), and sweet potato (Ipomoea batatas) (Austin, 2002). Architecture describes the spatial configuration of the root system (Lynch, 1995). Root system architecture is formed by intrinsic developmental and response to environment type cues. Intrinsic mechanisms are those that are an essential part of plant development and allow for the characteristic architecture of a particular root system. Response mechanisms are pathways that determine how plants respond to external signals (Malamay 2005). The ability of a plant’s root system to respond to particular enviromnental conditions to improve the survival in a specific enviromnent is know as plasticity (Sultan, 1987). Natural variation exists between and within plant species for intrinsic and response type mechanisms of root architecture development. For example, in a recombinant inbred line population from two ecotypes of Arabidopsis, two QTL were identified for lateral root length. One QTL acted intrinsically and the other was responsive to mannitol concentration (Gerald et al. 2006). Arabidopsis has been used successfully to identify root system architectural responses to nutrients N, P, and S as seen in Figure 1. Both intrinsically and response regulated root system architectural traits can have a role in stress tolerance, and a growing number of studies measured genetic variation for root architecture and related that to stress avoidance or yield gains (Gerald et al., 2006; Laperche et al., 2006; F ita etial., 2006). A study of cultivated (Lactuca sativa L.) and wild (L. serriola) lettuce found that a QTL for taproot length co-localized with a QTL for moisture extracted fiom a deep soil zone (50-100 cm) (Johnson et al., 2000). Regions of the maize genome that explained genotypic variation for adventitious root mass co-localized with QTL for grain yield under water stress and non-water stress conditions, such that an increase in root mass was associated with an increase in yield at four separate QTL (Tuberosa et al., 2002). Adaptive plasticity for root mass and root length has also been observed in response to soil moisture, including drought and flooding stress in two Polygonum species (Bell and Sultan 1999). The scope of this review is to examine the literature closely for attempts at breeding better root systems and relate where possible applications to common bean improvement. Common bean (Phaseolus vulgaris L.) is an excellent candidate for improved root system architecture. It is an important grain legume often planted in marginal environments (Wortmann et al., 1998). Common bean is also very susceptible to a number of stresses. Common bean roots have a large root length and root length density as compared to other annual legumes, soybean (Glycine max L.) and field pea (Pisum sativum L.). Common bean was found to have a smaller median root depth (0.46m) than soybean (0.56m) and field pea (0.48m) based on the average of three years of data on a Haplustoll soil (Merrill et al., 2002). Within P.vulgaris, genetic variability for root architecture traits exist and some of these traits have been implicated in stress resistance. For example, basal root shallowness is associated with bean genotypes tolerant to low phosphorus soils (Liao et al., 2001). The remainder of this literature review will focus on two stresses detrimental to bean production in many parts of the world, the abiotic stress, phosphorus deficiency, and the biotic stress, Fusarium root rot. The nature Of each of these stresses and root system traits were examined along with other factors that may be useful for genetic control of these ailments. Abiotic Stress: Phosphorus Deficiency Phosphorus is an essential mineral for plant metabolic processes, including photosynthesis, synthesis and breakdown of carbohydrates, and energy transfer. Phosphorus is also a structural element of nucleic acids and phospholipids. Phosphorus levels of 0.3-0.5% plant dry weight are required for optimal plant growth (Marschner, 1995). Insufficient plant phosphorus levels result in reduced leaf size and number, flower set, and shoot to root ratios (Lynch et al., 1991). Phosphorus is extremely reactive and exists in the soil as organic P or one of at least 170 different mineral phosphates. Organic P makes up 20 to 80% of the soil P, a large fraction of which is phytic acid (inositol hexaphosphate) (Richardson, 1994). Both mineral and organic P forms found in soils are insoluble, and therefore there are very low levels of P in the soil solution available - for plant absorption at any one time (Holford, 1997). For most soils, the amount of P in solution is insufficient for crop production. There is an equilibrium that occurs with P in soil solution as plants absorb P (Whitney, 1988) (Figure 2). At the extreme end of low soil P availability are acid weathered soils, where P is in a complex with iron and aluminum. Such soils, specifically Andosols, Ultisoils, and Oxisols, make up 43% of the land area in the tropics. Many acid weathered soils adsorb or fix P so that 70 to 90% of P fertilizer applied will become part of these low solubility compounds (Sanchez and Salinas, 1981). Cultivation leads to mineralization of organic P and mineralization occurs rapidly in tropical soils as compared to temperate soils (Hedley et al., 1995). Plants take up phosphorus from the soil solution in the form of orthophosphate (l-IPO4'2 or HzPO4'1). Thetrelative amount of each anion depends on soil pH. Acid soils favor the formation of H2P04'l, and alkaline soils favor the formation of HPO4‘2. Phosphorus uptake by plants roots across the cell boundary occurs against a steep concentration gradient via an ATP dependent proton A syrnport (Bucher et al., 2001). Phosphorus uptake is greatest. at the root tip, and is reduced by soil drying. Phosphorus diffusion is limited in the soil to less than 6 mm, therefore roots must be very close to phosphorus in the soil for uptake to occur (Whitney, 1988). Root system surface area is very important for P uptake because the greater the surface area, the greater the potential contact with P anions (Lynch, 1995). Phosphorus is translocated from the roots to young leaves in the xylem as inorganic P. Remobilization of P (in organic or inorganic form) from older leaves to younger leaves and roots occurs via the phloem and is more prevalent in P deficient plants (Schachtman et al., 1998). Phosphorus deficient soils will continue to be a problem in the coming years, because P is a finite resource and the world’s supply of rock phosphate is. estimated to expire in 60-80 years (Beebe et al., 2006). Common bean is especially susceptible to low P soils in part due to the high P requirement associated with energy-intensive nodulation and N2 fixation. For this reason, common bean genotypes tolerant to low P soils when N fertilizer is added may not be as tolerant if dependent on fixation to supply N to the plant (Christiansen and Graham, 2002). Low P soils are a major constraint to bean production in regions of Africa and Latin America where farmers lack access to sufficient P fertilizer (Wortmann et al., 1998). Plant adaptation to low P soil: ' Nutrient efficiency is the effectiveness with which a plant system utilizes nutrient inputs to produce outputs. Phosphorus efficiency, has been defined in numerous ways, here it is considered as the ability to produce plant growth and yield in relation to the amount of available phosphorus (Lynch and Beebe, 1995). Tolerance to low P soils, or phosphorus efficiency, can be achieved by two distinct routes: acquisition efficiency and utilization efficiency. Acquisition efficiency reflects a plants ability to extract P from the environment. It has been shown to be related to root system traits that increase the root surface area and allow capture of more P form the soil, including specific root length, root hair density and mycorrhizae (Gahoonia and Nielsen, 2003). Root exudates that can free inorganic P from insoluble P forms are another form of acquisition efficiency (Marshcner, 1995). Utilization efficiency is the superior ability of a plant to convert acquired P into plant biomass and yield. Possible P utilization efficiency mechanisms . include reduced requirements and modified phenology. Increased seed P reserves is another form of utilization efficiency that may allow plants to get a better start in low P soils (Lynch and Beebe, 1995). Phosphorus acquisition efficiency: Root system traits that improve phosphorus acquisition efficiency, and act to capture more P from the soil by increasing the surface area of the root, have been studied extensively in bean and other crops. Complex root architectures may have evolved for the capture of immobile nutrients, such as P, from the soil. Even with greatly reduced lateral branching, Arabidopsis can capture sufficient nitrate ions (mobile), but require more lateral branching to effectively uptake sufficient phosphate ions (Fitter et al., 2002). Genotypic variability for P acquisition efficiency has been identified in many crops including rice, soybean, and common bean (W issuwa and Ae, 2001a; Yan et al., l995ab; Li et al., 2005). Landraces of rice showed improved tolerance to low P soils than improved varieties, suggesting there is still unexplored variation for P uptake that can be exploited for rice improvement (Wissuwa and Ae, 2001a). In common bean, wild gennplasm was not more adapted to low P soils than cultivated genotypes (Araujo et al., 1998). A screening of thousands of both Mesoamerican landraces and Mesoamerican and Andean improved common ' bean varieties for tolerance to low fertility soil revealed variability within each of the groups, suggesting that selected genotypes of either type can be useful as parental material in a low fertility breeding program (Singh et al., 2003). According to Yan et al. (1995b), Andean common beans are superior to Mesoamerican beans in low P soils. They often have higher yields, although Mesoamerican lines are more responsive to P fertilization (Y an et al., 1995b). Phosphorus deficient soils of the tropics are of two major classes: mineral low P, including oxisols and utisols, with inorganic bound P, and volcanic low P soils, mostly andisols, with high organic matter and organic bound P. In a study of 16 Common bean lines from the Andean and Mesoamerican gene pools, large genetic variation for P efficiency was discovered. The P acquisition efficiency appeared to be general and not soil specific (Y an et al., 1995a). Root system architecture Most studies that relate root architecture to P acquisition efficiency examine genotypes with diverse plant performance in P-lirniting environments, eg. media with high and low levels of P and observe root responses. In maize, P- deficient media promotes larger root systems in P-efficient and P-inefi‘icient genotypes. However, the P-efficient genotype had a larger root system and higher total root length than the P-inefficient genotype in low P conditions (Lui et al., 2004). A plant growth simulation model for rice grown in low P showed that improving root growth was the best way to enhance tolerance to P deficiency (Wissuwa, 2003). Low P nutrient solution was associated with altered basal root growth angle of some common bean genotypes, a change not exhibited with deficiencies of N, K, Ca, Mg, S or micronutrients (Bonser et al., 1996). Rubio et a1. (2003) found the shallowness of basal roots of common bean added a competitive advantage to shoot biomass production as compared to deep basal roots when P is not evenly distributed in soil depth but instead is found only in the topsoil. Phosphorus acquisition efficient bean genotypes have more lateral roots and shallower basal roots under low P conditions as compared to P inefficient genotypes grown under similar conditions (Lynch and Brown 2001). Low phosphorus insensitive Arabidopsis mutants did not switch from primary to lateral root growth as seen in wild type plants. The mutants also showed reduced expression of genes important in phosphate uptake, including a phosphate transporter and acid phosphatases (Sanchez-Calderon et al., 2006). In a study of four common bean lines, efficient lines had more adventitious roots than inefficient lines. Differences were more pronounced early in development (22 days after planting) than later in development (43 days after planting) (Miller et al., 2003). This type of root system modification has implications for acquisition efficiency because adventitious roots are shallow and so are located within the P-emiched topsoil. Adventitious rooting can also be categorized as a P use efficiency mechanism because they tend to have a lower carbon construction cost than other root types as they are less dense (Lynch and Ho, 2005). The next step after the identification of root architecture traits associated with nutrient efficiency is to understand the genetic basis of these traits, and determine if the inheritance of a given trait is such that it can be readily introduced to inefficient germplasm. Variability for tolerance to low P soils has been identified in the Andean gene pool (Yan et al., 1995a). In a population of recombinant inbred lines developed from a P acquisition efficient, large seeded Andean bean,G19833 and a P acquisition inefficient small seeded Mesoamerican bean DOR364, Liao et a1. (2004) also found root gravitropism to be correlated with P. use efficiency. These traits include basal root growth angle and shallow basal root length. Results from this growth pouch assay were correlated with field responses of bean to low P. Three QTL for root gravatropism traits all from G19833 detected using'the growth pouch assay were associated with QTL» for P uptake in low P field conditions. In the same population of RILS, a QTL for root length and one for specific root length on two different linkage groups co-localized with QTL for phosphorus uptake. The uptake QTL and the root QTL each explained about 15% of phenotypic variation for these traits (Beebe et al., 2006). Root hairs Root hairs, or trichoblasts are tubular outgrowths of rhizodennal cells. Since root hairs are thin, they are better able to explore soil for nutrients with low mobility in soil such as P. They are able to penetrate cracks in the soil and also can penetrate moderately compacted soils. Root hairs therefore aid in plant anchorage. Root exudates are mainly released from root hairs. The growth of the root system and the movement of root hair zone allow the nutrient availability to be maintained spatially. (J ungk, 2001). Root hairs have many plasmodesmata between cells to aid in symplastic transport of nutrients to the vascular tissue. Root hair characteristics including length, radius, and density are important for P uptake (Bucher et al., 2001). Root hairs have been shown to be important for P use efficiency in bean, white clover, cowpea, wheat, and barley. Yan et al. (2004) conducted a QT L mapping study of root traits associated with P use efficiency in 86 RILS of an Andean efficient (G19833) by Mesoamerica inefficient (DOR 364) cross in common bean. The efficient parent had higher root hair density, average root hair length and total root hair length. Root hair traits were induced by low P (2 uM) in both cultivars as compared to high P (1000 uM). H+ exudation and total acid exudation increased in low P as compared to high P in both parents, and was higher in G19833 than DOR364. Wang et al. (2004) analyzed root hair traits associated with P use efficiency in RILS of a wild x cultivated soybean cross. The wild genotype had a higher shoot biomass and‘P uptake. It also had higher root hair density and total root hair length per unit root length but lower average root hair length. None of these traits were however correlated with total P in the plant. 10 Mycorrhizae Mycorrhizae are symbiotic associations between plant roots and fimgi. They are widespread in nature. An estimated 83% of dicots and 79% of monocots have mycorrhizal associations (Marschner, 1995). Vesicular-arbuscular (VAM) is the only type of mycorrhizae associated with annual crop plants. In this symbiosis, host plants provide carbohydrates to the fungi and fungi primarily provide increased access to uptake of mineral nutrients to the host plant, especially for low mobility ions. Mycorrhizal plants take up 2-3 times more P per unit root length than non mycorrhizal plants (Tinker et al., 1992). The predominant mechanism by which VAM increase plant P uptake is by extending the surface area of the root system (Cardoso and Kuyper, 2006). In addition VAM can acquire inorganic P from some of the soil pools marginally available to plants, as shown in an experiment in an Oxisol soil with mycorrhizal and nonmycorrhizal maize (Cardoso et al., 2006). In the case of the phosphorus efficient bean cultivar ‘Carioca’, mycorrhizal plants were able to acquire more P fi'om the soil than nonmycorrhizal plants because hyphae had a greater affinity for P in the soil solution than the plant roots and therefore a greater influx of P occurred. The beneficial effects of the mycorrhizal symbiosis on plant phosphorus levels were evident at flowering and mid podfill (deSilveira and Nogueira Cardoso, 2004). Reot exudates Root exudates are able to increase phosphorus available to plants for uptake fi'om the inorganic or organic pools not immediately available for plant 11 uptake (Figure 2). The type and amount of root exudate released by a plant is genotype dependent. Proteoid roots of white lupin, which are clusters of closely spaced determinate rootlets are very effective in acquiring P from the soil. They are known to release citrate and acid phosphatases which release P from organic and inorganic compounds (Gilbert et al., 1999). Lupin roots when compared to soybean roots, were able to accumulate 4.8 times more P per unit root length than soybean, suggesting that where soybean relied on increased root length to capture more P, lupin used root exudation to increase P availability (Watt and Evans, 2003). A study of exudates of three grain legumes, fava bean (Viciafaba L.), field pea (Pisum sativum L.) and white lupin (Lupinus albus L.) and the cereal spring wheat (Triticum aéstivum L.) demonstrated all species’ roots exude acid phosphatase, but the legumes more so than the wheat. There was also variability among species as to which soil P pools (inorganic and organic) were accessed (Nuruzzaman et al., 2006). A purple acid phosphatase gene isolated from Medicago truncatula was expressed in Arabidopsis roots and the transgenic plants were able to use organic P in the form of phytate as their sole P source (Xiao et al., 2006). This study illustrated the potential for P acquisition efficient crop plants to be developed via mechanisms that make plant unavailable P available. Another impOrtant point is that plant species and even genotypes have significant variability for the types and amounts of P solubilizing enzymes they contain and express. In common bean, genotypic differences for root exudation of organic acids (citrate, acetate, and tartrate) also suggest a mechanism for P acquisition 12 efficiency. The large seeded Andean genotypes G19833 and G19839 exude more organic acids from their roots than Mesoamerican genotypes DOR364 and G212121. These organic acids are useful to mobilize Al and Fe bound P from soils for uptake by the plant roots (Shen et al., 2002). Low P was associated with an increase in acid phosphatase activity in the roots and on the root surface of the P-efficient genotype, whereas only slight increased acidity was found on the root surface of the inefficient genotype (Lui et al., 2004). Root exudation is also soil type dependent. The cation-anion balance of the plant plays a role in the pH of the rhizosphere, which in turn can effect the P mobilization in the soil. The value of rhizosphere pH in mobilizing P is soil type dependent. In a luvisol, where P was mainly bound to Ca, a decrease in rhizosphere pH enhanced P mobilization, whereas in an oxisol where P was mainly bound to Al and Fe an increase in rhizosphere pH mobilized more P (Gahoonia and Nielsen 2004). Phosphorus utilization efficiency: . An evaluation of 26 bean lines with varying levels of P efficiency, showed that the efficient and inefficient lines did not differ in P uptake, but differed in pods per plant and seeds per pod under low P conditions. At 65 days after emergence the efficient lines had a greater proportion of total P in pods and less in stems and leaves than that of the inefficient lines that had more P comparatively in stems and leaves and less in the pods (Y oungdahl, 1990). A more in depth study of the P efficient common bean cultivar ‘Calima’ looked at remobilization of P from stem and leaves in low and high P growing media. The 13 study found that in low P media remobilization occurs earlier (56-70 days after emergence) than in high P media (70 days after emergence). Under low P conditions 80% of P stayed in the root system, whereas in high P only 20% stayed in the roots (Snapp and Lynch, 1996). P use efficiency can also demonstrate itself in terms of carbon cost. A root system that requires less energy (i.e. lower root respiration) can increase P uptake per unit carbon invested. A comparison of P utilization in four common bean genotypes indicated that efficient and inefficient genotypes have a similar rate of P absorption per root dry weight. Phosphorus efficient bean genotypes G19839 and G1937 had a significantly lower rate of root respiration than the P inefficient genotypes DOR364 and Porillo Sintetico. The efficient genotypes had a higher percentage of their biomass as roots than the P-inefficient genotypes in low P soils (Nielsen et al., 2001). A high level of seed P is considered to be a form of P use efficiency as it is indicative of the effect of P remobilization. In upland rice, the amount of shoot and root dry weight produced at six weeks of growth was highly correlated with the amount of phosphorus in the seed, suggesting high seed P as a trait to select for when breeding cultivars tolerant to low P soils (Hedley et al., 1994). A greenhouse experiment with four annual legume species Wedicago polymorpha, T nfolium subterranean T. balanSae, Ornithopus compressus) indicated under low P fertilizer levels demonstrated that increases in seed P concentration was associated with increased plant yields (Bolland and Paynther, 1990) 14 Brachiaria species are the most widely planted forage grasses worldwide, and are well adapted to low P soils. In addition to a root system efficient in P acquisition, this species also uses P efficiently in the plant, as seen in the cultivar Mulato. P use efficiency appeared to be related to enhanced sugar catabolism and increased phosphohydrolase activity in low P conditions, which allows rapid turnover of P (Nanamori et al., 2004). The QTL for phosphorus uptake and use efficiency were each found on chromosomes 2 and 12 of the rice genome, with the P uptake QTL coming from one parent and the use efficiency coming from the other parent used in the study. While P uptake was positively correlated with plant dry weight, use efficiency was negatively correlated. This indicates that P use efficiency may not be a positive adaptation worth exploring by rice breeders, but may be a response of plants with poor P uptake (Wissuwa et al., 1998). Summary ' The use of root properties as selection criteria to improve P use efficiency through plant breeding requires a trait that is easily identifiable, for which large genetic variability exists and the mode of inheritance is known (Gahoonia and Nielsen, 2004). This poses a challenge using root traits in breeding as characterization is notoriously difficult below ground. Success in breeding rice to be more tolerant to low P soils was attained through a QTL for variation in P uptake. The trait considered in that study was simply plant P concentration; no root system measurements were taken. This QTL was transferred from one 15 cultivar to another and the resulting near isogenic line more than tripled its grain yield (Wissuwa et al., 2001b). The same success with breeding for improved tolerance to low P soils has yet to be realized in common bean, despite the large body of literature regarding potential mechanisms of P efficiency and genetic control of related traits. One should keep in mind that seed yield is the bottom line in improving common bean gennplasm for tolerance to low P soils. Singh et al. (2003) advocated field selections under general low soil fertility instead of just low soil P. In addition, and in direct contrast to previous finding by Yan et al. (1995a, 1995b) when field selections of more than 5000 landraces and improved common bean lines were conducted based on yield, it was Mesoamerican, not Andean races that were identified as better adapted to low soil fertility (Singh et al., 2003). The discrepancies observed most likely arise from what was measured as an indicator of P efficiency, overall yield or plant P uptake. Biotic Stress: Fusarium Root Rot In contrast to the large amount of research that relates low soil phosphorus tolerance to root architectural traits, there has been limited research that associates root architecture with tolerance to Fusarium root rot. There have been numerous studies illustrating that the severity of Fusarium root rot is highly influenCed by environmental stresses that impact root growth. This is consistent with an important role for root architecture in determining resistance. 16 Nature of disease: Root rot, caused by a combination of fungi, iS a major limiting factor in the production of Phaseolus vulgaris (Abawi and Pastor Corrales, 1990). The complex of pathogens believed to contribute to root rot include: F usarium solam’ f.sp. phaseoli, Rhizoctom'a solam' Kuhn, Pythium ultimum, Thielavz'opsis basicola (Berk. & Br.) Ferr. and Aphanomyces euteiches f.sp. phaseoli. In the majority of dry bean and snap bean producing regions of the United States, including Michigan, F. solani f.sp. phaseoli (Fsp) is considered the major pathogen of the complex (Chatterjee 1958; Keenan et al., 1974; Burke and Kraft 1974; Steadman et al., 1975; Saettler 1982). Yield losses of up to 84% have been attributed to Fusarium root rot (Keenan et al., 1974). Susceptibility to Fsp is developmentally regulated in beans. Fsp is not able to effectively cause infection until the plant develops a root system (10 days after planting). Earlier in development, considerably fewer spores attach to plant tissue than in older bean plants. Disease symptoms in younger plants are limited to a hypersensitive response at the inoculation site by the outermost cell layers (V ogeli-Lange et al., 1995). Symptoms of Fusarium root rot include red longitudinal lesions on the hypocotyls, taproot and lateral roots of the bean plants, followed by root rot. In severe infections complete rotting of a root system is observed (Schneider and Kelly, 2000). Increased adventitious rooting is also often a sign of disease infection (Chatterjee 1958; Steadman et al., 1975). Severity of Fsp increases as the plant develops, and with time it becomes possible to see symptoms on upper parts of the plant including chlorosis, stunting, and 17 premature defoliation, along with reduction in pod production (Burke and Barker, 1966). Although Fsp infection is readily visible on the hypocotyl and taproot, lateral root infection most significantly impacts crop yield. A plant with a healthy lateral root system can outgrow the detrimental effects of localized hypocotyl and taproot infection (Burke and Barker 1966). The nature of the infection by Fsp may be the reason for the importance of lateral roots. Infection is defined as direct penetration of fungal hyphae into healthy epidermis, wound sites on hypocotyl or. root tissue, or through hypocotyl stomata. The hyphae spread extracellularly in the cortex until cell death occurs, at which time hyphae begin to spread intracellularly. The pathogen is unable to pass the endodermis of the taproot, but can enter the vascular system of lateral root during their emergence (Christou and Snyder, 1962) suggesting that large loss of lateral roots leads to the plant’s inability to transport enough water and nutrients to the shoot, but this may not be a factor with taproot infection because the vascular system is able to remain functional. The disease cycle is shown in Figure 3. Environmental Factors: Under optimal growing conditions, Fsp is a miner pathogen of beans. Environmental conditions that increase the stress on plant roots and impede root growth, including drought, flooding, (Miller and Burke, 1977) soil compaction (Burke 1968; Miller and Burke, 1985), low soil fertility, low soil temperature (Burke et al., 1980), and plant competition (Burke, 1965) exacerbate incidence of Fusarium root rot. 18 Soil temperature Soil temperature plays an important role in root rot disease incidence caused by Fsp on common bean. Soil temperatures fluctuate seasonally and are impacted by air temperature, with the lowest average soil temperatures occurring early and late in the growing season. Common bean grows best at soil temperatures of 22 to 26°C. Temperatures above 30°C cause an inhibition in root growth by reducing cell division and root elongation. Temperatures below 18°C also significantly reduce yield. Chilling temperatures reduces membrane permeability and decrease the ability of the root system to uptake water. Studies conducted by Reddick (1917) and Sippell and Hall (1982a) indicated Fsp has a temperature tolerance of 12 to 35°C. Disease caused by Fsp appears to be favored by temperatures in the range of 16 to 22°C as compared to warmer temperatures (26 to 34°C). At temperatures above 30 °C there is reduced attachment of firngal spores to the-root surface which Can decrease amount of infection and disease severity. Suboptimum soil temperatures for bean grth increase Fusarium root rot incidence. Suboptimal soil temperatures cause decreased root growth, water uptake, and nutrient uptake. Such conditions exacerbate root rot caused by Fsp by limiting root growth which reduces a plant’s ability to compensate for a portion of the root system malfunctioning from rot. Beans: planted in warm soils are often able to escape the yield depressing effects of root rot that occur under cold soil temperatures. There is also an interaction between moisture, temperature, and root rot severity. Although moist soil is most favorable for germination of resting Fsp spores, it also favors the growth of the bean seedling and incidence of disease 19 is much greater in dry soils. In dry soils there is a larger temperature range over which infection occurs as compared to moist soils. (Buerkert and Marschner, 1992). Under low soil temperatures (21°C days 16°C nights) root growth was more reduced in the presence of Fsp as compared to inoculum free soil (Burke et al,1980) Drought Deep root systems appear to be an important factor in enhancing resistance to Fusarium root rot under dryland conditions. Fsp spores most widely inhabit soil at depths of 15-45 cm below ground, as do the majority of bean roots. . Dryden and Van Alfen (1984) found lower levels ostp infection on roots growing below 45 cm as compared to those roots above 45 cm. Under dryland conditions, soil water availability is often greatest below 60 cm, therefore, deep rooted cultivars may tolerate drought by accessing water 45 to 60 cm below the soil surface. Intact deep roots which are able to take up water may help plants avoid Fsp infection (Miller 1985; Burke and Miller, 1983). I Compaction Soil compaction is an increase in soil bulk density and a decrease in air filled pores as a result of force applied to the soil (Allrnaras et al., 1988). Excessive soil compaction, which impedes root growth and function, may be observed at a critical penetration resistance of 2 MPa and a DzDo ratio (oxygen diffusion coefficient in the soil relative to air) of zero at 10% air-filled porosity (Allrnaras et al.,-1988). Such soil compaction levels are frequently present in agricultural fields as a result of machinery used for planting, harvest, and other 20 crop maintenance activities (Harverson et al., 2005). Plant root systems respond to increases in soil bulk density or mechanical impedance by exerting a growth pressure to get through soil channels that are smaller than the root diameter (Clark et al., 2003). There is a positive correlation between root diameter and elongation rate under impeded conditions among different species. In addition cultivar differences for ability to penetrate soil with high bulk density have been shown in rice using a wax layer technique in a greenhouse study (Yu et al., 1995). Excessive soil compaction deprives roots of oxygen and they must undergo anaerobic respiration, this process produces less than half the energy of aerobic reSpiratibn and halts root growth and maintenance (Alhnaras et al., 1988). Grimes et al. (1975) observed in corn and cotton that in the presence of a compacted layer 30 cm below the soil, roots proliferate above the compacted layer. Asady et al. (1984) developed a controlled environment method to test bean root tolerance to compacted soil and showed that it was strongly correlated (R = 0.91) to field tests. Fewer roots were able to penetrate compacted soil layers, causing an increase in root mass above the compacted layer and an overall reduction in root biomass. Excessive soil compaction can slow root tip advance up to 75%, making interception and infection by soil fimgal pathogens much more likely than in non- compacted soils (Huisman 1982; Alhnaras et al., 1988). Studies with bean and Fsp showed that inoculation reduced root ability to penetrate a compacted soil layer. Fusarium species are generally likely to infect the hypocotyl or more stationary parts of the root system because they do not germinate quickly and 21 usually exist largely as colonies epiphytic on the root surface (Allrnaras et al., 198 8). In compacted fields with Fusarium spores present, tillage increased bean seed yields by 40% and reduce disease severity scores from 3.4 to 1.4 over non- tilled fields (Harverson et al., 2005). Bean genotypes that perform well in the presence of Fsp inoculum also often demonstrate tolerance to other soil stresses. For example the resistant genotype NY-2114-12 was able to increase root growth above the point of soil compaction to a greater extent than susceptible cultivars (Miller, 1985). The resistant snap bean ‘FR266’ also has been shown to have high tolerance to compacted soil (Silbemagel, 1990). NY 2114-12 also had higher shoot and root growth under water deficit'as compared to susceptible cultivars (Miller 1985). Fusarium root rot resistant cultivars outperform susceptible cultivars under each of the environmental stresses listed above except flooding stress. Resistant cultivars were unable to overcome disease under low soil oxygen stress (Miller 1985; Burke and Miller, 1983). Methods of Control: Root rot severity in beans is related to environmental conditions that limit root growth. Cultural practices that allow the development of a vigorous root system are essential to reducing root rot, especially when resistant cultivars are unavailable. Thus, planting in warm soil and maintaining optimum soil water are important to establish vigorous root grth which allows plants to better withstand root rot. Irrigation is especially important in dry years because water stressed plants are more susceptible to root rot than non water stressed plants. 22 Increasing the spacing between plants within rows and between rows may reduce root rot severity by reducing competition among root systems for water and nutrients. Crop rotation: Crop rotation can be used to reduce Fsp populations in the soil. Soil population density (cfu/ g) of Fsp is highest in continuous bean production and rotation can reduce the amount of inoculum in the soil. Fsp is very immobile in the soil with 14 years of continuous corn, soybean or fallow having very little Fusarium (Sippell and Hall 1982b). Recommendations are generally for 4 year rotations to reduce inoculum. Rotation with small grains and alfalfa add residue to the soils and can reduce compaction, improve the soil structure and increase water holding capacity of the soil. However, residues must be incorporated at least one month before planting or phytotoxicity can result (Burke and Miller, 1983). Organic amendments can also increase overall microbial population in the soil which in some cases reduCe habitat and nutrient availability in soil which could reduce competitiveness of Fsp. Deep Tillage: Deep tillage of the soil has been very effective to reduce root rot severity in compacted soils. Both susceptible and resistant cultivars exhibit decreased root rot severity with deep tillage (Tan and Tu, 1995). Deep tillage in the presence of Fsp inoculum improves plant yields more so than in non- infested soil. A study by Miller and Burke (1986) found that susceptible cultivars showed a larger yield increase fiom tillage than resistant cultivars illustrating the importance of deep root growth to overcome infection. 23 Chemical Treatments: Seed treatments are generally not effective against Fusarium root rot (Sippell and Hall, 1982b). Soil fumigants such as chloropicrin and methyl bromide may suppress F usarium root rot but are economically and environmentally impractical (Burke and Miller, 1983). Genetics: Sources of Resistance: N203 (P1203958) is considered the best source of resistance to F usarium root rot in P. vulgaris. Although this accession does not exhibit complete immunity, it does have a high degree of resistance. N203 is a late maturing, indeterminate accession collected in Mexico (Wallace and Wilkinson 1965). P. coccineus, also a primary source of resistance has shown slightly higher levels of resistance that N203 (Wallace and Wilkinson 1965). Both principle known sources of resistance, N203 and P. coccineus, have been reported to exhibit higher levels of resistance in the field as compared to the greenhouse (Wallace and Wilkinson, 1965; Baggett and Frazier, 1959), suggesting root avoidance or escape of infection may be an important component to resistance in the field. The first F usarium root rot resistant dry bean cultivars were released in 1974. They included the pink cultivars ‘Viva’, and ‘Roza’ and the small red cultivar ‘Rufus’ (Boomstra and Bliss, 1977). Each cultivar derived resistance from N203 (Burke, 1982). The first snap bean cultivars with resistance to Fusarium root rot were released in 1978. The source of resistance of these cultivars (RRR 77) and (RR 83) was not specified (Hagadom and Rand, 1978). Available genotypes with the highest level of resistance are late maturing with a 24 vine growth habit. It has been difficult to transfer resistance to early maturing short-vine and bush type beans (Burke and Miller, 1983). ‘FR266’ is a resistant snap bean with ‘Blue Lake’ pod quality characteristics released in 1987 which derived resistance from N203 and P. coccineus (Silbemagel, 1987). It is one of the few genotypes available with Fusarium root rot resistance and a bush type growth habit. Inheritance of Resistance: Bravo et al. (1969) found the resistance from N203 and P. coccineus to be inherited in a similar manner. They found resistance with either of the parents to be mostly dominant and controlled by 3 Or more genes (Bravo et al., 1969). Based on comparisons of the inheritance of resistance in N203 and 2114-12 (resistance from P. coccineus), Hassan et al. (1971) found N203 to possess 4 resistance genes and 2114-12 to posses 5-6 resistance genes, with 4 of the resistant genes in 2114— 12 also present in N203. Hassan et al. (1971) used a hypocotyl disease severity index to rate disease and did not consider root rot. Thus, their results may not be applicable to ‘root rot’. Hassan et a1 (1971) obtained similar results as Bravo et a1. (1969), who scored disease index on the roots. Overall there are many conflicting reports on the number of genes controlling resistance and it is clearly i . quantitatively inherited. The exact number of genes controlling the trait remains unclear. Numerous conflicting reports on the inheritance of resistance to Fsp suggest that growth and environmental conditiOns influence study results (Boomstra and Bliss 1977). 25 Schneider et al. (2001) observed while studying a population of RILS derived from a cross between FR266 and Montcalm that there is no hypersensitive response in the resistant genotype, only a slowing of symptom development. In the same population QTL were found to be associated with resistance in the field or in the greenhouse but there was no overlap. In addition, no single QTL explained more than 15 % of phenotypic variation, indicating that many genes are involved. Intermediate heritability scores also indicated a high effect on the environment for this trait. In a recent QTL study by Roman-Aviles and Kelly (2005) an inbred backcross line population was used to reduce the background effects and isolate more thoroughly the genetic mechanism for resistance. In this study a total of 5 QTL on two linkage groups explained 73% of the phenotypic variation. Again, QTL identified in field were different than those in the greenhouse. This may be in part because greenhouse plants were scored after two weeks whereas in the field they were scored after flowering or at maturity, resulting in selection for early expressed genes for resistance. One QTL was linked to a marker by Schneider et a1 (2001) that was identified (F R266 x Montcalm) as important for resistance. Root Architecture and Disease: Root pathogen interactions are largely dependent onroot growth, as roots are typically much more mobile in the soil than fungal pathogens. Growth rates of root length can serve as an indicator for the potential for root pathogen contact (Huisman, 1982). Root length also serves to indicate how much area is available for water and nutrients in the presence of root rot (Burke and Barker, 1966). 26 Genotypic variability for root length has been shown to be associated with resistance in common bean. Under field conditions, in the presence of Fsp spores, the root rot resistant genotype FR266 had a root dry Weight twice that of the susceptible cultivar ‘Montcalm’, along with a greater length and surface area, indicating an overall larger root system (Roman-Aviles et al., 2004). In addition average root system diameter was negatively correlated with root rot severity (r2: 0.4) in RILS developed from a cross between Montcalm and FR266. The average number of adventitous in the top cm was positively correlated with root rot tolerance (r2=0.6) (Snapp et al., 2003). In a screening of 10 bean cultivars from different market classes and with varying susceptibility to Fusarium root rot the three most resistant cultivars also had the most adventitious root mass (Roman- Aviles et al., 2004). Adventitious roots are commonly considered a host plant response to compensate for the infection of taproots and lateral roots (Chatterj ee, 1958; Steadman et al., 1975). A few studies have examined the role of root system architecture in resistance to root pathogens. Vine decline of muskrnelon (Cucumis melo L.) is caused by a complex of soil borne fungi, the most prominent fungus, Monosporascus cannonballus Causes the root rot symptoms. As with Fusarium root'rot in bean, disease severity is strongly influenced by the environment and no highly resistant cultivars exist (Dias et al., 2002). Root morphology variability has been discovered in tolerant and susceptible cultivars infected with the pathogen. Tolerant cultivars had more vigorous root systems, including a greater total root length, more root tips, and a greater fine root length than susceptible 27 cultivars (Crosby et al., 2000). The highly resistant wild melon accession Pat81 had a larger root system and more lateral roots that a highly susceptible commercial cultivar ‘Amarillo Canario’ both in the presence and absence of fungal inoculum (Dias et al., 2002). Pat8l was used as a parent in a backcross breeding program, and progeny with larger root systems demonstrated root rot tolerance (Dias et al., 2004). Root and Shoot Growth Habit Coordination Common bean shoot architecture has been classified into four types: determinate (Type I), indeterminate upright (Type H), indeterminate prostrate (Type III), and vine (Type IV) (Kelly 2001). Limited observational studies support a coordinated grth habit of the root and shoot (Lynch and van Beem 1993; Kelly, 1998). ‘Carioca’ a type III plant with highly branched shoot architecture and little apical dominance also exhibited highly branched root architecture (Lynch and van Beem 1993). In the case of Fusarium root rot resistance, it has been much easier to develop resistance in indeterminate type cultivars than in determinate cultivars. Therefore, if root architecture is important for resistance to Fusarium root rot, it may be strongly influenced by shoot architecture traits such as total leaf area. To effectively breed determinate Fusarium resistant bean cultivars an understanding of root and shoot traits is required to detennine how to separate them. 28 Summary: A half century of research has shown that Fusarium root rot severity in common bean is strongly influenced by environmental conditions that affect plant root growth. In addition all known sources of resistance provide partial resistance to infection. These factors of disease have indicated that chemical resistance is not a reliable selection tool as seen with differences between genetic control of field and greenhouse resistance. Breeding of cultivars resistant to Fsp is typically done in combination with breeding for general tolerance to stress. Only within the last few years have researches began to study the relationship between root architecture and F usarium root rot. In that time, it has been shown that indeed, the resistant cultivars do have larger root systems in field conditions than susceptible cultivars. Much more research needs to be conducted, however to determine the genetic control of this relationship and what root traits may serve as selection tools to facilitate breeding efforts that can provide cultivars that are less susceptible to Fusarium root rot in the presence of environmental stresses. 29 References: 2004. The Phosphorus Cycle [Online]. Available by Potash and Phosphate Institute http://msucares.com/crops/soils/phosphorus.html. Abawi, GS, and MA. Pastor Corrales. 1990. Root rots of beans in Latin America and Africa: Diagnosis, research methodologies, and management strategies CIAT, Cali, Colombia. Allmaras, R.R., J .M. Kraft, and DE. Miller. 1988. Effects of Soil Compaction and Incorporated Crop Residue on Root Health. Annual Review of Phytopathology 26:219-243. Araujo, A.P., M.G. Teixeira, and BL. de Alrneida. 1998. Variability of traits associated with phosphorus efficiency in wild and cultivated genotypes of common bean. Plant and Soil 203: 173-182. Asady, G.H., A.J.M. Smucker, and M.W. Adams. 1985. Seedling Test for the Quantitative Measurement of Root Tolerances to Compacted Soil. Crop Science 25:802-806. Austin, DR 2002. Roots as a Source of Food, p. 1025-1043, In Y. Waisel, Eshel, A., and Kafltafi U., ed. Plant Roots The Hidden Half, 3rd ed. Marcel Dekker, Inc, New York. Baggett, J .R., and WA. Frazier. 1959. Disease Resistance in the Runner Bean, Phaseolus coccineus L. Plant Disease Reporter 43: 137-143. Beebe, S.E., M. Rojas-Pierce, X.L. Yan, M.W. Blair, F. Pedraza, F. Munoz, J. Tohme, and JP. Lynch. 2006. Quantitative trait loci for root architecture traits . correlated with phosphorus acquisition in common bean. Crop Science 46:413- 423. Bell, D.L., and SE. Sultan. 1999. Dynamic phenotypic plasticity for root growth in Polygonum: A comparative study. American Journal of Botany 86:807-819. Bolland, M.D.A., and RH. Paynter. 1990. Increasing Phosphorus Concentration in Seed of Annual Pasture Legume Species Increases Herbage and Seed Yields. Plant and Soil 125: 197-205. Bonser, A.M., J. Lynch, and S. Snapp. 1996. Effect of phosphorus deficiency on growth angle of basal roots in Phaseolus vulgaris. New Phytologist 132:281-288. Boomstra, AG, and FA. Bliss. 1977. Inheritance of Resistance to F usarz’um? Solam' F Sp Phaseoli in Beans (Phaseolus vulgaris L) and Breeding Strategy to 30 Transfer Resistance. Journal Of the American Society for Horticultural Science 102:186-188. Bravo, A., D.H. Wallace, and Wilkinso.Re. 1969. Inheritance of Resistance to Fusarium Root Rot of Beans. Phytopathology 59: 1930-&. Bucher, M., C. Rausch, and P. Daram. 2001. Molecular and biochemical mechanisms of phosphorus uptake into plants. Journal of Plant Nutrition and Soil Science 164:209-217. Buerkert, A., and H. Marschner. 1992. Calcium and Temperature Effects on Seedling Exudation and Root-Rot Infection of Common Bean on an Acid Sandy Soil. Plant and Soil 1472293-303. Burke, D.W. 1968. Root Growth Obstructions and Fusarium Root Rot of Beans. Phytopathology 58:1575-&. Burke, D.W. 1982. Registration of Pink Beans Viva, Roza, and Gloria. Crop Science 22:684-684. Burke, D.W., and AW. Barker. 1966. Importance of Lateral Roots in Fusarium Root Rot of Beans. Phytopathology 56:292-&. Burke, D.W., and J.M. Kraft. 1974. Responses of Beans and Peas to Root Pathogens Accumulated During Monoculture of Each Crop Species. Phytopathology 64:546-549. Burke, D.W., and DE. Miller. 1983. Control of Fusarium Root-Rot with Resistant Beans and Cultural Management. Plant Disease 67:1312-1317. Burke, D.W., D.E. Miller, and AW. Barker. 1980. Effects of Soil-Temperature on Grth of Beans in Relation to Soil Compaction and Fusarium Root-Rot. Phytopathology 70: 1047-1049. Cardoso, I.M., and T.W. Kuyper. 2006. Mycorrhizas and tropical soil fertility. Agriculture Ecosystems & Environment 116172-84. Cardoso, I.M., C.L. Boddington, B.H. Janssen, O. Oenema, and T.W. Kuyper. 2006. Differential access to phosphorus pools of an oxisol by mycorrhizal and nonmycorrhizal maize. Communications in Soil Science and Plant Analysis 37:1537-1551. * Chatterjee, P. 1958. The Bean Root Rot Complex in Idaho. Phytopathology 48:197-200. 31 Christiansen, 1., and PH. Graham. 2002. Variation in di-nitrogen fixation among Andean bean (Phaseolus vulgaris L.) genotypes grown at low and high levels of phosphorus supply. Field Crops Research 73:133-142. Christou, T., and WC. Snyder. 1962. Penetration and Host-Parasite Relationships of Fusarium—Solani-F-Phaseoli in Bean Plant. Phytopathology 52:219-&. I Clark, L.J., W.R. Whalley, and PB. Barraclough. 2003. How do roots penetrate strong soil? Plant and Soil 255:93-104. Crosby, K., D. Wolff, and M. Miller. 2000. Comparisons of root morphology in susceptible and tolerant melon cultivars before and after infection by Monosporascus cannonballus. Hortscience 35 :68 l -683. da Silveira, A.P.D., and E. Cardoso. 2004. Arbuscular mycorrhiza and kinetic parameters of phosphorus absorption by bean plants. Scientia Agricola 61 :203- 209. . Dias, R.D.S., B. Pico, A. Espinos, and F. Nuez. 2004. Resistance to melon vine - decline derived from Cucumis melo ssp agrestis: genetic analysis of root structure and root response. Plant Breeding 123:66-72. Dias, R.D.S., B. Pico, J. Herraiz, A. Espinos, and F. Nuez. 2002. Modifying root . structure of cultivated muskmelon to improve vine decline resistance. Hortscience 37: 1092-1097. Dryden, P., and N.K. Vanalfen. 1984. Soil-Moisture, Root-System Density, and Infection of Roots of Pinto Beans by F usarz'um-Solani 1F Sp Phaseoli under Dryland Conditions. Phytopathology 74: 132-135. Fita, A., B. Pico, and F. Nuez. 2006. Implications of the genetics of root structure in melon breeding. Journal of the American Society for Horticultural Science 131:372-379. . Fitter, A. 2002. Characteristics and Functions of Root Systems, p. 15-32, In Y. Waisel, Eshel, A., and Kafltafi U., ed. Plant Roots The Hidden Half, 3rd ed. Marcel Dekker, Inc., New York. Fitter, A., L. Williamson, B. Linkohr, and O. Leyser. 2002. Root system architecture determines fitness in an Arabidopsis mutant in competition for immobile phosphate ions but not for nitrate ions. Proceedings of the Royal Society of London Series B-Biological Sciences 269:2017-2022. Gahoonia, TS, and NE. Nielsen. 2003. Phosphorus (P) uptake and growth of a root hairless barley mutant (bald root barley, brb) and wild type in low- and high- P soils. Plant Cell and Environment 26:1759-1766. 32 Gahoonia, TS, and NE. Nielsen. 2004. Root traits as tools for creating . phosphorus efficient crop varieties. Plant and Soil 260:47-57. Gerald, J.N.F., M.D. Lehti-Shiu, P.A. Ingram, K.I. Deak, T. Biesiada, and J .E. Malamy. 2006. Identification of quantitative trait loci that regulate Arabidopsis root system size and plasticity. Genetics 172:485-498. Gilbert, G.A., J.D. Knight, C.P. Vance, and D.L. Allan. 1999. Acid phosphatase activity in phosphorus-deficient white lupin roots. Plant Cell and Environment 22 : 8O 1 -8 10. Grimes, D.W., R.J. Miller, and PL. Wiley. 1975. Cotton and Corn Root Development in Two Field Soils of Different Strength Characteristics. Agronomy Journal 67 :5 19-523. Hagedom, DJ, and RE. Rand. 1978. Bean Breeding Lines Wisconsin (Rrr) 77 and Wisconsin (Rrr) 83. Hortscience 13:202-202. Harveson, R.M., J .A. Smith, and W.W. Stroup. 2005. Improving root health and yield of dry beans in the Nebraska panhandle with a new technique for reducing soil compaction. Plant Disease 89:279-284. Hassan, A.A., D.H. Wallace, and Wilkinso.Re. 1971. Genetics and Heritability of Resistance to F usarium-Solani F Phaseolr' in Beans. Journal of the American Society for Horticultural Science 96:623-&. Hedley, M.J., G.J.D. Kirk, and MB. Santos. 1994. Phosphorus Efficiency and the Forms of Soil-Phosphorus Utilized by Upland Rice Cultivars. Plant and Soil 1 58 : 53-62. Hedley, M.J., J.J. Mortvedt, N.S. Bolan, and J .K. Syers. 1995. Phosphorus Fertility Management in Ecosystems, p. 59-92, In H. Tiessen, ed. Phosphorus in the Global Environment. John Wiley and Sons, New York. Holford, I.C.R. 1997. Soil phosphorus: its measurement, and its uptake by plants. Australian Journal of Soil Research 35:227-235. Huisman, CC. 1982. Interrelations of Root-Growth Dynamics to Epidemiology of Root-lnvading Fungi. Annual Review of Phytopathology 20:303-327. Johnson, W.C., LE. Jackson, 0. Ochoa, R. van Wijk, J. Peleman, D.A. St Clair, and R.W. Michehnore. 2000. Lettuce, a shallow-rooted crop, and Lactuca serriola, its wild progenitor, differ at QTL determining root architecture and deep soil water exploitation. Theoretical and Applied Genetics 10111066-1073. 33 Jungk, A. 2001. Root hairs and the acquisition of plant nutrients from soil. Journal of Plant- Nutrition and Soil Science-Zeitschrift Fur Pflanzenemahrung Und Bodenkunde 164:121-129. Keenan, J.G., H.D. Moore, N. Oshima, and LE. Jenkins. 1974. Effect of Bean Root-Rot on Dryland Pinto Bean Production in Southwestern Colorado. Plant Disease Reporter 58:890-892. Kelly, J .D. 1998. Bean roots — A plant breeder’s perspective. Bean Improvement Cooperative Annual Report 41 :214-215. Kelly, J .D. 2001. Remaking bean plant architecture for efficient production. Advances in Agronomy 71 :109-143 Laperche, A., F. Devienne-Barret, O. Maury, J. Le Gouis, and B. Ney. 2006. A . simplified conceptual model of carbon/nitrogen functioning for QTL analysis of winter wheat adaptation to nitrogen deficiency. Theoretical and Applied Genetics 1 13:1 131- 1 146. Li, Y.D., Y.J. Wang, Y.P. Tong, J .G. Gao, J.S. Zhang, and S.Y. Chen. 2005. QTL mapping of phosphorus deficiency tolerance in soybean (Glycine max L. Merr.). Euphytica 142:137-142. Liao, H., G. Rubio, X.L. Yan, A.Q. Cao, K.M. Brown, and JP. Lynch. 2001. Effect of phosphorus availability on basal root shallowness in common bean. Plant and Soil 232:69-79. Liao, H., X. L. Yan, G. Rubio, S. E. Beebe, M. W. Blair, and J. P. Lynch. 2004. Genetic mapping of basal root gravitropism and phosphorus acquisition efficiency in common bean. Functional Plant Biology 31 9-.59 970. Liu, Y., G.H. Mi, R]. Chen, J .H. Zhang, and ES. Zhang. 2004. Rhizosphere effect and root growth of two maize (Zea mays L.) genotypes with contrasting P efficiency at low P availability. Plant Science 167:217-223. Lopez-Bucio, J., A. Cruz-Rarnirez, and L. Herrera-Estrella. 2003. The role of nutrient availability in regulating root architecture. Current Opinion 1n Plant Biology 6: 280- 287. Lynch, J. 1995. Root Architecture and Plant Productivity. Plant Physiology 109. 7- 13. Lynch, 1., and J .J . Vanbeem. 1993. Growth and Architecture of Seedling Roots of Common Bean Genotypes. Crop Science 33:1253-1257. 34 Lynch, JP, and SE. Beebe. 1995. Adaptation of Beans (Phaseolus- Vulgaris L) to Low Phosphorus Availability. Hortscience 30:1165-1171. Lynch, J .P., and KM. Brown. 2001. Topsoil foraging - an architectural adaptation of plants to low phosphorus availability. Plant and Soil 237:225-23 7. Lynch, J .P., and MD. Ho. 2005. Rhizoeconomics: Carbon costs of phosphorus acquisition. Plant and Soil 269245-56. Lynch, J., A. Lauchli, and E. Epstein. 1991. Vegetative Growth of the Common Bean in Response to Phosphorus-Nutrition. Crop Science 3 1 :380-3 87. Malamy, J.E. 2005. Intrinsic and environmental response pathways that regulate root system architecture. Plant Cell and Environment 28:67-77. Marschner, H. 1995. Mineral Nutrition of Higher Plants. 2nd ed. Academic Press, New York. Merrill, S.D., D.L. Tanaka, and J .D. Hanson. 2002. Root length grth of eight crop species in haplustoll soils. Soil Science Society of America Journal 66:913- 923. Miller, DE, and D.W. Burke. 1977. Effect of Temporary Excessive Wetting on Soil Aeration and Fusarium Root-Rot of Beans. Plant Disease Reporter 61: 175- 179. Miller, DE, and D.W. Burke. 1985. Effects of Soil Physical Factors on Resistance in Beans to Fusarium Root-Rot. Plant Disease 69:324-327. Miller, DE, and D.W. Burke. 1986. Reduction of F usarirun Root-Rot and Sclerotinia Wilt in Beans with Irrigation, Tillage, and Bean Genotype. Plant ‘ Disease 70: 163-166. Miller, OR, I. Ochoa, K.L. Nielsen, D. Beck, and J .P. Lynch. 2003. Genetic variation for adventitious rooting in response to low phosphorus availability: potential utility for phosphorus acquisition fiom stratified soils. Functional Plant Biology 30:973-985. Nanamori, M., T. Shinano, J. Wasaki, T. Yamarnura, I.M. Rao, and M. Osaki. 2004. Low phosphorus tolerance mechanisms: Phosphorus recycling and photosynthate partitioning in the tropical forage grass, Brachiaria hybrid cultivar mulato compared with rice. Plant and Cell Physiology 45:460-469. Nielsen, K.L., A. Eshel, and JP. Lynch. 2001. The effect of phosphorus availability on the carbon economy of contrasting common bean (Phaseolus vulgaris L.) genotypes. Journal of Experimental Botany 52:329-339. 35 Nuruzzarnan, M., H. Lambers, M.D.A. Bolland, and E.J. Veneklaas. 2006. Distribution of carboxylates and acid phosphatase and depletion of different phosphorus fractions in the rhizosphere of a cereal and three grain legumes. Plant and Soil 281:109-120. Reddick, D. 1917. Effect of soil temperature on the growth of bean plants and on their susceptibility to a root parasite. American Journal of Botany 4:513-519. Richardson, AB. 1994. Soil Microorganisms and phosphorus availability, p. 50- 62, In C. Pankhurst, et al., eds. Soil biota management in sustainable farming systems. CSIRO Publishing, Melbourne. Roman-Aviles, B., and JD. Kelly. 2005. Identification of quantitative trait loci conditioning resistance to F usarium root rot in common bean. Crop Science 45:1881-1890. ’ Roman-Aviles, B., S.S. Snapp, and J .D. Kelly. 2004. Assessing root traits associated with root rot resistance in common bean. Field Crops Research 86: 147- 156. Rubio, G., H. Liao, X.L. Yan, and JP. Lynch. 2003. Topsoil foraging and its role in plant competitiveness for phosphorus in common bean. Crop Science 43:598- 607. Saettler, AW. 1982. Bean Disease and Their Control, In R. D. Robertson, ed. Dry Bean Production: Principles and Practices. MSU Cooperative Extension, East Lansing, MI. Sanchez, RA, and JG. Salinas. 1981. Low-input technology for managing Oxisols and Utisols in tropical America. Advances in Agronomy 34:280-406. Sanchez-Calderon, L., J. Lopez-Bucio, A. Chacon-Lopez, A. Gutierrez-Ortega, E. Hemandez-Abreu, and L. Herrera-Estrella. 2006. Characterization of low phosphorus insensitive mutants reveals a crosstalk between low phosphorus- induced determinate root development and the activation of genes involved in the adaptation of Arabidopsis to phosphorus deficiency. Plant Physiology 140:879- 889. Schachtrnan, D.P., R.J. Reid, and SM. Ayling. 1998. Phosphorus uptake by plants: From soil to cell. Plant Physiology 116:447-453. Schneider, K.A., K.F. Grafton, and JD. Kelly. 2001. QTL analysis of resistance to Fusarium root rot in bean. Crop Science 41 :535-542. 36 Schwartz, HF. 2006. Root Rot of Dry Beans [Online]. Available by Colorado State University Cooperative Extension http://www.ext.colostate.edu/pubs/crops/02938.htrnl (posted June 2006). Shen, H., X.L. Yan, M. Zhao, S.L. Zheng, and X.R. Wang. 2002. Exudation of organic acids in common bean as related to mobilization of aluminum- and iron- bound phosphates. Environmental and Experimental Botany 4811-9. Silbemagel, M.J. 1987. Fusarium Root Rot-Resistant Snap Bean Breeding Line Fr-266. Hortscience 22: 1 337-1338. Silbemagel, M.J., and LI. Mills. 1990. Genetic and Cultural-Control of Fusarium Root-Rot in Bush Snap Beans. Plant Disease 74:61-66. Singh, S.P., H. Teran, C.G. Munoz, J.M. Osorno, J .C. Takegami, and M.D.T. Thung. 2003. Low soil fertility tolerance in landraces and improved common bean genotypes. Crop Science 43:110-119. Sippell, D.W., and R. Hall. 1982a. Effects of pathogen species, inoculum concentration, temperature, and soil moisture on bean root rot and plant growth. Canadian Journal of Plant Pathology 4:1-7. Sippell, D.W., and R. Hall. 1982b. Effects of F usarium solani phaseoli, Pythium ultimum, and F. oxysporum on yield components of white bean. Canadian Journal of Plant Pathology 4:54-58. Snapp, SS, and J .P. Lynch. 1996. Phosphorus distribution and remobilization in bean plants as influenced by phosphorus nutrition. Crop Science 36:929-935. Snapp, S., W. Kirk, B. Roman-Aviles, and J. Kelly. 2003. Root traits play a role in integrated management of Fusarium root rot in snap beans. Hortscience 38: l 87- 191. Steadman, J.R., E.D. Kerr, and RE. Murnm. 1975. Root-Rot of Bean in Nebraska - Primary Pathogen and Yield Loss Appraisal. Plant Disease Reporter 59:305-308. Sultan, SE. 1987. Evolutionary Implications of Phenotypic Plasticity in Plants. Evolutionary Biology 21 :127-178. Tan, CS, and J .C. Tu. 1995. Tillage Effect on Root-Rot Severity, Growth and Yield of Beans. Canadian Journal of Plant Science 75:183-186. Tinker, P.B., M.D. Jones, and D.M. Durall. 1992. A functional comparison of ecto- and endomycorrhizas, p. 303-310, In D. J. Read, eta1., eds. Mycorrhizas in Ecosystems. CAB International, Wellingford, UK. 37 Tuberosa, R., M.C. Sanguineti, P. Landi,‘M. Michela Giuliani, S. Salvi, and S. Conti. 2002. Identification of QTLS for root characteristics in maize grown in hydroponics and analysis of their overlap with QTLS for grain yield in the field at two water regimes. Plant Molecular Biology 48:697-712. Vogelilange, R., U. Mohr, A. Wiemken, M. Duggelin, R. Guggenheim, and T. Boller. 1995. Developmental Regulation of Susceptibility and Tolerance to Fusarium Root-Rot in Beans. Botanica Acta 108:387-395. Wallace, DH, and Wilkinso.Re. 1965. Breeding for Fusarium Root Rot Resistance in Beans. Phytopathology 55:1227. Wang, L.D., H. Liao, X.L. Yan, B.C. Zhuang, and Y.S. Dong. 2004. Genetic variability for root hair traits as related to phosphorus status in soybean. Plant and Soil 261 :77-84. Watt, M., and JR. Evans. 2003. Phosphorus acquisition from soil by white lupin (Lupinus albus L.) and soybean (Glycine max L.), species with contrasting root development. Plant and Soil 248:271-283. Whitney, DA. 1988‘. Phosphorus Facts Soil, Plant, and Fertilizer. Extension Bulletin. Kansas State University Cooperative Extension Service, Manhattan, Kanas. Wissuwa, M. 2003. How do plants achieve tolerance to phosphorus deficiency? Small causes with big effects. Plant Physiology 133:1947-1958. Wissuwa, M., and N. AC. 2001. Further characterization of two QTLS that increase phosphorus uptake of rice (Oryza sativa L.) under phosphorus deficiency. Plant and Soil 237:275-286. Wissuwa, M., and N. Ae. 2001. Genotypic variation for tolerance to phosphorus deficiency in rice and the potential for its exploitation in rice improvement. Plant Breeding 120:43-48. Wissuwa, M., M. Yano, and N. Ae. 1998. Mapping of QTLS for phosphorus- deficiency tolerance in rice (Oryza sativa L.). Theoretical and Applied Genetics 97:777-783. Wortrnann, C.S., R.A. Kirkby, C.A. Eledu, and DJ. Allen. 1998. Atlas ‘ of common bean (Phaseolus vulgaris L.) production in Africa CIAT, Cali, Colombia. Xiao, K., H. Katagi, M. Harrison, and Z.Y. Wang. 2006. Improved phosphorus acquisition and biomass production in Arabidopsis by transgenic expression of a purple acid phosphatase gene fi'om M. truncatula. Plant Science 170:191-202. 38 Yan, X.L., J .P. Lynch, and SE. Beebe. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .1. Vegetative Response. Crop Science 35:1086-1093. Yan, X.L., S.E. Beebe, and J .P. Lynch. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .2. Yield Response. Crop Science 35: 1094-1099. Yan, X.L., H. Liao, M.C. Trull, S.E. Beebe, and JP. Lynch. 2001. Induction of a major leaf acid phosphatase does not confer adaptation to low phosphorus availability in common bean. Plant Physiology 125:1901-1911. l Yan, X.L., H. Liao, S.E. Beebe, M.W. Blair, and JP. Lynch. 2004. QTL mapping of root hair and acid exudation traits and their relationship to phosphorus uptake in common bean. Plant and Soil 265:17-29. Youngdahl, L]. 1990. Differences in Phosphorus Efficiency in Bean Genotypes. Journal of Plant Nutrition 13:1381-1392. Yu, L.X., J.D. Ray, J.C. Otoole, and HT. Nguyen. 1995. Use of Wax-Petrolatum Layers for Screening Rice Root Penetration. Crop Science 35:684-687. Zobel, R.W. 1991. Root Growth and Development, p. 61-71, In P. B. Keister, ed. The Rhizosphere and Plant Growth. Kluwer Academic Publishers, Boston. 39 Figure 1. Responses of Arabidopsis root systems to supply of nutrients P, N, and S. Figure reproduced from Lopes-Bucio et al. (2003). 40 Soil phosphorus cycl HARVEST _ H-Iliiiii' I.Ill|l'|-li a H‘ :1, lillr‘.;:idl' ll'lllll ‘l'l‘v UHHIIH ‘5. l triuiiili ’ ‘ BIOITIE—‘SS it, Uptake A' , *4,” I". ‘ Weathering P fixation Figure 2: The Phosphonrs Cycle (Potash and Phosphate Institute, 2004) 41 s» lnltlel leelone i on root: and Older lesions and general discoloration. Chlamydoepore “ {a Chlamydoeporee i germinatee end ' V " it" « form In Infected . penetrates nearby = ‘ m roots and root and hypocotyl .‘ . . , ......... .. . _ Figure 3: Fusarium root rot infection in beans (Schwartz, 2006) 42 Chapter 2: Fusarium Root Rot Incidence and Root System Architecture in Grafted Common Bean Lines Abstract There is minimal understanding of the complex genotype by environment interaction controlling Fusarium root rot (caused by F usarium solani f.sp. phaseoli) disease expression in common bean (Phaseolus vulgaris L.). Severity of the disease is increased by environmental factors such as soil compaction that add additional stress to the plant root system. It has proven difficult to develop resistant cultivars, especially with a bush growth habit as seen in snap and kidney. bean market classes. One resistant determinate snap bean line, ‘FR266’, has been developed. The current study used reciprocal grafting techniques with resistant ‘FR266’ and susceptible ‘Montcalm’ genotypes to determine if the genetic control of resistance lies in the root genotype. The influence of a compacted layer on the root and shoot genotype and root rot resistance was studied. Root rot resistance was found generally to be controlled by the root genotype. However, in the presence of a compacted layer, the root and the shoot genotype play a part in the resistance. Root traits were also observed under these conditions and root mass and root diameter Were shown to be controlled by the root genotype without compaction but by the root and shoot genotype when a compacted. layer was present. 43 Introduction Root rot of common bean (Phaseolus vulgaris L.) is caused by a combination of fungi and environmental stress and is a major limiting factor in production (Abawi and Pastor Corrales 1990). The complex of pathogens believed to contribute to root rot include: F usarium solani f.sp. phaseoli, Rhizoctonia solani Kuhn, Pythium ultimum, flielaviopsis basicola (Berk. & Br.) Ferr. and Aphanomyces euteiches f.sp. phaseoli. In the majority of dry bean and snap bean producing regiOns of the United States, F. solani f.sp. phaseoli (Fsp) is considered the major pathogen and severe infections have caused yield losses of up to 84% (Chatterjee, 1958; Keenan et al., 1974; Burke and Kraft, 1974; Steadman etal., 1975; Saettler, 1982). I Symptoms Of Fusarium root rot include red longitudinal lesions on the hypocotyls, taproot and lateral roots of the bean plants which lead to rot of infected roots. Severity of infection increases as the plant develops, and with time complete rotting of the root system can occur and above ground symptoms develop including leaf chlorosis, stunting of stems, and reduction in pod production (Schneider and Kelly, 2000; Burke and Barker, 1966). Under optimal growing conditions, Fsp is a minor pathogen of beans. Environmental conditions such as exocssive soil compaction can cause plant stress and constrain optimal root development and enhance Fusarium root rot development (Burke, 1968). Soil compaction is an increase in soil bulk density and a decrease in air filled pores as a result of force applied to the soil; excessive soil compaction impedes root grth and function (Alhnaras et al., 1988). Soil 44 compaction is frequently present in agricultural fields as a result of the frequent passage of heavy machinery over soil used for planting, harvest, and other crop maintenance activities (Harverson et al., 2005). Excessive soil compaction can slow root tip advance up to 75%, making interception and infection by soil fungal pathogens much more likely than in non compacted soils (Huisman, 1982, Allrnaras et al., 1988). In compacted soil in the presence of inoculum of Fsp, tillage enhanced bean seed yields by 40% and reduced disease severity scores from 3.4 to 1.4 (on a scale of 0-4, where 0 indicates no disease and 4 indicates more than 75% of the root system is rotted) over non tilled fields (Harverson et al., .2005) Bean genotypes with the highest level of Fusarium root rot resistance are late maturing with an indeterminate plant growth habit. It has been difficult to transfer resistance to beans with a determinate bush type plant growth habit (Burke and Miller, 1983). FR266, a resistant snap bean was released in 1987 and derived resistance from N203 a late manning, indeterminate accession collected in Mexico (Wallace and Wilkinson, 1965) and Phaseolus coccineus (Silbemagel 1987). FR266 has no hypersensitive response, only a slowing of symptom development. Schneider et a1 (2001) investigated QTL for resistance in a population of recombinant inbred lines derived from a cross between FR266 and a highlysusceptible kidney bean variety ‘Montcalrn’. QTL were identified, however, no single QTL explained more than 15% of the phenotypic variation and different QTL were identified in greenhouse as compared to field. tests. Intermediate heritability scores also indicated a high effect of the environment for 45 this trait. Breeders have typically not relied only on root rot resistance scores from greenhouse screening alone, but have combined root rot scoring with field screening for stress tolerance that involved compaction and drought interactions (Burke and Miller, 1983). The effect of environmental factors such as compaction (Miller and Burke, 1985) on disease severity and the necessity to develop stress tolerant plants in general raised the question, what traits besides biochemical resistance may be responsible for the resistance seen in the cultivar FR266. Characteristics such as root system architecture (Snapp et al., 2003) may be involved. Under field conditions, in the presence of Fsp chlarnydospores, FR266 had a root dry weight twice that of Montcalm, along with a greater length and surface area, indicating an overall larger root system (Roman-Aviles et al., 2004). The average number of lateral roots in the top 1.0 cm of soil was positively correlated (r2=0.6) with root rot tolerance in snap beans (Snapp et al., 2003). Lateral root density per volume of soil may be predictive of stress tolerance rather than resistance to Fsp and has been associated with drought tolerance in lettuce (Johnson et al., 2000). FR266 also has been shown to grow well in compacted soil (Silbemagel, 1990). Previous studies on compaction tolerance in corn. (Zea mays L.) and soybean (Glycine max L.) have identified cultivar specific variability in the ability of roots to grow through compacted layers (Busharnuka and Zobel, 1998). These findings suggest that in compacted soils biochemical resistance to a pathogen is not the only factor required for a plant to grow well. 46 The difficulty in introducing root rot resistance into Andean beans, has led to the hypothesis that there are inherent physiological factors in F usarium root rot resistant bean cultivars that are expressed only under field conditions. Root and shoot growth is genetically integrated, therefore determining if a trait such as F usarium root rot resistance is expressed in the root, the shoot, or the whole plant is difficult to discern in whole plant or hybridization studies, but can be differentiated by interchanging root and shoot genotypes via grafting (White and Castillo, 1989; Izquierdo and Hosfield, 1982). The grafting technique has been demonstrated to be a sound tool for genetic research in common bean. Research has demonstrated the technique is free of artifacts by employing the self graft as a control for the technique (White and Castillo, 1989; Izquierdo and Hosfield, 1982). Using the reciprocal grafting technique to study drought tolerance in common bean, the root genotype, or more specifically, genes expressed only in the root system, had a large effect on seed yield under drought conditions. The shoot genotype had some effect on the trait, but much less than that of the root (White and Castillo, 1989; White and Castillo, 1992). A study of resistance to the soil borne pathogen, Phialophora gregata which causes brown stem rot in soybean was also found to be controlled by the root system (Bachman and Nickell, 1999). Pantalone et a1. (1999) found that the root system architecture trait “fibrous-like rooting” was controlled by the root system and enhanced seed biomass when grafted onto some soybean genotypes. In addition there is evidence that the shoot genotype influences the rOot system 47 phenotype as seen with nodule development and N fixation in soybean (Sheng and Harper, 1997). The objectives of this research were to determine a) if the expression of Fusarium root rot resistance was mediated by root or shoot factors and if specific root architectural traits were associated with resistance in reciprocally grafted Fusarium root rot resistant and susceptible bean lines and to evaluate b) how root rot resistance and root architectural traits were mediated by root and shoot factors when reciprocally grafted plants were grown with the added stress of soil compaction. Material and Methods Plant material Two common bean genotypes with differences in susceptibility to. Fusarium root rot were used in this study. They were the snap bean line ‘FR266’ characterized as root rot resistant (Silbemagel, 1987) andthe kidney bean ‘Montcalm’ characterized as root rot susceptible (Schneider et al., 2001). FR266 was developed by USDA/ARS in Prosser, WA (Silbemagel, 1987). Montcalm is representative of commercial kidney bean types and was developed at Michigan State University (Copeland and Erdmann, 1977). Select characteristics of FR266 and MOntcalm are described in Table 1. Reciprocal Grafting In all experiments reciprocal grafting was employed to determine if genetic control of Fusarium root rot resistance originated from the shoot or root system. Plants were grafted in one of four combinations. The first combination 48 was F R266 shoot grafted onto Montcalm root and the second combination was Montcalm Shoot grafted onto FR266 root. The third and fourth graft types were self grafts of FR266 shoot grafted onto FR266 root and Montcalm shoot grafted onto Montcalm root, respectively The self grafts were used as controls to determine the effect of grafting on root rot and plant grth and these were compared to ungrafted FR266 and ungrafted Montcalm plants alSo included in all experiments. Seeds were surface sterilized by soaking 2 min in 70% ethanol, followed by 2 min in 0.5% NaOCl. Seeds were rinsed with distilled water 4 times and incubated at 18°C in the dark in moist germination paper for 2 days. Seeds were planted as described below for each experiment. Five days after germination, seedlings were grafted using the cleft graft technique (Izquierdo and Hosfield, 1982). A razor blade was used to cut through the stem 1.5 cm above the cotyledon. The bottom 1.5 cm of the stem of the shoot was cut into a wedge shape with a razor blade by removing a thin slice of the epidermis on 2 sides of the stern. A longitudinal cut of 1.5 cm was made into the top part of the stern of the root System. The wedge shaped base of the shoot’s stem was placed into the stem above the root system. The shoot and root were held together with parafilm. Grafted plants were kept in a controlled environment at a temperature of 25°C day and a relative humidity of 90%. Plants received 14 hr of light (350-450 mE). Plants were misted with water by hand with a spray bottle every 4 hours until callus tissue formed (5-7 days) and water could be supplied through the root system. 49 Experiments: Two distinct controlled environment plant growth experiments were conducted and each experiment was repeated twice. All experiments were set up as randomized complete block designs with two treatments, inoculation with Fusarium solani spores and no inoculation. There were 6 plant genotypes grown under each treatment, including each of the 4 graft types described above and ungrafted plants of FR266 and Montcalm. There were four replications of each treatment/ genotype combination. Experiment 1 Non-compacted soil: Media: Sand harvested from Sandhill Farm, Michigan State University, East I Lansing, MI was sieved, sterilized via autoclaving, and mixed with perlite (60:40) by volume. Containers: Seeds were planted into 72 cell flats. Five day old plants were grafted as described above. One week following grafting, plants were transplanted into containers referred to as root observation boxes. These containers promoted the plant’s root system to grow flat and therefore improved the ability to harvest roots for measurement. The containers were constructed of glass panes 21.6 cm by 27.9 cm separated by 5.1 cm wooden strips for a dimension of 5.1 x 21.6 x 27.9 cm (Figure 1a). Plant growth conditions: Root observation boxes were completely covered with aluminum foil to prevent light from interfering with root growth. Individual root observation boxes were placed together into the growth chamber with other boxes, each at a 45° angle. Plants were exposed to 14 h of light (350-450 mE) at 28°C 50 and 10 h dark cycle at 25°C. Plants were watered as needed and were fertilized weekly with ‘/2 strength Hoagland’s solution (Hoagland and Arnon, 1938). Fusarium spore preparation and inoculation: Fourteen day old plants were inoculated with 15 ml of 2.0 x 105 conidia ml'l suspension of a mix of 2 isolates of F. solani f. Sp. phaseoli (Fsp). The Hawks 2B isolate was collected by Schneider and Kelly (2000) in Presque Isle County, MI and S-602 was received from the I Penn State Fusarium Research Center, University Park, PA. Potato dextrose agar (PDA, Invitrogen TM, Life Technologies, MD; 39 g L'1 water) was the media used to grow the Fusarium colonies. Inoculum discs, 4 mm diameter were cut from margins of 1- week old cultures grown on PDA and transferred mycelial side down, to the. center of individual Petri plates (15 m depth x 90 mm diameter) containing 25 ml of PDA. Plates were stored at room temperature for 2 weeks prior to inoculum preparation. The inoculum was prepared by scraping mycelium and conidia from PDA plates into distilled water, then strained through cheesecloth. The conidia] concentration was adjusted to 2 x 105 rrrl'1 using a hemocytometer. Fifteen ml of the inoculum was applied to the soil at the base of the plant. Harvest: Plants were harvested 32 days after planting and shoots were separated from roots and soil was washed from the root system with distilled water. The experiment was carried out twice, in August 2003 and in August 2004. 51 Experiment 2: Compacted soil layer: Media: The growing media was a Capac loam harvested from the Michigan State University Horticulture Farm, East Lansing, MI. soil was sieved, autoclaved, air dried and brought to 18% moisture before adding to containers. Containers: PVC tubes of 7.6 cm diameter and 0.5 cm wall thickness were used for plant grth containers based on the design of Asady et al (1985). The containers consisted of three layers. The top layer was 12.5 cm long, the middle layer was 2.6 cm long, and the bottom layer was 15 cm long (Figure lb). Soil was added to the top and bottom layers to achieve a bulk density of 1.1g/cm3. The middle layer was compacted to a bulk density of 1.7 g/cm3 with a hydraulic compactor machine. Bulk densities of each layer were determined based on soil weight. The bulk density of the compacted layer was chosen because the soil strength at that density reduces common bean root growth (Asady et al., 1985). The three layers of the container were held together with duct tape. Fusarium spore preparation and inoculation: The same mix of 2 isolates of F. solani f. sp. phaseoli (Fsp) were used in this experiment as described in Experiment 1, Fusarium spore preparation and inoculation. Colonies were first grown on potato dextrose agar (PDA, Invitrogen TM, Life Technologies, MD; 39 g L'1 water). After one week of growth, fusariurn inoculum discs, 4 mm diameter were cut from margins and transferred mycelial side down to sterile pearl millet seed prepared as follows. Millet seed (1000 ml) was poured in an aluminum pan (30.5 cm x 25.4 cm x 11.4 cm) mixed with 600 ml of distilled water, covered with aluminum foil and was kept at room temperature for 12 h. The pan was 52 autoclaved for 4 h and cooled 4 h. Next, 12 inoculum disks (1.3 cm wide) were evenly spaced over the seed. Two-week old cultures were used to prepare inoculum. Millet seed was mixed to homogeneity. Mycelium and conidia of a 1 gm sample of the seed was collected in distilled water. The inoculum concentration of the sample was determined with a hemocytometer. The weight of seed for a conidia] concentration of 2 x 105 ml'1 was determined. The inoculum/millet mixture was combined with the soil for the top layer of the container. Plant growth conditions: Surface sterilized pregerminated seeds were planted directly into containers, two seeds per container. After germination, seeds were thinned to one per container. Grafting, as described above, was conducted directly in containers. Plants were exposed to 14 h of light (350—450 mE) at 28°C and 10 h dark cycle at 25°C. Plants were watered as needed. Percent moisture of the soil was kept at 18% and the amount of water to add to each plant was determined by container weight. Plants were fertilized weekly with 1/2 strength Hoagland’s solution (Hoagland and Amon, 1938). Harvest: Thirty-two day old plants were harvested and shoots were separated fiom roots. Roots were harvested separately from each of the three layers of the containers and soil was washed from the roots with distilled water. The experiment was carried out once in November 2005. In October, 2004 the experiment was carried out with ungrafted and reciprocal grafts only. 53 Plant Measurements: Following harvest of plants in each experiment, root rot scores were measured based on a scale of l to 7 as described by Schneider et al (2000) where “1: healthy root system with no discoloration of root or hypocotyls tissue and no reduction in root mass compared to the uninoculated control; 2 =appearance of small reddish-brown lesions, 0.1-0.2 cm in length, at base of hypocotyls with size and appearance of root mass normal; 3 = increase in intensity and size and coalescing of localized root/hypocotyl lesion circling 3 180° of stem with lesions from 0.5-1cm and 10% to 20% discoloration on roots but no reduction in size of root mass; 4 =increasing intensity of discoloration and size of hypocotyl lesions with lesions becoming extended and completely encircling the stem; 5% to 10% reduction in root mass with 95% of the roots discolored, 5 = increasingly discolored and extended hypocotyls lesions. 100% of the roots intensely reddish- brown with a 20% to 50% reduction in root mass, 6 = hypocotyl lesions encircling stem become more extended (2 cm); root mass is intensely discolored and reduced from 50% to 80%; 7 = pithy or hollow hypocotyl with very extended lesions, where root mass is reduced from 80% to 100% and is functionally dead.” Following root rot scoring, cleaned roots were placed in a transparent plastic tray (22 x 28 x 6.5 cm) and covered with water. The tray containing roots was then scanned with a flatbed scanner with a top and bottom light source at 300 dpi to produce a 2 dimensional scanned image. The images were analyzed with WinRHIZO software (Regent Instruments Inc., Quebec, Canada). With this program root length, root surface area and average root diameter of the samples 54 was determined. Following image capture, roots and shoots were dried at 65°C for two days and dry weights were determined. Data Analysis: Statistical analysis was conducted with SAS for Windows V8 (SAS Institute, Cary, NC). The command PROC GLM was used to determine treatment, genotype and interaction effects. An initial model with 2 runs, 4 replications, 2 treatments, and 6 genotypes, was run for experiment 1 to determine the effect of Fusarium inoculation and grafting on plant growth. The same model was developed for experiment 2, but there was 1 run, instead of 2 runs in the model. A second model was then developed with 2 runs (1 run for experiment 2), 4 replications, 1 treatment (inoculated plants only) and 2 shoot genotypes (FR266 or Montcalm) and 2 root genotypes (FR266 or Montcalm. Ungrafted plants and uninoculated plants were excluded from this model. This model was used to determine the specific effects of each root and shoot genotype in the presence of Fusarimn inoculum with out confounding results with ungrafted plants. Least significant differences (LSD) values were used to determine significant differences among treatments. Results: Experiment 1: Consistent with earlier results (Schneider et al., 2001), ungrafted Montcalm plants grown in the presence of Fsp spores had a significantly higher root rot score than ungrafted FR266 plants (Table 2). No significant differences were observed between FR266 and Montcalm for root dry weight in the presence of Fsp inoculum, which is not unexpected based on length 55 of experiment. Root measurements were taken on 32 day old plants, and at this time root rot symptoms of root lesions and discoloration were visible, but a longer grth period would be required to observe differences in grth response. Similar results have been observed by other researchers (Roman-Aviles et al., 2004) Root system architectures of FR266 and Montcalm were characterized by root weight, length, and diameter. The most striking difference between FR266 and Montcalm root architecture was average root diameter. In both inoculated and uninoculated plants, FR266 had a greater average root diameter than Montcalm (Table 3). The reciprocal grafting technique was used with FR266 and Montcalm plants to understand genetic control of resistance to Fusarium root rot and root architecture. Self grafts of each genotype were included and compared to the ungrafted genotypes as controls to determine if the technique itself impacted reaction to Fusarium. The ungrafted and grafted plants did not differ in root rot scores (Table 2). The ungrafted plants exhibited larger root dry weights than their self grafted counterparts (Table 2). The comparison of root traits in self grafts and reciprocal grafts in the presence of Fsp inoculum showed that root rot score is controlled by the root 1 genotype (Table 4) and it was the F R266 root in any combination that exhibited the lower score (Tables 4 and 5). Root dry weight was also controlled by the root genotype (Table 4) and the FR266 root in any graft combination produced a larger root system that the Montcahn root (Table 5). Average root diameter was 56 controlled by the shoot genotype (Table 4). The FR266 shoot in any graft combination increased the average diameter of the root system (Table 5). For each of the traits described above, there were no significant root*shoot interactions, which allowed for the presentation of data averaged over root or shoot genotype as in Table 5. Experiment 2: Ungrafted Montcalm plants had higher root rot incidence than FR266 plants when they were grown in containers with a compacted soil layer (1.7 g /cm3 ) at 12.5 cm below the surface of the container (Table 6). Differences in root architecture between FR266 and Montcalm were observed in the soil below the compacted layer. In the presence of Fsp inoculum, F R266 plants had greater total root length and average root diameter than Montcalm plants (Table 7). Comparison of ungrafted and self grafted plants indicate that grafting did not affect root rot scores (Table 7). Grafting, however, reduced shoot and root dry weight of both FR266 and Montcahn plants (Table 7). Root rot incidence of grafted plants in this experiment was affected not only by the root genotype as it was in Experiment 1, but was affected by root and shoot genotype (Table 8). The graft combination FR266 root with Montcalm shoot had the same root rot incidence as the graft combination Montcalm root with FR266 shoot. The root architecture variables root dry weight, root length, and average root diameter also were influenced not by root or shoot genotype alone, but by the interaction of root and shoot genotypes (Table 8). The graft combination 57 Montcalm root/F R266 shoot, which had a different reaction to root rot in this experiment than in Experiment 2, also exhibited greater root dry weight and root length in the middle compacted and bottom layers of soil than the Montcahn self graft (Table 8). Ungrafted and self grafted Montcalm plants exhibited reductions in root length below the compacted soil layer in the inoculated treatment (Tables 7 and 8) not exhibited by the Montcalm root/FR266 shoot grafted plants (Table 8). The Montcalm root/F R266 shoot graft combination did not, however, have a larger average root diameter than the Montcalm self graft. Each of the four graft combinations had consistently similar root diameters in this experiment, except in the top layer where the Montcalm root/F R266 shoot combination had a lower average root diameter (Table 8). 9 Discussion: Experiments 1 and 2 showed that root rot incidence was controlled by genes expressed only in the root system in the absence of a compacted layer and by genes expressed throughout the plant in the presence of a compacted layer. These results demonstrated that different mechanisms of resistance to F sp were at work under different enviromnental conditions. I looked at a few basic root architectural traits in each experiment to determine their involvement in F sp resistance. One of the potential Fsp resistance mechanisms of FR266 is its thick roots (Snapp etal., 2003). The reciprocal grafting technique alloWed the testing of the hypothesis that thicker roots have less root rot incidence than thinner roots. In Experiment 1, it was found that the root genotype dictated root rot score, whereas 58 the shoot genotype dictated average root diameter. Since the graft combination Montcalm root/FR266 shoot had thicker roots than the Montcalm self graft, but did not have reduced root rot incidence, therefore root thickness was not a factor in root rot resistance under the experimental conditions. It would be interesting to determine if the same relationship holds true later in plant development when root rot symptoms are more severe. In Experiment 2, there were no clear differences in root diameter between the genotypes. In the compacted layer there was a uniform reduction of average root diameter of each graft combination, indicating both FR266 and Montcalm responded in the same way to compaction. Small diameter roots have the ability to grow through smaller pores in compacted soils than larger diameter roots (Bennie, 1991). This may be the reason why a reduction in average root diameter was observed in the compacted layer. Average root diameter was influenced by the root and shoot genotypes. Gibberellin (GA) may be a possible signal from the shoot that influenced average root diameter in this study. Gibberellin moves from shoot to root and vice versa and inhibitors of GA biosynthesis have been shown to increase root diameter (Tanimoto, 2005). Total root length was not an indicator of root rot resistance in Experiment _ 1. In Experiment 2, however, increased root length below the compacted layer was observed in the resistant genotypes and graft combinations, indicating that FR266 (in any root or shoot combination) was better able to penetrate the compacted layer than Montcalm. One of the mechanisms by which roots move through soils with an increased bulk density and a smaller pore size is by 59 increasing root diameter and displacing soil to push through an area of compaction (Goodman and Ennos, 1999). The different interaction of root and shoot signals under varying soil environmental conditions is a likely reason for the difficulty in developing common bean lines tolerant to Fusarium root rot. Conclusions Root genotype controlled root rot incidence in the absence of compaction, but with the addition of a compacted soil layer, the interaction of the root and shoot genotype dictated root rot incidence. In the absence of compaction, root traits were clearly influenced by root or shoot genotype. In the soil with a compacted layer, interaction of the root and shoot genotypes was important in determining root system architecture. 60 References: Abawi, GS, and MA. Pastor Corrales. 1990. Root rots of beans in Latin America and Afiica: Diagnosis, research methodologies, and management strategies CIAT, Cali, Colombia. Allmaras, R.R., J.M. Kraft, and DE. Miller 1988. Effects of Soil Compaction and Incorporated Crop Residue on Root Health. Ann. Rev. Phytopathol. 26:219- 243. Asady, G.H., A.J.M. Smucker, and M.W. Adams. 1985. Seedling Test for the Quantitative Measurement of Root Tolerances to Compacted Soil. Crop Science ' 25:802-806. Bennie, ATP. 1991. Growth and Mechanical Impedance. Pages 393-414 in: Plant Roots, the Hidden Half. Y.Eisel, A. Eisel, and U. Kaflcafi, eds. Marcel Dekker, Inc., New York. Bachman, MS. and Nickell, CD. 1999. Use of Reciprocal Grafting to Study Brown Stem Rot Resistance in Soybean. Phytopathology 89:59-63 Burke, D. W. 1968. Root Growth Obstructions and Fusarium Root Rot of Beans. Phytopathology 58: 1575- 1576. Burke, D.W., and A.W. Barker. 1966. Importance of Lateral Roots in Fusarium Root Rot of Beans. Phytopathology 56:292-294. Burke, D.W., and J .M. Kraft. 1974. Responses of Beans and Peas to Root Pathogens Accumulated during Monoculture of Each Crop Species. Phytopathology 64:546-549. Burke,.D.W., and DE. Miller. 1983. Control of Fusarium Root Rot with Resistant Beans and Criltural Management. Plant Disease 67: 1312-1317. Bushamuka, V.N. and Zobel R.W. 1998. Differential Genotypic and Root Type Penetration of Compacted Soil Layers. Crop Sci. 38:776-781. Chatterj ee, P. 1958. The Bean Root Rot Complex in Idaho. Phytopathology 48: 197-200. Copeland, L. O., and M. H. Erdmann. 1977. Montcalm and Mecosta, halo blight tolerant kidney bean varieties for Michigan. Ext. Bull. 957. Michigan State University, East Lansing Goodman, A.M., and AR. Ennos. 1999. The effects of soil bulk density on the 61 morphology and anchorage mechanics of the root systems of sunflower and maize. Annals of Botany 83:293-302. Harveson, R.M., Smith, J .A., and Stroup, W.W. 2005. Improving Root Health and Yield of Dry Beans in the Nebraska Panhandle with a New Technique for Reducing Soil Compaction. Plant Dis. 89: 279-284. Hoagland, D. R. & Amon, D. I. (1938) The Water-Culture Method for Growing Plants Without Soil (Univ. of California College of Agriculture Agricultural Experiment Station, Berkeley, CA). Huisman, CC. 1982. Interrelations of Root Growth Dynamics to Epidemiology of Root Invading Fungi. Ann. Rev. Phytopathol. 20:303-327. Izquierdo, J .A., and G.L. Hosfield. 1982. A Simplified Procedure for Making Cleft Grafts and the Evaluation of Grafting Effects on Common Bean. Hortscience 17:750-752. Johnson W.C., L.E. Jackson, 0. Ochoa, R. van Wijk, J. Peleman, D.A. St Clair, R.W. Michelmore 2000. Lettuce, a shallow-rooted crop, and Lactuca serriola, its wild progenitor, differ at QTL determining root architecture and deep soil water exploitation. Theor. App. Genet. 101: 1066-1073. Keenan, J .G., H.D. Moore, N. Oshima, and LE. Jenkins. 1974. Effect of Bean Root Rot on Dryland Pinto Bean Production in Southwestern Colorado. Plant Disease Reporter 58:890-892. Miller, DE, and D.W. Burke. 1985. Effects of Soil Physical Factors on Resistance in Beans to Fusarium Root-Rot. Plant Disease 69:324-327. Pantalone, V.R., Rebetzke, G.J., Burton, J .W., Carter, TE, and Israel, D.W. 1999. Soybean PI 416937 Root System Contributes to Biomass Accumulation in Reciprocal Grafts. Agron.J. 91: 840-844. Roman-Aviles, B., Snapp, SS, and Kelly, J .D. 2004. Assessing root traits associated with root rot resistance in common bean. Field Crops Res. 86:147-156. Saettler, A.W. 1982. Bean Disease and Their Control, In (L. S. a. F. Robertson, R.D., ed. Dry Bean Production: Principles and Practices. MSU Cooperative Extension, East Lansing. Schneider, K.A., and J .D. Kelly. 2000. A Greenhouse Screening Protocol for Fusarium Root Rot in Bean. HortScience 35: 1095-1098. Schneider, K.A., K.F. Grafton, and J .D. Kelly. 2001. QTL Analysis of Resistance to Fusarium Root Rot in Bean. Crop Sci. 41 :535-542. 62 Sheng, C. and J .E. Harper 1997. Shoot versus R00t Signal Involvement in Nodulation'and Vegetative Growth in Wild—Type and Hypemodulating Soybean Genotypes. Plant Physiol. 113:825-831. Silbemagel, M.J. 1987. Fusarium Root Rot-resistant Snap Bean Breeding Line FR266. HortScience 22:1337-1338. Silbemagel, M.J. and LJ. Mills 1990. Genetic and Cultural Control of Fusarium Root Rot in Bush Snap Beans. Plant Dis. 74: 61-66 Snapp, S., W. Kirk, B. Roman-Aviles, and J. Kelly. 2003. Root Traits Play a Role in Integrated Management of F usarium Root Rot in Snap Beans. HortScience In Press. ' Steadman, J .R., E.D. Kerr, and RF. Murnm. 1975. Root Rot of Bean in Nebraska: Primary Pathogen and Yield Loss Appraisal. Plant Disease Reporter 59:305-308. Tanimoto, E. 2005. Regulation of root grth by plant hormones - Roles for auxin and gibberellin. Critical Reviews in Plant Sciences 24:249-265. White, J .W. and Castillo, J .A. 1989. Relative Effect of Root and Shoot Genotypes on Yield of Common Bean under Drought Stress. Crop Sci. 29:360-362. Wallace, DH, and RE. Wilkinson. 1965. Breeding for Fusarium Root Rot Resistance in Beans. Phytopathology 55: 1227-1231. 63 Table 1. Selected plant and seed characteristics of the common bean cultivars FR266 and Montcalm. Cultivars FR266 Montcalm Reaction to resistant susceptible F usarium solani Gene pool Andean Andean Market class Snap Kidney Plant growth habit determinate determinate Seed color white red Seed size medium large 100 seed weiLht 30.7 g 61.3 g 64 Table 2. Mean response root rot score, shoot dry weight, and root dry weight of different graft combinations of common bean lines FR266 and Montcalm in the presence (+ F us) or absence (-Fus) of F usarium solani f.sp phaseoli inoculum. Data analyzed across 2 runs of experiment 1. Root rot Shoot dry Root dry score (1-7)j weight (mg) weight (mg) Graft combination Root Shoot -Fus +Fus -Fus +Fus -F‘us + Fus FR266 ungrafted 1.5 2.9 785 834 604 580 FR266 FR266 1.0 2.1 634 718 330 413 FR266 Montcalm 1.4 2.5 726 831 470 526 Montcalm ungrafted 1.6 4.1 770 868 449 530 Montcalm Montcalm 1.8 4.5 816 782 381 371 Montcahn FR266 1.6 4.3 705 710 459 408 LSD 0.051 0.84 172 132 ANOVA Source df p value Run (R) 1 0.0001 0.0006 <0.0001 Replication 3 0.8838 0.0649 0.8174 Graft (G) 5 <0.0001 0.0461 <0.0001 Fusarium (F) 1 <0.0001 0.1068 0.3075 G*F 5 0.0018 0.7542 0.3324 R*G 5 0.5961 0.4347 0.1013 R*G*F 5 0.3825 0.1603 0.0264 T Root rot score is on a scale of 1-7, where 1 is no root rot and 7 is completely rotted ( Schneider and Kelly, 2000). I LSD value to compare any values within root rot score, shoot dry weight, and root dry weight variables. 65 Table 3: Mean response for total root length and average root diameter of ungrafted Montcalm and F R266 in the presence (+ Fus) or absence (-Fus) of F usarr’um solani f.sp phaseoli inoculum. Data analyzed across 2 runs of experiment 1. ' Genotype Total root Average root length (cm) diameter (mm) -Fus +Fus -Fus +Fus FR266 3475 at 3402 a 0.461 a 0.467 a Montcalm 3273 a 2990 a 0.417 b 0.399b T= Values that do not share a letter are significantly different as determined by LSD (0.05). Letters are compared down a column. 66 Table 4: Mean response and analysis of variance for root traits of different graft combinations of common bean lines FR266 and Montcalm in the presence of F usarium solani f.sp phaseoli inoculum. Data analyzed across 2 runs of experiment 1 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response. Graft Combination Root Shoot Root Rot Root dry Root Avg root Score wt. (mg) length diam. (1-7)L (cm) me) FR266 FR266 2.1 at 453 ab 3070 ab 0.45 a FR266 Montcalm 2.5 a 527 b 3209 b 0.44 a Montcalm FR266 4.3 b 351 a 2920 ab 0.46 a Montcalm Montcalm 4.5 b 371 a 2689 c 0.40 b ANOVA Source df p value Run (R) l <.0001 <.0001 0.04 <.0001 Replication 3 0.40 0.95 0.18 0.17 Root 1 <.0001 0.01 0.17 0.23 Shoot 1 0.26 0.23 0.86 0.04 R*Root l 0.82 0.18 0.67 0.29 R*Shoot l 0.26 0.1 1 0.47 0.98 Root*Shoot l 0.82 0.34 0.42 0.13 R*Root* l 0.26 0.75 0.81 0.61 Shoot 1' Root rot score is on a scale of 1-7, where 1 is no root rot and 7 is completely rotted (Schneider and Kelly, 2000). I Values that do not share a letter are significantly different as determined by LSD (0.05). Letters are compared down a column. 67 Table 5: Mean response for root traits of different graft combinations of common bean lines FR266 and Montcalm in the presence of F usarium solani f.sp phaseoli inoculum. Data analyzed across 2 runs of experiment 1 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response, means are‘averaged by root genotype and shoot genotype. Root 1' Root Rot Root dry f Root Avg. root Score (1-7) * wt. (mg) length diam. (cm) (mm) FR266 2.31 a** 492 a 3144 a 0.446 a Montcalm 4.38 b 362 'b - 2805 a 0.426 a Shoot it FR266 3.19 a 402 a 2991 a 0.455 a Montcalm 3.50 a 449 a 2949 a 0.418 b T Root heading indicates that means of each variable are averaged over root genotype. i Shoot heading indicates that means of each variable are averaged over shoot genotype. . * Root rot score is on a scale of 1-7, where 1 is no root rot and 7 is completely rotted (Schneider and Kelly, 2000). ** Values that do not share a letter are significantly different as determined by LSD (0.05). Letters are compared down a column; comparisons under ‘Root’ are separate from those under ‘Shoot’. 68 Table 6. Mean response of root rot score, shoot dry weight, and root dry weight of different graft combinations of common bean lines FR266 and Montcahn grown in containers with a layer of compacted soil in the presence (+ Fus) or absence (- Fus) of F usarium solani f.sp phaseoli inoculum, Data analyzed across 1 run (Nov. 2005) of experiment 2. Root rot Shoot dry Root dry score (1 -7) 1‘ weight (mg) weight (113) Graft combination Root Shoot -Fus +Fus -Fus +Fus -Fus + Fus FR266 ungrafted 1.0 2.4 1424 1436 350 398 FR266 FR266 1.0 2.3 717 647 129 153 FR266 Montcahn 1.0 2.5 . 710 651 140 144 Montcalm ungrafted 1.0 4.0 1158 932 198 256 Montcalm Montcalm . 1.0 4.5 61 1 321 129 1 15 Montcahn F R266 1.0 2.5 700 692 1 10 129 LSD 0.05:1: 0.74 252 94 ANOVA Source (11‘ ' p value Replication 3 0.5330 0.2752 0.8888 Graft (G) 5 <0.0001 <0.0001 <0.0001 Fusarium (F) 1 <0.0001 0.0230 0.2062 G*F 5 <0.0001 0.3361 0.8516 T Root rot score is _on a scale of 1-7, where 1 is no root rot and 7 is completely rotted (Schneider and Kelly, 2000). 1 LSD value to compare any values within root rot score, shoot dry weight, and root dry weight variables. 69 Table 7: Mean response for total root length and average root diameter of ungrafted Montcalm and F R266 grown in containers with a layer of compacted soil in the presence (+ Fus) or absence (-F us) of F usarium solani f.sp phaseoli inoculum, Data analyzed across 2 runs of experiment 2. Soil layer Genotype Total root Average root lengh (cm) diameter (mm) Top -Fus +Fus -Fus +Fus FR266 1577 at 677 a 0.454 a 0.43 a Montcalm 1412 a 525 a 0.419 b 0.42 a Middle FR266 181 a 215 a 0.380 a 0.37 a Montcalm 248 a 180 a 0.344 a 0.37 a Bottom FR266 2483 a 4322 a 0.406 a 0.45 a Montcalm 2141 a 3082 b 0.376 a 0.40 b TValues that do not share a letter are significantly different as determined by LSD (0.05). Letters are compared down a column. 70 Table 8. Mean response and analysis of variance for root traits of different graft combinations of common bean lines FR266 and Montcalm grown in containers with a layer of compacted soil in the presence of F usarium solani f.sp phaseoli inocultun. Data analyzed across 1 run (Nov. 2005) of experiment 2 and the category graft combination is separated into root and shoot components for analysis to determine which components are affecting a response. Graft combination Soil IayerT Root Shoot Root Root Root Avg Rot dry wt. length root Score (mg) (cm) diam. (1-7)l (mm) Top FR266 FR266 2.25 a * 89 a 461 a 0.44 a FR266 Montcalm 2.50 a 84 a 326 a 0.46 a Montcahn FR266 2.50 a 50 b 449 a 0.39 b Montcahn Montcalm 4.50 b 78 a 436 a 0.46 3 AN OVA Source df p value Root (R) 1 0.002 0.019 0.481 0.0019 Shoot (S) 1 0.002 0.213 0.300 0.004 R*S 1 0.0113 0.063 0.387 0.053 Middle FR266 FR266 NA 7 ab 84 ab 0.37 a compacted FR266 Montcalm -- 9 a 94 ab 0.37 a Montcalm FR266 -- 7 ab 119 a 0.36 a Montcalm Montcalm -- 5 b 40 b 0.37 3 ANOVA Source (if p value Root (R) 1 NA 0.07 0.698 0.75 Shoot (S) 1 -- 0.917 0.177 0.95 R*S 1 -- 0.099 0.087 0.77 Bottom FR266 FR266 NA 57 a 1204ab 0.36 a FR266 Montcalm -- 95 b 1667 a 0.36 a Montcahn FR266 -- 76 ab 1568 a 0.34 a Montcalm Montcalm -- 28 c 667 b 0.37 a ANOVA Source df p value Root (R) 1 NA 0.018 0.252 0.49 Shoot (S) 1 -- 0.534 0.421 0.33 R*S 1 -- 0.006 0.026 0.07 T Soil layer is defined as top, middle, and bottom. Root traits were measured separately for each layer. I Root rot score is on a scale of 1-7, where l is no root rot and 7 is completely rotted (Schneider and Kelly, 2000). *Values that do not share a letter are significantly different as determined by LSD (0.05). Letters are compared down a column. 71 Figure 1. Photos of containers used in Experiment 1 (a) and Experiment 2 (b). The bulk densities of soil are listed in g/cm3 for each of the 3 layers of containers used in experiment 2. 72 Chapter 3: The Relationship among Root Architecture Traits, Plant Growth Habit and Tolerance to Low Soil Phosphorus Levels in an Andean Bean Population Abstract: Tolerance to low P soils is a desirable characteristic for common bean line grown in acid-weathered soils of the tropics. Variability for tolerance to low P soils exists in the Andean gene pool. The objective of this research was to identify mechanisms of tolerance to low P soils in a By recombinant inbred line (RH) population developed fiom two common bean lines in the Andean gene pool, AND696, an improved line, susceptible to low P soils, with a determinate grth habit and G19833, a landrace tolerant to low P soils, and an indeterminate growth habit. The 77 RILs were grown in low and high soil phosphorus levels in a field in Darien, Colombia. P uptake and rootarchitecture traits were measured in the population, along with P uptake efficiency and seed P content. Genetic differences were identified for root architecture traits, including root length density, specific root length, and average root diameter in the top 16 cm of soil. Specific root length was weakly correlated with P uptake (r=0.16) in both the high and low P treatments. P uptake and P use efficiency were correlated with yield. Plant growth habit affected seed yield, and root length density, but was not related to P uptake. 73 Introduction: Common bean (Phaseolus vulgaris L.) is an important grain legume worldwide, produced on 20 million hectares of land each year. The highest production and consumption of common bean occurs in Latin America (5.6 million metric tons) and Afi'ica (2.8 million metric tons) (FAOSTAT, 2006). One of the major constraints to bean production in Afi'ica and Latin America is low available soil phosphorus (Wortmann et al., 1998). An estimated 43% of the lands in the tropics are acid weathered soils, specifically Andosols, Ultisoils, and Oxisols, that adsorb or fix P so that 70 to 90% of P fertilizer applied. reacts with iron or aluminum to form compounds of low solubility (Sanchez and Salinas, 1981). Common bean is susceptible to low P soils in part due to the high P requirement associated with energy-intensive nodulation and N2 fixation (Christiansen and Graham, 2002). Tolerance to low P soils or P efficiency is defined as the ability to produce plant growth and yield in relation to the amount of available phosphorus (Lynch and Beebe, 1995) and can occur by two distinct routes: acquisition efficiency and utilization efficiency. Acquisition efficiency reflects the plants ability to extract P fiom the environment. It has been shown to be related root system traits that increase the root surface area and allow capture of more P from the soil (Gahoonia and Nielsen, 2003). Utilization efficiency is the superior ' ability of a plant to convert acquired P into plant biomass and yield and is related to reduced tissue P requirement (Lynch and Beebe, 1995). 74 Genetic variability for tolerance to low P soils has been identified in common bean (Beebe et al., 1997; Singh et al., 2003). Within the primary gene pool of common bean there are multiple sites of domestication divided into two major groups, Mesoamerican (northern Mexico and Central America) and Andean (South America) (Singh, 2001). The most studied bean lines tolerant to low P soils have been landraces from the Andean gene pool (Lynch and Beebe, 1995; Yan et al., 1996; Bonser et al., 1996; Yan et al., 2004; Beebe et al., 2006). The Andean landraces identified as tolerant to low P soils have been shown to be efficient in P acquisition from the soil (Liao and Yan, 2001; Liao et al., 2001). Controlled environment physiology studies have identified specific plant root traits related to P uptake in these P efficient lines, including increased lateral and adventitious root number and more shallow basal roots. (Lynch and Brown, 2001; Miller et al., 2003). QTL studies employing inter-gene pool crosses have identified regions of the bean genome important for both root grth and P uptake (Liao et a1, 2004; Beebe et al., 2006; Ochoa et al., 2006). Within a - recombinant inbred line population developed fi'om a P efficient, Andean bean, G19833 and a P inefficient, Mesoamerican bean DOR364, QTL were identified for total root length and specific root length on two different linkage groups that co-localized with QTL for phosphorus uptake. The uptake QTL and the root QTL each explained about 15% of phenotypic variation for these traits (Beebe et al., 2006). Low P tolerant, Andean landraces perform well in low P soils, but typically do not yield well in P sufficient soils. They have an indeterminate 75 growth habit, long maturation time, and are day length sensitive (Beebe et al., 1997). There is question as to whether the features of these landraces that enhance yields in low P environments are the same features that limit their yield potential in other environments. Previous observational studies with common bean support a coordinated grth habit of the root and shoot such that indeterminate plant types with highly branched shoot architecture and little apical dominance also exhibited highly branched root architecture (Lynch and van Beem, 1993). The parental material used in this study had contrasting plant growth habit, and the population was segregating for grth habit, which allowed the opportunity to look at the relationship among grth habit, root architecture, and P acquisition efficiency. The first objective of this study was to identify the broad mechanisms (P acquisition and P use efficiency) of tolerance to low P soils in a bean recombinant inbred line (RIL) population developed from an Andean intra-gene pool cross between a P efficient, unadapted landrace, G19833 and a P inefficient line, AND696. Specific root architecture traits were also examined for their relationship to P acquisition efficiency. The RH. population was grown in both P deficient and P sufficient soils and plants were evaluated in each environment to clearly identify traits generally important for plant yield vs. those Specifically important for adaptationto low P soils. The second objective of this study was to determine if there were differences in shoot growth habit associated with differences in root growth and if they were related to P uptake in common bean. 76 Materials and Methods: Seventy seven F 5:7 recombinant inbred lines were developed from a cross between two common bean lines from the Andean gene pool. The initial cross G19833 x AND696 was advanced to the F 5 generation by single seed descent and seed of the RILS was increased for field studies. The parent G19833 is a Peruvian landrace with an indeterminate (Type III) growth habit. G19833 has large yellow and red mottled seed with an average 100 seed weight of 41 g. G19833 is relatively unadapted (i.e. low yield potential) but has been identified as tolerant to low P soils (Y an et al., 199Sab). The parent AND696 is a CIAT improved line from the race Nueva Granada. It has a determinate grth habit (Type I) and has large red and white mottled seed with an average 100 seed weight of 51 g. AND696 has been identified as susceptible to low phosphorus soils (CIAT, 2000). The RIL population developed from G19833 and AND696 segregated for plant growth habit, 56 of the lines have a determinate growth habit and 21 lines have an indeterminate growth habit. Growth habit in common bean is controlled by a single gene, fin (Norton, 1915). The number of determinate and indeterminate lines deviated significantly from a 1:1 ratio that is expected for a single gene trait. This deviation is likely due to the selection of RILs with i uniform days to maturity, which biased the selection to favor detemrinate lines. In 2000 and 2005, the 77 RILs, G19833, AND696, and two check varieties, G4017 (Carioca) and Gl6140 were planted in Darien, Colombia in a 9x9 lattice design with three replications at two soil P levels, low P (45 kg/ha triple super phosphate) and high P (300 kg/ha triple superphosphate). Phenotypic 77 data was collected on 75 of the 77 RILS. The soil of this site in an Andisol with a native soil P of 2 mg/kg based on Bray II extraction method. Seed was hand- planted at 10 cm spacing in 4 row plots. Each row was 3 m in length. The middle tworows were planted to the genotypes of interest and the outer two rows were border rows of Mesoamerican tan seeded cultivar BAT 477. The border rows were included to improve uniformity in soil P availability to the different genotypes and also aided in identification of genetic material. During the 2005 growing season, various plant measurements were taken to elucidate underlying factors involved in differences between the parental genotypes in tolerance to low soil P. Adventitious root number was counted on twoplants for each P treatment and in each replication at 21 days after planting. This was done by excavating entire root systems from the ground. At mid-pod fill (which ranged from 54 to 67 days after planting) shoots from 50 cm row length (5-10 plants) were harvested and oven dried at 70°C until free of moisture (about 5 days). Dry weights were recorded and shoots were ground to a fine powder in a Wiley mill with a 60-mesh screen. Plant tissue samples were Kj eldahl acid digested and subsequently measured for concentration of P according to the method of Murphy and Riley (1962) using colorimetric spectrometry. At mid pod fill, a single root sample was taken per plot using a soil anger with a diameter of 7.1 cm and to a depth of 16 cm. Samples were taken directly adjacent to the plant stem. Roots were separated fiom soil and washed with water. Soil was dried and weighed to determine volume. Cleaned roots were placed in a transparent plastic tray (22 x 28 x 6.5 cm) and covered with water. The tray 78 containing roots was then scanned with a flat bed scanner with a top and bottom light source at 300 dpi to produce a 2 dimensional scanned image. The images were analyzed with WinRHIZO software (Regent Instruments Inc., Quebec, Canada). Root length, root surface area and average root diameter of the samples was determined. Following image capture, roots were dried at 65°C for five days and dry weights were determined. Seeds reached maturity at 76 to 98 days after planting. At maturity, seeds were hand—harvested. Seeds were dried and seed yield was determined at 18% moisture. Seed weight (of 100 seed) was measured at 18% moisture for the 2000 and 2005 plantings. A sub sample of seed fi'om the 2000 and 2005 harvest was analyzed for total phosphorus concentration. Five grams of seed of each treatment for each of the three replications were cleaned with distilled water and dried. Samples were then freeze dried to remove all moisture. Freeze-dried seed samples were placed into aluminum tubes containing three silver balls. Mechanical agitation was used to crush seeds into a fine powder. Seed samples were Kjeldahl acid digested and subsequently measured for concentration of P according to the method of Murphy and Riley (1962) using colorimetric spectrometry. Statistical analysis was conducted with SAS for Windows V8 (SAS Institute, Cary, NC). The command PROC GLM was used to determine treatment, genotype and interaction effects. Data was analyzed as a split plot with soil P level as the whole plot and plant genotype as the split plot. The coMand PROC CORR was used to determine Pearson correlation coefficients among variables. 79 Effects of plant growth habit on measured variables were determined by means comparisons of determinate vs. indeterminate growth habit types and differences were established using Tukey’s test for significance. Bartlett test for variance homogeneity was conducted for growth habit because there were unequal number of determinate (56) and indeterminate (21) RILS. Results: P uptake: Significant differences in P uptake were observed between environments such that mean uptake of the RILs was more than 3 times higher in the high P than the low P environment (Table 1 and Table 2). Genotypic variability was observed for P uptake among the RIL population (Table 1). P uptake by the inefficient parent, AND 696 was below the RIL mean in the high P environment, and similar to the RIL mean in the low P environment (Table 2). Significant genotype x environment (G x E) interactions were observed for this trait (Table 1). P use efficiency: In addition to P uptake efficiency, P use efficiency also exists as a tolerance mechanism to low P soil. In this study P use efficiency was defined as the amount of seed yield per unit of P uptake by the plant. There was variability for P use efficiency between environments and among genotypes. G x E was not significant (Table 1). Efficient use of P was greater in the low P environment, but genotypic differences in the population were only observed in the high P environment (Table 2). 80 Seed P content: Another measure of P efficiency is the content of P in the seed. There were significant environmental and genetic differences for seed P content observed, and a G x E effect was present in 2000, but not 2005 (Table 1 and Table 2). The seed P content of AND696 was higher than the mean RIL content in both the high and low P soil treatments (Table 2). Yield: Significant differences in seed yield were observed between the high and low soil P environments in the RIL population (Table 1). The mean yield in low P was 597 kg/ha and 1345 kg/ha in high P in 2005. Similar yield differences as those seen in 2005 were also recorded between the environments in 2000 (Table 2). The differences in yield in each environment demonstrate the effects of limiting P on plant growth in this population. Yield of the check variety, Carioca was higher than that of the RIL population across environments and years (Table 2). In both years, there was significant genotypic variability for yield and in 2005 there was a significant G x E interaction (Table 1). Correlations: P uptake was significantly correlated with seed yield in .2005 in each P treatment, and the correlation coefficient was higher in the low P (1:053) than the high P (r=0.48) treatment (Table 3). This is consistent with P uptake having an important role in yield detennination in a P limiting environment. P use efficiency was more strongly correlated with seed yield in the high P than the low P treatment (Table 3). There was a negative correlation between P uptake and P use efficiency. This was expected because by definition, P use efficiency is based on getting the highest yields with the least amount of P 81 uptake. Seed P content was positively correlated with seed yield in 2005, but not in 2000 (Table 9). i Root traits and P uptake: A selection of root traits was measured in the 2005 growing season to determine their importance in P uptake. Adventitious root number was counted at 3 weeks after planting. These roots arise fiom shoot tissue, and have a lower carbon construction cost than roots arising from root tissue, as they have low density (Lynch and Ho, 2005). They are generally shallow in orientation and deploying in the topsoil therefore closer to P enriched sources of the soil, such as organic matter. Environmental differences were observed for adventitious root number (Table 5). No genotypic differences were detected for adventitious root number in this population (Tables 4 and 5). Mean squares fiom analysis of variance identified significant genotypic differences in root length density, specific root length, root surface area, and average root diameter (Table 4). Further inspection of these traits indicated that in the low P soil treatment, only root length density and root surface area were different among the RILS in this study (Table 5). These traits were not however correlated with P uptake. Only specific root length was weakly correlated with P uptake in the low P soil treatment (Table 6). i In the high P so'il environment, root length density, root surface area, and average root diameter showed genotypic variation within the RILs (Table 5). In this environment there were also genotypic differences in P uptake (Table 2). There was a weak correlation between specific root length and P uptake and a negative correlation between average root diameter and P uptake (Table 6). 82 _ The weak correlations presented in this study between root architecture traits and P uptake are not surprising because roots are extremely plastic, especially in field environments. However, despite the plastic nature of root growth it was possible to detect genetic" differences in these traits. The observation that there was genotypic variability for root architecture traits, but these traits were not correlated with P uptake shows that these traits should not be considered adaptive for tolerance to low P soils in this population. Differences by Growth Habit In this study, the RIL population segregated for indeterminate and determinate plant grth habit, this trait is controlled by a single gene, the fin gene (Koinange et al., 1996; Norton et al., 1915). Plant grth habit has been shown to be linked to maturity in common bean (Coyne and Schuster, 1.97 4; White et al., 1991) as observed here. Detenninate RILs in the population reached maturity on average in 80 to 82 days whereas indeterminate RILs required 88 to 89 days to reach maturity depending of soil P level (Table 7). There were differences in seed yield in the determinate and indeterminate RILS in both 2000 and 2005. The trend however was not the same across years. In 2000, the indeterminate RILS had greater seed yield than the determinate RILS under low P and no differences in yield under the high P treatment (Table 7). Detenninate plants have separate vegetative and reproductive stages of development, whereas indeterminate plants continue to grow vegetatively during flowering and seed development (Huyghe, 1998). In 2000, the indeterminate plants yielded higher on average than the determinate plants, perhaps because of 83 the ability to grow longer and acquire nutrients whereas the determinate plants had switched fully to seed production. ' In 2005 there were no differences is seed yield in low P between the determinate and indeterminate plants. Yield was however greater in the determinate plants in the high P treatment (Table 7). Comparison of the 2000 and 2005 yield data suggest the presence of an additional stress present in 2005 that affected the indeterminate RILs to a greater extent than the determinate RILS. Such a stress may have become a factor later in the life cycle of the plants because the indeterminate lines took 6-9 days longer to reach maturity. In 2005, no significant differences in shoot dry weight or P uptake were observed for determinate as compared to indeterminate RILs (Table 7). There were differences observed in two root traits, but these differences were only present in the high P soil treatment. The indeterminate plants had greater root length density and root surface area than determinate RILS in the population (Figures 1 and 2). Specific root length was weakly correlated with P uptake in the determinate RILS in the low P treatment and with the indeterminate RILS in the high P treatment (Table 8). The determinate RILs had a higher P use efficiency in high P than the indeterminate RILS. In 2005 the seed P content was higher in the indeterminate RILS in both the high and low P treatments. In 2000 differences in seed P content were only present in the high P treatment (Table 7). Correfations for seed yield and seed P content were significant and positive in 2005 but not in 2000 (Table 9). Since there were unequal number of determinate and indeterminate RILS in the 84 population, Bartlett’s test for variance homogeneity was conducted to identify bias that may have existed in trait analysis by growth habit. Most traits exhibited variance homogeneity although there were exceptions, including seed P content (high P), yield (low P, 2000) and root length density, specific root length, and root surface area (high P) (Appendix, Tables A3-l; A3-2). Discussion: Seed yields were greatly reduced in the low P treatment, and genetic variability was observed for response to low soil phosphorus. The population also exhibited genetic differences in P uptake and P use efficiency, although only P uptake efficiency was important in the low P soil treatment. Tolerance to low soil P exhibited by the efficient parent G19833 has been shown to be based on its efficient uptake of P from the soil, which can be attributed to its extensive root system (Liao et al., 2001; Nielson et al., 2001). Root architecture traits that increase the area of soil exploration may in turn increase the capture of immobile P ions by a plant root system (Gahoonia and Nielsen, 2003). Each of the root traits measured in this study are indicators of explorative capacity and we hoped to be able to use the RILs to identify those root traits important for P uptake. None of genetic variability observed for the root architecture traits studied here factored into improved P uptake efficiency. Previous studies have, however, identified adventitious root number as important for P uptake (Miller et al., 2003) and studies of RILs developed by a cross between an Andean and Mesoamerican line have shown increased numbers of adventitious roots to be important for P uptake (Ochoa et al., 2006). Root 85 length density and root surface area are both measures of root spacing in a volume of soil. Specific root length and average root diameter indicate root thickness. Thinner roots have an advantage in P uptake over thicker roots because they require less construction cost in the form of carbon, to explore the same area of soil (Lynch and Brown, 2001). Most plant available P is found in the topsoil, therefore in this study, measurements on root traits were conducted on roots in the top 16 cm of soil. Perhaps within the Andean gene pool there is less variability for root traits than in the Mesoamerican gene pool. 9 Genetic variability was uncovered for P use efficiency in the high P treatment. This trait was associated with the determinate plants (Table 7), and may be indicative of the different grth strategies of indeterminate and determinate plants. In rice (Oryza sativa L.), for example, P use efficiency was negatively correlated with plant dry weight and found to be an adaptation to low soil P found in plants inefficient in P uptake (Wissuwa et al., 1998). Selection of genotypes with greater seed P content may improve plant growth in low P soils. Nutrients stored in the seed are important for germination and early plant growth (Liptay and Arevalo, 2000). Increased seed P content partially compensated for the negative effects of low P scils on early seedling grth in wheat (Triticum aestivum L.), by increases in shoot dry weight, leaf size, and root length (De Marco, 1990). A positive correlation between seed P and yield was observed in 2005. The correlation between these traits was stronger for the indeterminate than the determinate lines and therefore may be a more 86 important factor when specifically selecting indeterminate lines adapted to low soil P. Another important factor that may be related to plant growth habit is adaptation, because the parents exhibited differences for this trait. The unadapted nature of G19833 can be seen in Figures 3 and 4, where seed yield in high P is regressed by seed yield in low P in 2000 and 2005. G19833 did not perform well in the high P environment (Figure 4). AND696 on the other hand is an improved line and demonstrated an average yield potential in the high P environment, although it performed poorly in the low P environment (Figure 3). Based on the yield regressions, it would appear that selecting a determinate plant type, would be the best way to maximize yield potential across environments. In 2005, the top yielding lines across environments possessed a determinate growth habit (Figure 3), whereas in 2000, the indeterminate RILS on average performed better than the determinate lines, but numerous determinate lines performed well in both environments (Figure 4). In a previous study of Andean beans conducted in three environments in Colombia, determinate lines were found to have greater yield stability (White et al., 1992). Detenninate lines are not always more stable and other studies in temperate environments have shown indetenninate lines to be more stable (Kelly et al., 1987). Conclusions Genotypic differences were observed for yield in 2000 and 2005. There were also genotypic differences for P uptake in‘the low P treatment and P uptake 87 and P use efficiency in the high P treatment. Both P uptake and P use efficiency were correlated with yield. Genetic variability for root length density and root surface area was observed among the RILS. These differences did not appear to improve adaptation to low P soils, however, because they were not correlated with P uptake. Growth habit affected seed yield differently between years, and was not related to P uptake in 2005. Root length density and root surface area were greater in indeterminate plants in the high P treatment only. Determinate lines were more P use efficient in high P and indeterminate lines had greater seed P content in high and low P in 2005 and in low P in 2000. The differences in trait response in 2000 and 2005 suggest that while the indeterminate lines were (on average) better adapted to the low P environment than the determinate lines, the determinate lines were better able to perform across the two environments. 88 References: Beebe, S., J. Lynch, N. Galwey, J. Tohme, and I. Ochoa. 1997. A geographical approach to identify phosphorus-efficient genotypes among landraces and wild ancestors of common bean. Euphytica 95:325-336. sBeebe, S.E., M. Rojas-Pierce, X.L. Yan, M.W. Blair, F. Pedraza, F. Munoz, J. Tohme, and J .P. Lynch. 2006. Quantitative trait loci for root architecture traits correlated with phosphorus acquisition in common bean. Crop Science 46:413- 423. Bonser, A.M., J. Lynch, and S. Snapp. 1996. Effect of phosphorus deficiency on growth angle of basal roots in Phaseolus vulgaris. New Phytologist 132:281-288. Christiansen, 1., and PH. Graham. 2002. Variation in di-nitrogen fixation among Andean bean (Phaseolus vulgaris L.) genotypes grown at low and high levels of phosphorus supply. Field Crops Research 732133-142. CIAT. 2000. Inheritance of low phosphorus tolerance in the Andean population AND696 x G19833. Bean Improvement for the Tropics Unit Annual Report. CIAT, Cali, Colombia. Coyne, DP, and ML. Schuster. 1974. Inheritance and Linkage Relations of Reaction to Xanthomonas-Phaseoli Smith,Ef Dowson (Common Blight), Stage of Plant Development and Plant Habit in Phaseolus-Vulgaris L. Euphytica 23: 195- 204. Demarco, D.G. 1990. Effect of Seed Weight, and Seed Phosphorus and Nitrogen Concentrations on the Early Growth of Wheat Seedlings. Australian Journal of Experimental Agriculture 30:545-549. FAO. 2006. FAOSTAT [Online] http://faostat.fao.org (posted November 14, 2006; verified November 14). Gahoonia, TS, and NE. Nielsen. 2003. Phosphorus (P) uptake and growth of a root hairless barley mutant (bald root barley, brb) and wild type in low- and high- P soils. Plant Cell and Environment 26:1759-1766. Huyghe, C. 1998. Genetics and genetic modifications of plant architecture in grain legumes: a’review. Agronomic 18:383—411. Kelly, J .D., M.W. Adams, and G.V. Vamer. 1987. Yield Stability of Detenninate and Indeterminate Dry Bean Cultivars. Theoretical and Applied Genetics 74:516- 521. 89 Koinange, E.M.K., S.P. Singh, and P. Gepts. 1996. Genetic control of the domestication syndrome in common bean. Crop Science 36:1037-1045. Liao, H., and X.L. Yan. 2001. Genotypic variation in root morphological characteristics of common bean in relation to phosphorus efficiency. Acta Botanica Sinica 43:1161-1166. Liao, H., G. Rubio, X.L. Yan, A.Q. Cao, K.M. Brown, and J .P. Lynch. 2001. Effect of phosphorus availability on basal root shallowness in common bean. Plant and Soil 232:69-79. Liao, H., X.L. Yan, G. Rubio, S.E. Beebe, M.W. Blair, and J .P. Lynch. 2004. Genetic mapping of basal root gravitropism and phosphorus acquisition efficiency in common bean. Functional Plant Biology 31:959-970. Liptay, A., and AB. Arevalo. 2000. Plant mineral accumulation, use and transport during the life cycle of plants: A review. Canadian Journal of Plant Science 80:29-38. Lynch, J ., and J .J . Vanbeem. 1993. Grth and Architecture of Seedling Roots of Common Bean Genotypes. Crop Science 33:1253-1257. Lynch, J. P. and S. E. Beebe. 1995. Adaptation of Beans (Phaseolus-Vulgaris L) to Low Phosphorus Availability. Hortscience 30: 1165- 1171. Lynch, J .P., and KM. Brown. 2001. Topsoil foraging - an architectural adaptation of plants to low phosphorus availability. Plant and Soil 237:225-237. Miller, CR, 1. Ochoa, K.L. Nielsen, D. Beck, and J .P. Lynch. 2003. Genetic variation for adventitious rooting in response to low phosphorus availability: potential utility for phosphorus acquisition from stratified soils. Functional Plant Biology 30:973-985. Murphy, J ., and JP. Riley. 1962. A modified single solution for determination of phosphate in natural waters. Anal Chem Acta 27:31-36. Nielsen, K.L., A. Eshel, and J .P. Lynch. 2001. The effect of phosphorus availability on the carbon economy of contrasting common bean (Phaseolus vulgaris L.) genotypes. Journal of Experimental Botany 52:329-339. Norton, J .B. 1915. Inheritance of habit in the common bean. Am. Nat. 49:547-561. Ochoa, I..,E M. W. Blair, and J. P. Lynch. 2006. QTL analysis of adventitious root formation In common bean under contrasting phosphorus availability. Crop Science 46: 1609- 1621. 90 Park, 80., DP. Coyne, J .M. Bokosi, and J .R. Steadman. 1999. Molecular markers linked to genes for specific rust resistance and indeterminate growth habit in common bean. Euphytica 105:133-141. Sanchez, PA, and J .G. Salinas. 1981. Low-input technology for managing Oxisols and Utisols in tropical America. Advances in Agronomy 342280-406. Singh, SP. 2001. Broadening the genetic base of common bean cultivars: A review. Crop Science 41 :1659-1675. Singh, S.P., H. Teran, C.G. Munoz, J .M. Osomo, J .C. Takegami, and M.D.T. Thung. 2003. Low soil fertility tolerance in landraces and improved common bean genotypes. Crop Science 43:110-119. White, J .W., J. Komegay, J. Castillo, C.H. Molano, C. Cajiao, and G. Tejada. 1992. Effect of Growth Habit on Yield of Large-Seeded Bush Cultivars of Common Bean. Field Crops Research 29: 15 1-161. Wissuwa, M., M. Yano, and N. Ae. 1998. Mapping of QTLs for phosphorus- deficiency tolerance in rice (Oryza sativa L.). Theoretical and Applied Genetics 97:777-783. Wortmann, C.S., R.A. Kirkby, C.A. Eledu, and DJ. Allen. 1998. Atlas of common bean (Phaseolus vulgaris L.) production in Africa CIAT, Cali, Colombia. Yan, X.L., J .P. Lynch, and SE. Beebe. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .1. Vegetative Response. Crop Science 35:1086-1093. Yan, X.L., S.E. Beebe, and J .P. Lynch. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .2. Yield Response. Crop Science 35 : 1 094-1099. Yan, X.L., J .P. Lynch, and SE. Beebe. 1996. Utilization of phosphorus substrates by contrasting common bean genotypes. Crop Science 36:936-941. Yan, X.L., H. Liao, S.E. Beebe, M.W. Blair, and J .P. Lynch. 2004. QTL mapping of root hair and acid exudation traits and their relationship to phosphorus uptake in common bean. Plant and Soil 265: 1 7-29. 91 Table 1. Analysis of variance of traits related to phosphorus uptake and use, and seed yield in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien, Colombia in 2000 and 2005 under two treatments: high and low soil phosphorus. Mean squares Source df P P use Days to Seed P Seed yield uptake eff. Maturity content 2005 Genotype (G) 74 263‘" 0.04" 113'" 2092'" 161674'" Env. (E) 1 120770‘" 526"" 217‘" 324541‘" 62929473” G x E 74 220‘" 0.02 "S 5.4 "S 420 "S 103183” 2000 Genotype (G) 74 —————————— 55'" 1672'" 465740‘" Env. (E) 1 ---------- 195*" 205516‘" 88298547” G x E 74 ---------- 3.2 “s 673‘” 211912 “S “8 indicates not significant * Indicates significance at P < 0.05. ”Indicates significance at P <0.01. *** Indicates significance at P<0.001 92 Table 2. Mean plant growth traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and G19833 and the means and ranges of 75 recombinant inbred lines (RILS) developed from the parents. Means are also included for the check variety, Carioca. The experiment was planted in 2000 and 2005 in Darien, Colombia. Mean values are of 3 replications. P value indicates level of significant genotypic differences among the RILS for each trait. Parents Recombinant inbred lines Check Traits P AND G mean range p Carioca level 696 19833 value (G4017) P uptake HP 36.5 92 46.7 25-92 0.0010 70.9 (mg ~plant") LP 14.5 ----- 13.7 6.6-26.1 0.0112 8.9 P use HP 0.33 ----- 0.29 0.11-0.50 0.0016 0.29 efficiency LP 0.27 ----- 0.51 0.25-0.78 0.2787 0.79 (g seed mg P") Days to HP 77.3 ----- 82.9 77-98 <.0001 86.3 maturity LP 77 .3 ----- 84.2 77—94 <.0001 84.7 2005 Days to HP ----- 93 83.4 77-90 <.0001 86.7 maturity LP ----- 93 84.7 80-93 <.0001 83.3 2000 Seed P HP 198 165 193 146-261 <.0001 l 1 1 content LP 147 ----- 139 97-181 <.0001 87 (mg 100 seed") 2005 Seed P HP ----- 161 160 107-256 <.0001 93 content LP ---------- 1 16 76-151 <.0001 68 (mg 100 seed") 2000 Seed yield HP 1347 ----- 1345 582-1962 <.0001 2096 (kg-ha“) 2005 LP 465 ----- 597 323-944 0.0542 830 Seed yield HP ----- 1080 1516 681-2871 <.0001 3038 (kg ha") 2000 LP ----- 649 642 268-1064 <.0001 1344 93 Table 3. Phenotypic correlations among P uptake, P use efficiency, and seed yield in a population of 75 recombinant inbred lines from a AND696/G19833 cross grown in high (HP) or low (LP) soil phosphorus in Darien, Colombia in 2005. P P use Seed yield level efficiency P uptake HP -0.51m 0.48m 0*. LP E059“ 0.53 ti. P use HP ------ 0.43 efficiency it! LP ------ 0.24 *Indicates significance at P < 0.05. “Indicates significance at P <0.01. *** Indicates significance at P<0.001. 94 Table 4. Analysis of variance of root traits in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien, Colombia in 2005 in two environments: high and low soil phosphorus. 11 Mean Squares s Source df Adv. Root Specific Root Average roots length root surface root i density lergth area diameter nGenotype (G) 74 5 “S 0.56"" 21‘ 2210'" 0.008‘" dEnvironment (E) 1 142‘" 9.06‘" 31 "S 691 "S 0.065‘" 11G x E 74 7 “S 0.31 "s 13 "5 1538 "5 0.006 "5 “3 indicates not significant * Indicates significance at P < 0.05. “Indicates significance at P <0.01. *** Indicates significance at P<0.001 95 Table 5. Mean root growth traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and G19833 and the means and ranges of 75 recombinant inbred lines (RILS) developed from the parents. Means are also included for the check variety, Carioca. The traits are from the 2005 field experiment planted in Darien, Colombia. Mean values are of 3 replications. P , value indicates level of significant genotypic differences among the RILS for each trait. Parents Recombinant inbred lines Check Traits P AND G mean range p value Carioca level 696 19833 _ (G4017) Adventitious HP 7.2 8.5 9.9 7-12 0.8931 8.2 roots ( #) LP 11.2 ----- 8.8 6-14 0.5291 9.2 Root length HP 1.35 1.1 1.08 0.57-2.30 0.0008 0.70 density LP 1.59 ----- 1.36 0.59—2.59 0.0281 0.66 (cm-cm3-1) Specific root HP ’ 11.1 10.7 8.4 3.3-17.6 0.5144 5.7 length LP 9.71 ----- 8.8 4.0-18.7 0.3887 4.4 (cmmg'l) Root surface HP 87.9 57.3 71.7 33-147.4 0.0040 53.4 area (cmz) LP 603 ----- 74.0 32.2-156.5 0.0289 37.7 Average HP 0.24 0.27 0.33 0.25-0.53 0.0128 0.37 root LP 0.26 ----- 0.31 0.24-0.40 0.3464 0.32 diameter (mm) 96 Table 6. Phenotypic correlations between root traits and P uptake and seed yield in a population of 75 recombinant inbred lines developed from a AND696/G19833 cross grown in high (HP) or low (LP) soil phosphorus in Darien, Colombia in 2005. P Adv. Root _ Specific Root Average level root length root surface root number density length area diameter P HP -004 -0.06 0.16" —0.16" -018'" uptake .. LP 008 0.001 0.16 0.002 -0.01 Seed HP 0.09 -0.098 -0.04 -0.10 -0.05 yield LP 0099 0.03 0.10 0.02 -0.022 *Indicates significance at P < 0.05. **Indicates significance at P <0.01. *** Indicates significance at P<0.001. 97 Table 7. Means of root and shoot traits of 75 RILS developed from a AND696/G19833 cross and grown under high and low soil phosphorus treatments in Darien, Colombia in 2000 and 2005, grouped and averaged by plant growth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.). Traits HigLP Low P Det. Ind. Bet. 2005 Adventitious roots 9.8 a 8.7 a 8.9 a H") Root length density 1.01 b 1.42 a 1.34 a (cm-cm3-1) Specific root length 8.2 a 8.8 a 8.9 a (curing!) Root surface area 63.5 b 76.9 a 71.3 a (cm’) Average root 0.32 a 0.31 a 0.30 a diameter (mm) Shoot dry weight 13.7 a 6.6 a 6.3 a (3 'Plant") P uptake 46.6 a 13.7 a 13.7 a (mg ~plant") P use efficiency 0.30 b 0.47 a 0.52 a (g-seed mg P") Days to maturity 80.6 b 89.4 a 82.2 b Seed P content 186 b 147 a 136 b (mg 100 seed") Seed yield 1435 b 573 a 605 a (kg°ha") 2000 Days to maturity 82 b 88.3 a 83.3 b Seechontent 153b 112a 117a (mg 100 seed") Seed yield (kg-ha") 1551 a 707 a 614 b ‘1 Significant differences are based on Tukey tests and are at alpha = 0.05. Values that do not share the same letter are significantly different. Tests should be read across rows and within each phosphorus treatment level, High P or Low P. 98 Table 8. Phenotypic correlations between root traits and P uptake and seed yield in a population of 75 recombinant inbred lines (RILS) developed from a AND696/G19833 cross grown under high or low soil phosphorus treatments in Darien, Colombia in 2005. For the analysis, RILs were grouped according to plant grth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.). Growth Adv. . Root Specific Root Average habit root # length root surface root density length area diameter Low P P Det. -0.10 0.02 0.20*** 0.02 -0.04 uptake Indet. -0.01 -0.03 -0.08 -0.03 0.05 Seed Det. -0.04 0.13 0.17** -0.10 0.10 yield Indet. -0.27*** -0.17 -0.05 -0.15 ' 0.19 High P P , Det. -0.09 -0.12 0.10 -0.17** -0.14* ' uptake Indet. 0.04 0.03 0.24*** -0.13 -0.27*** Seed Det. 0.15* 0.10 0.14* 0.05 -0.04 yield Indet. 0.04 -0. 18 0.02 -0. 13 —0.005 * Indicates significance at P < 0.05. **Indicates significance. at P <0.01. *** Indicates significance at P<0.001 99 Table 9. Phenotypic correlations between P uptake and seed yield and P seed content and seed yield in a population of 75 recombinant inbred lines (RILS) from a AND696/G19833 cross grown under high or low soil phosphorus treatments in Darien, Colombia in 2000 and 2005. For the analysis, RILS were grouped according to plant grth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.). ' High P-Ind. High P-Det. Low P-Ind. Low P-Det Seed P uptake (2005) yield 0.03 0.18** 0.46*** 0.42*** P seed content 2005 0.17 0.30*** 0.42*** 0.30*** 2000 -0.06 0.12 0.08 0.08 * Indicates significance at P < 0.05. **Indicates significance at P <0.01. *** Indicates significance at P<0.001 100 .mnoumomnon m mo. $226 2: 2a 832 .nnonnoob £80.26? «and 03222235 . fins magnet mo 83? neonn 58232 292.34 ...Een 8.30% 3 consanom 8? 2m mtmm nnfiw some nag? .Ewn 05 no magnum 43 mm una a2 05 no 23an .3 w2n22n2 AA 6.9» 26— m now 43 Bamanom 2a mnounnwnmwu mononvonm mm noun: nouns—Ema Beam Emu ofi no 392» no “on“ $2 2: no £3me 2: no 850% mm .3 E mam mo nonnfihmmo 2F .mom agonnmonn SE 32 one EB 83 a 88 a 2828 gonna a eoeoaa 225 eds 282... oooEop aeo a see eooo_o>oe Gina men: 3.55 “935882 mu 3 awn»: noon 058% end bane“. fimnfl noon mo nounflbmmo zononwonm ._ 03mm Gian—953 bane—o finne— «com md 0N m... . 0... m6 mN . 0N m.—. oé m6 F l r I c Q 4 Es - fim‘ S _ o l N . :fiN‘ l N r. V I V T 0 l m a . 1 l w fi l w t S . . . t or L L , . 1 NF . 1 NF ofimfidhuoonwfi g. 22:55qu § . . I t E EmEEQouc. D t 3 d... n: d 1.2. [or Aouenbeld 101 or mw vw 6358.235 we. 26.55.23 I . ~ n_I Aswan—av finned “com “.59on to? lwv or mr vr NF Ow SmEEnouonc. § . , 22.5.98 U a.— o v . . Hm . 0 IN ..,. to In ..2 ..s lvw ..e ..e . .283 n 8&2 Kouenber :1 102 0850332 m no 0m20>0 06 0.8 0502. .nnonnmob £000 38?», m5 Begun—003 one 2022200 no 003? n88 E2052 032.24 £3. £3on .3 883mg 003 20 mam nnfiw n80 n55? .332 05 no 232» 43 mm one #2 05 no £32» 3 00800052 .5 an? ~22 m mom \3 @2232 20 0555508 xenon—00¢ .mm Bonn, nounnwnmfi Bonm Emu 05 no 232w no fins c2 05 no 322w 05 no gonm mm .5 E mam mo noun—95.0% 2E. .mom 0223023 8.: 32 one And an a moon 5. 2628 .eofin a 383 82 no one 232... copies none a Bee eoooeeoe fig men: 0205 “923882 mu 3 220820 noon 0w20>0 one 02a 00250 noon mo nounnwnmfi mononwnnm .N 0.5me $53 «0.3 000......» 30% . 8F 9: cm. 8. 8 8 8 9:. e: on. . 8F 8 8 2. ac . - . _ .F s O t N m , . WI t N .I V . . .I V t 0 . _ L _ ~ .. 0 t w T 0 fi T2 . i 2 1 NF _ , i a. . V IL 2055.20.05 I .. 3 2058.200... V///////. l t 2 2055.98 I . . . oeosésoo D g . t 9 . i S d: _ fl , . . n... Aouenbelg 103 90.55.3005 ems... 2055.900 I a: Annnv 22255. «00.. 03203» 2.0 . 13. law INN IVN 9058.200... § 32.5.0.8 D d._ s m \ w a m \ m \ m __..._..l .8388 N 035 Kouenbeld 104 .38 S 0253 “on 003 anon—an 0:: no 0000 003000 moon non ooom 80.0 8530 20 0020., 200» mmmeU .0002 .m nmE 2005 205 508 00n002n2 .80 0820., 05 m .52 .00n: 22% 508 00n002n2 .03 nanownom .mmwanw 05 000% 303000 0020 0 22m 0320.60 00n: 00.53 «539802 E. S non—«Egon 05 mo no0000 mnrsonm moon 20 E m :30 2005. 205 0000 .3 m 32 008» 05005 0000 we nom002w0m .m 0.32m .72 9: o :02 5? 203 590 ooou. coup . com? 003 cam? Door com com . h p e .— . — — - b b p b. r - F CON 1 com 9 . m. I 00? U 4. M. . . . ..W r Com 0- e w. . m..- r 08 IOI . M 1 02. d I ) I I I 4 . . mun I 1 000 U. I I . I w B. . '7 0005500005 4 I 5 0005:0200 I t 82 ‘ 105 . .m :30 0005 200A 508 03000500 03 0000.02, 0:0 m 32 000:: 20% 0008 03000090 an. anagram .900 a 0:0 000% 803000 0080 0 800 00020060 00:: 005E 0903:8000." mu E ”800153 06 .00 80000 wEBSw CO8 08 E m :mE 000:: 2000 0000 3 m 32 00:: 200A 0000 mo 03000Hw0m .0 0.5mm A 72 9; a :02 £3, 0.0.; 5.20 _ 800 800 88 89 82 80 P _ p p F L 5’ , . . n b 80 4 I .. 80 f. ‘ I f mu . 08 m. . U u 80 .M. w. r 80 D. . .M: w con 1U... n W I . 80 M .. 80 d 1 82 .W . u. u 8: e j V . 80F ( 0005500005 0. . . 0005:0900 I .180? u 83 106 Chapter 4: Identification of QTL related to Root Architecture Traits and Low Phosphorus Tolerance in an Andean Bean Population Abstract Low soil phosphorus is a major constraint to common bean production. The identification of QTL (quantitative trait loci) for phenotypic traits associated with tolerance to low phosphorus soils will aid in the development of cultivars that can perform well in such environments. Root architecture was identified as a determinate of P uptake in a cross between an Andean tolerant and Mesoamerican susceptible bean genotype. This experiment was conducted to determine if root architecture aids in tolerance to low P soils in Andean intra gene pool cross between a tolerant (G19833) and susceptible (AND696) genotype. A linkage map was developed from 77 F57 recombinant inbred lines of Gl9833/AND696 using SSR, AF LP and RADP markers with 12 linkage groups for a total of 633cM. QTL were identified for root length density (cm/cm3) and root surface area in the top 16 cm of soil. The QTL were derived fi'om G19833, but did not co-localize with QTL for P uptake. A QTL for P use efficiency, derived from G19833 was identified on linkage group B11. QTL were identified for seed yield on linkage group B1 near the fin gene for determinacy, while G19833 was the source of the increased effect in the low P environment in 2000, AND696 was responsible for the yield increase in this location in the high P environment in 2005. 107 Introduction Quantitative trait loci (QTL) analysis is a powerful tool for understanding genetic variation and control of complex traits. QTL are regions of the genome statistically associated with phenotypic variation of a quantitative trait (Doerge, 2002). QTL identification requires phenotypic data for a trait of interest in a segregating population and a genetic linkage map (Collard et al., 2005). . . Mapping of QTL permits the identification QTL controlling multiple traits in numerous genetic backgrounds. This technique also can be used to dissect epistatic interactions and predict genotypic by environment interactions at the genetic level (Dekkers and Hospital, 2002). The identification of QTL serves as a starting point for marker assisted selection (MAS), which offers potential to improve traits with low heritabilities or difficult to measure traits without a heavy reliance on phenotypic selection (Collard et al., 2005). QTL mapping has also been successfully employed as a first step to identify genes underlying phenotypic variation for a trait of interest via positional cloning (Li et al., 2006). In common bean (Phaseolus vulgaris L.) QTL studies have been conducted for the major diseases, insects, and abiotic stress that limit productivity. To date, QTL analysis for disease resistance has been the most fruitful area of research, and MAS based on QTL studies for bean golden mosaic virus and common bacterial blight resistance are currently employed in breeding programs (Miklas et al., 2006). . QTLanalysis for abiotic stresses affecting common bean has yet to reach the level of success of that seen with biotic stress, nonetheless, this tool is 108 currently being used to identify underlying mechanisms of tolerance to low P soils (Yan et al., 2004; Beebe et al., 2006). Phosphorus deficiency is one of the most prevalent stresses, especially in bean growing regions of the tropics. An estimated 60% of bean production in Latin America and Africa occurs under deficient soil P conditions (Wortman et al., 1 998). Tolerance to P deficiency has been identified in bean lines from diverse genetic backgrounds in both the Andean and Mesoamerican gene pools (Beebe et al., 1997). A number of Peruvian landraces, including G19839 and G19833, have been found to grow especially well in low P soils (Y an et al., 1995a; 1995b). Strong genotype x environment interactions for seed yield, however, necessitate the use of other indicators of tolerance to low P soils that exhibit higher heritabilities (Beebe et al., 1997). A number of physiological traits correlated with tolerance to low P soils have been identified. In G19833, research has shown P efficiency to be related to greater P uptake from the soil. RoOt system phenotypes that exhibit shallower basal root angle, greater total root length and root surface area, and root length of basal roots in top 3 cm have been shown to enhance P uptake in P limiting environments (Bonser et al., 1996; Liao etal., 2004; Beebe etal., 2006). The measurement of root traits in the field is laborious and root growth is very plastic even with small changes in the soil environment (Snapp et al., 1995), making selection for plants with desired root traits challenging. The identification of QTL for root traits that improve P uptake in low P soils is a first 109 step to conducting MAS for P efficiency. Studies (Beebe et al., 2006 and Liao et al., 2004) using G19833 as the efficient parent in crosses with DOR364, a Mesoamerican small seeded black bean as the inefficient parent, have identified QTL for root traits that co-localize with QTL for P uptake efficiency. QTL for both root length and P uptake in low P soil were found in the same region of linkage group B4 with r'2 values of 0.21 and 0.13, respectively. Additional QTL for specific root length and P uptake under the same environmental conditions co- localized to a region of linkage group B10 and had 12 values of 0.19 and 0.14, respectively (Beebe et al., 2006). Liao et al. (2004) also identified QTL on linkage group B4 in the G19833/DOR364 population for percent of basal roots in the top 3 cm in a growth pouch assay that co-localized with QTL for P uptake in the field. The identification of QTL for root architecture traits with QTL for P uptake efficiency in crosses between the Andean and Mesoamerica gene pools raises the question if similar mechanisms for P efficiency would be observed in an Andean x Andean cross using the same P efficient parent, G19833. The first objective of this study was to construct a linkage map of two Andean common bean lines with different tolerance to low P soils. The second objective was to identify QTL related to tolerance to low P soils using an Andean common bean recombinant inbred line population. 110 Materials and Methods: Plant Material The common bean genotypes G19833 and AND696 from the Andean gene pool were used as parents to develop 77 F57 recombinant inbred lines (RILS). The initial cross was advanced to the F5 generation by single seed descent. G19833 is a Peruvian landrace with an indeterminate grth habit and large yellow and red mottled seed with an average 100 seed weight of 41 g. G19833 has also been identified as tolerant to low P soils (Yan et al., 1995ab). AND696 is a CIAT improved line from the race Nueva Granada with a determinate growth habit and large red and white mottled seed with an average 100 seed weight of 51 g. AND696 has been identified as susceptible to low phosphorus soils (CLAT, 2000). The RILs developed fiom G19833 and AND696 segregated for determinacy, with 56 determinate and 21 indeterminate lines. Growth habit in common bean is controlled by a single gene, fin (Norton, 1915). The number of determinate and indeterminate lines deviated significantly from a 1:1 ratio that is expected for a single gene trait. This deviation is likely due to the selection of RILS with uniform days to maturity, which biased the selection to favor determinate lines. Field Trials In 2000 and 2005, the 77 RILS, G19833, AND696, and two check varieties, G4017 (Carioca) and G16140 were planted in Darien, Colombia in a 9x9 lattice design with three replications at two soil P levels, low P (45 kg/ha triple super phosphate) and high P (300 kg/ha triple superphosphate). Phenotypic 111 data was collected on 75 of the 77 RILS. The soil of this site in an Andisol with a native soil P of 2 mg/kg based on bray II extraction method. Seed was hand- planted in 4 row plots where each row was 3 m long. During the 2005 growing season, various plant measurements were taken to elucidate underlying factors involved in the parental genotypes differences in tolerance to low soil P. Plant measurements include adventitious root number at 3 weeks after planting, shoot dry weight and total P concentration at mid pod (which ranged fi'om 54 to 67 days after planting), root length and weight to a depth of 16 cm below ground at mid pod fill (Chapter 3, Materials and Methods). Seeds reached maturity at 76 to 98 days after planting, as determined by seed dryness. At maturity, seeds were hand-harvested. Seeds were dried and seed yield was determined at 18% moisture. Seed weight of 100 seed was measured at 18% moisture for the 2000 and 2005 plantings. A sub sample of seed from the 2000 and 2005 harvest was acid digested and analyzed for total P concentration according to the method of Murphy and Riley via colorimetric spectrometry (1962) DNA Isolation and Molecular Marker Analysis Plant tissue was harvested from 5 plants of greenhouse grown seed of G19833, AND696 and 77 F57 recombinant inbred lines. Total DNA was extracted from plant tissue with 24:1 Chloroform: Isoarnyl alcohol. The rniniprep procedure was carried out according to the method of Edwards et al. (1991). DNA was quantified with a fluorometer (Hoefer DyNA Quant 200, San Francisco, CA). Molecular markers screened in the study for polymorphisms between the 112 parents (319833 and AND696 included 125 simple sequence repeats (SSR) (Metais et al., 2002; Blair et al., 2003; Gaitan-Solis et al., 2002; Yu et al., 2000; Caixeta et al., 2005), 50 Random amplified polymorphic DNA (RAPD) (Integrated DNA Technologies, Inc., Coralville, IA) and 41 amplified fragment length polymorphisms (AF LP) primer combinations. SSR Markers Amplification was performed with 2 pl of DNA diluted to 20 ng°pl", 0.2 pl of primer, 0.15pl of Taq polymerase, 1.2pl (2.5mM) MgClz , 1.2pl (10x) PCR buffer, and 0.12 pl of a 10mM mix of dNTPs. PCR was conducted in a 96 well FTC-100 Programmable Thermal Controller (MJ Research, Inc., Waltham, MA). The therrnocycler was programmed for 1 cycle of 5 minutes at 94°C, followed by 30 cycle of 1 minute at 94°C, 1 minute at 47°C, and 1 minute at 72°C, and a final extension step at 72°C for 5 minutes. Double stranded DNA was denatured for 5 minutes at 94°C. PCR amplification products were separated on 4% polyacryilimide gels and DNA bands were visualized with silver nitrate according to the procedure of Blair et al. (2003). RAPD markers Amplification was performed with 5 pl of DNA diluted to 10 ng°pl'l, 1 pl of primer (Integrated DNA Technologies,-Inc., Coralville, LA), 0.3pl pf Taq polymerase, 25 pl (2.5mM)MgC12 , 25pl (10x) PCR buffer, and 0.5pl of a 10mM mix of dNTPs. PCR was conducted in a 96 well FTC-100 Programmable Thermal Controller (MJ Research, Inc., Waltham, MA). The thermocycler was programmed for 2 cycles of 1 minute at 91°C, 15 seconds at 42°C, and 1 minute 113 and 10 seconds at 72°C, followed by 38 cycles of 15 seconds at 91°C, 15 seconds at 42°C, and 1 minute and 10 seconds at 72°C, and a final extension step at 72°C for 5 minutes. A total of 20 pl of PCR product were loaded ontol .5% agarose (diluted with 0.5x TBE buffer) gels with ethidium bromide and ran in electrophoresis chambers with 1.5L 0.5x TBE buffer for 3 hours at 110 volts. PCR amplification products were visualized under UV light. RAPD markers were named according to the primer used and the molecular weight of the band in kilobases. AFLP Markers AF LP reactions were performed according to the procedure described by Vos et al. (1995), using a commercially available kit (AFLP analysis System I, Invitrogen Corporation, Carlsbad, California) and following the manufacturer’s instructions. The digestion of 500 ng of DNA was performed with 2 pl EcoRI/MseI (1.25U/pl), incubated for two hours at 37°C. The restriction fragments were ligated to enzyme adapters using 1pl of T4 DNA ligase (1 U/ pl) incubated for two hours at 20°C. The prearnplification (primer +1 base) was performed with 2.5 p1 of DNA diluted 1:10 from the digestion-ligation product, 20 pl preamplification primer mix, 2.5 10x PCR plus Mg and 0.15 pl of DNA polymerase was added. PCR conditions were 20 cycles of 94°C for 30 seconds, 56°C for 60 seconds, and 72°C for 60 seconds. The PCR product was run in 1% agarose gel and diluted 1:50. The amplification (primer +3 bases) was performed with 5 pl of DNA's dilution 1:50 and 5 pl of Mix I (primers and dNTP's) and 10 p1 Mix II (10x buffer, MgCl and DNA polymerase). The PCR conditions were one 114 cycle at 94°C for 30 seconds, 65°C for 60 seconds, and 72°C for 60 seconds, then the annealing temperature was lowered 0.7°C per cycle for 13 cycles and 23 cycles at 94°C for 30 seconds, 56°C for 60 seconds, and 72°C for 60 seconds. PCR amplification products were separated on 4% polyacryilimide gels and DNA bands were visualized with silver nitrate according to the procedure of Blair et al. (2003). The AF LPs were named by combining the last 2 selective bases ligated to EcoRl with the last two selective bases ligated to Msel in the amplification step. Genotypic Data Analysis Linkage analysis was conducted with the software J oinMap 3.0 for Windows (Van Ooijen and Voonips, 2002) set to Kosambi’s map fiinction. Kosambi’s mapping function assumes the existence of interference that is negatively related to recombination frequency. Analysis was conducted with marker data of 160 molecular markers segregating in the population of 77 recombinant inbred lines. Mapping parameters were set to recombination frequency smaller than 0.300 and a LOD score larger than 3.0. Parameters were relaxed to a recombination frequency of smaller than 0.45 and a LOD score greater than 2.0 to join unlinked markers and to join fragmented linkage groups. LOD (log of odds) scores are = log (Ll/Lo), where L1 is the likelihood for the alternative hypothesis and L0 is the likelihood of the null hypothesis. A LOD score of 3 means the alternative hypothesis is 1000 times more likely than the null hypothesis. Linkage groups were identified according to location of SSR markers (Blair et al., 2003) and were named according to the core bean linkage map (Freyre et al., 1998). 115 QTL analysis was conducted with the genetic map developed with the J oinMap program and with the phenotypic means for each RIL collected from the field study. The computer software program Windows QT L cartographer version 2.5 (Wang et al., 2006) was used to identify QTL for root length density, root surface area, shoot P concentration, P uptake, seed yield and seed P content. The Composite Interval Mapping (CIM) feature set to a window size of 100M and with forward and backward regression model was used to identify QTL. CIM involves the use of maximum likelihood estimates and linear regression to identify QTL within marker intervals. Significant QTL were considered by defining the LOD score at p=0.01 afier 1000 permutation tests. Results and Discussion: Linkage Map DNA polymorphism levels with the molecular markers used in this study were low to moderate between the parent lines 619833 and AND696. Of the 125 SSR markers screened, all of which were developed for use in common bean, only 26.4% were polymorphic between the parent lines. Low polymorphism levels are often seen in intra gene pool crosses such as this one between two Andean genotypes. A recent study of SSR polymorphism levels compared 129 markers in 44 common bean genotypes fi'om both Andean and Mesoamerican gene pools, and in the study, the average inter gene pool polymorphism level (Andean x Mesoamerican) was 59.6% and intra gene pool polymorphisms were 37.9% (Blair et al., 2006). 116 The 103 locus linkage map developed from the SSR, RAPD, and AF LP markers screened in the G19833/AND696 RIL population spaned 633 cM. Common bean has 11 linkage groups corresponding to the genome’s 11 ' chromosomes. In this study, 12 linkage groups were identified, 10 of which were able to be identified and named according to the bean consensus map, based on the placement of SSR markers (Freyre et al., 1998; Blair et al., 2003) (Figure 1). Linkage group BS is missing from the map, although the identities of Groups A and D, (which are comprised solely of AF LP markers), have yet to be determined in relation to the consensus map. QTL identification Using composite interval mapping, 21 QTL were identified for 8 traits in 12 marker intervals on 7 linkage groups. One-third of the QT L were clustered on linkage grOups B1 and affected more than one trait. An additional 19% of the QTL were located on linkage group B6 and also affected multiple traits. Individual QTL explained 12 to 45% of the phenotypic variation, and total phenotypic variation explained for any one trait was 14 to 60% (Table 1). Root Architecture Frequency distribution graphs for four root traits measured at mid pod fill are shown in Figure 2. Means of RILS and parental genotypes for each trait are available in Chapter 3, Table 4. The mean root length density of the RILS was greater in the low P environment (1.36 cm/cm3) than in the high P environment (1.08 cm/cm3). QTL identified for root length density in the high P environment were distinct from the QTL identified for the same trait in the low P environment 117 (Table 1, Figure 6). There were no significant differences for root surface area based on P treatment. However, QTL for this trait were only detected for the high P environment. The QTL found for root surface area in high P mapped to the same linkage groups as the QTL for root length density in high P. The increased effect of each of the root traits was derived fiom G19833 (Table 1). A QTL for specific root length in greenhouse grown RILS from the population G19833/DOR364 was identified in the same region of linkage group B1 (Beebe et al., 2006) as QTL for root length density and root surface area under high P described here. Additional QTL for root traits in the G19833/DOR364 population were identified in the same region of B3 as the QTL for root length density in low soil P identified in this study. They include a QTL for taproot length in greenhouse grown RILS (Beebe et al., 2006) and a QTL for taproot root hair _ . length in solution culture grown plants (Yan et al., 2004). None of these QTL, however, overlap with QTL for P uptake. P uptake Shoot P concentratiOn and P uptake distribution among the RILs was affected by the P level of the soil (Figure 3; Chapter 3, Table 2). The range of values for P uptake was greater in the high P than the low P soil (Figure 3), suggesting luxury consumption of P occurred in some RILS in the high P treatment. QTL were detected on B8 for shoot P concentration and P uptake in the low P environment (Table 1; Figure 6). A QTL for P uptake in the high P environment was found on Bl 1. The increased effect of each of these traits was conferred by AND696. No QTL for P uptake were identified with the 2005 field 118 data where G19833 increased P uptake. This result is contrary to results found with the G19833/DOR364 population where QTL for p uptake from G19833 were identified (Beebe et al., 2006; Yan et al., 2004). This result, however, may be specific to the 2005 field season, as suggested by the 2000 field season data where a QTL was identified for yield in the low P environment derived from G19833 (Table 1). Seed Yield Frequency distribution graphs for seed yield in 2005 and 2000 show the large variability for yield across RILS and soil P levels (Figure 4, bottom; Chapter 3, Table 2). In the low P treatment in 2005, AND696 yielded less than the mean of the RILs (465 and 597 kg-ha'l, respectively). In the high P treatment in 2005, AND696 had the same yield as the mean of the RILs (1347 kg-ha‘l). One QTL was detected for seed yield in 2005 which explained 27% of the phenotypic variation. It was observed only in the high P treatment, and AND696 conferred the increased effect on yield (Table 1). In 2000, one QTL was found for yield in low P and one was found for yield in high P, explaining 17 and 45% of the phenotypic variation, respectively. Each QTL was derived from G19833 (Table 1). The QTL identified for yield under the high P treatment in 2005 and that for yield under the low P treatment in 2000 both map to the same region of B1 and the increased effects on yield for these QTL were derived from different parents . (Table 1; Figure 6). The region of Bl where these QTL are located is the same region where the single gene (fin) responsible for deterrninacy is located (Koinange et al., 1996), as the AND696/G19833 population segregated for 119 determinacy. The efficient parent, G19833 exhibits indeterminate plant growth and the inefficient parent, AND696, exhibits determinate plant growth. These differences were associated with differences in plant maturity in the population, such that the indeterminate plants averaged 5 days longer to reach maturity (Chapter 3, Table 7). Linkage group B1 has previously been shown to be important in common bean domestication, carrying both the fin gene and the de gene for photoperiod sensitivity. QTL for domestication syndrome, including earliness and seed size traits have been identified on this linkage group, near these genes (Koinange et al., 1996). The importance of B1 near the fin gene suggests either a pleiotrophic effect of this gene on yield, or linkage of this gene with other genes important for yield. The observation that this region of the bean genome had an opposite effect on seed yield under high and low soil P levels may be an indication that growth habit itself is a key mechanism to improve tolerance to low P soils, perhaps by increasing days to maturity and allowing more time for P uptake from the soil. The cluster of domestication genes on B1 have pleiotrophic effects on other traits in common bean, including leafhopper resistance (Murray et al., 2004). P use efficiency P use efficiency (PUE) was defined in this study as the amount of seed yield per unit of P taken up by the plant. In 2005, PUE was found to be higher in the low P environment than in the high P environment (Figure 5; Chapter 3, Table 2). AND696 was at the low end of the distribution for this trait under low P, with an efficiency of 0.27 compared to the mean of the RILs at 0.51 (Figure 5). A 120 QTL for PUB was detected under high P on El] that was positively contributed by G19833 (Table l). PUE was not an important mechanism for tolerance to low soil P in the G19833/DOR364 population (Beebe et al., 2006). PUB may be more important for tolerance to low P soils in intra gene pool crosses like G19833/AND696 compared to inter gene pool crosses. Preliminary QTL studies with the P efficient bean line G21212 identified PUE as a factor in tolerance to low P soils (Miklas et al., 2006). Seed P content P content of the seed may serve as an indicator of tolerance to low P soils. In P limiting environments, greater levels of P in the seed may aid early plant growth (Lynch and Beebe, 1995). In the low P soil treatment, seeds contained less P than in the high P environments in both 2000 and 2005 (Figure 4, top; Chapter 3, Table 2). P content was correlated with seed yield (1:030 p value <0.0001) in 2005 in the low P environment, but was not correlated with yield in any other environment/year combination. QTL identified for this trait were identified on B1, B2 and B6 across environments and years, and in all cases an increase in seed P content was derived from G19833 (Table 1, Figure 6). Although the QTL for seed P content were consistent across years and environments, based on phenotypic correlation with seed yield it does not appear to be a valuable trait for selection to improve tolerance to low P soils. Conclusions: QTL were identified for root growth traits including root length density and root surface area. These QTL were derived from the P efficient parent, 121 G19833, and co-localized with QTL for root traits identified in previous studies. The QTL for root traits identified here did not, however, co-locate with QTL for P uptake, and no QTL for P uptake detected in the current study were derived from G19833. - QTL for yield in low P 2000 and high P 2005 mapped to the same region of the genome (B1), but were derived from different parents. Plant grth habit appeared to play an important role in determining yield in the low P environment in this population. QTL for root length density, root surface area, and seed yield were also found in the same area of B1. The clustering of QTL in this region suggests pleitropic effects of the fin gene on a diversity of plant traits, and raises questions to the value of Selecting QTL in this region of the genome, to improve tolerance to low P soils. This assertion is supported by regression of yield data by soil P treatment, that identify determinate lines that yield better than indeterminate lines under both soil P levels (Chapter 3, Figures 3 and 4). Genetic differences in P use efficiency were identified in this p0pulation. QTL across years and environments were identified for seed P content. 122 References: Beebe, S., J. Lynch, N. Galwey, J. Tohme, and I. Ochoa. 1997. A geographical approach to identify phosphorus-efficient genotypes among landraces and wild ancestors of common bean. Euphytica 95:325-336. Beebe, S.E., M. Rojas-Pierce, X.L. Yan, M.W. Blair, F. Pedraza, F. Munoz, J. Tohme, and J .P. Lynch. 2006. Quantitative trait loci for root architecture traits correlated with phosphorus acquisition in common bean. Crop Science 46:413- 423. Blair, M.W., M.C. Giraldo, H.F. Buendia, E. Tovar, M.C. Duque, and SE. Beebe. 2006. Microsatellite marker diversity in common bean (Phaseolus vulgaris L.). Theoretical and Applied Genetics 113:100-109. Blair, M.W., F. Pedraza, H.F. Buendia, E. Gaitan-Solis, S.E. Beebe, P. Gepts, and . J. Tohme. 2003. Development of a genome-wide anchored microsatellite map for common bean (Phaseolus vulgaris L.). Theoretical and Applied Genetics 107:1362-1374. Bonser, A.M., J. Lynch, and S. Snapp. 1996. Effect of phosphorus deficiency’on grth angle of basal roots in Phaseolus vulgaris. New Phytologist 132:281-288. Caixeta, ET, A. Borem, and J .D. Kelly. 2005. Development of microsatellite markers based on BAC common bean clones. Crop Breeding and Applied Biotechnology 5:125-133. CIAT. 2000. Inheritance of low phosphorus tolerance in the Andean population AND696 x G19833. Bean Improvement for the Tropics Unit Annual Report. CIAT, Cali, Colombia. Collard, B.C.Y., M.Z.Z. Jahufer, J .B. Brouwer, and B.C.K. Pang. 2005. An introduction to markers, quantitative trait loci (QTL) mapping and marker- assisted selection for crop improvement: The basic concepts. Euphytica 142: 169- 196. Dekkers, J .C.M., and F. Hospital. 2002. The use of molecular genetics in the improvement of agricultural populations. Nature Reviews Genetics 3:22-32. Doerge, R.W. 2002. Mapping and analysis of quantitative trait loci in experimental populations. Nature Reviews Genetics 3:43-52. Edwards, K., C. J ohnstone, and C. Thompson. 1991. A Simple and Rapid Method for the Preparation of Plant Genomic DNA for Pcr Analysis. Nucleic Acids ' Research 19:1349-1349. 123 Freyre, R., P.W. Skroch, V. Geffroy, A.F. Adam-Blondon, A. Shinnoharnadali, W.C. Johnson, V. Llaca, R.O. Nodari, P.A. Pereira, S.M. Tsai, J. Tohme, M. Dron, J. Nienhuis, C.E. Vallejos, and P. Gepts. 1998. Towards an integrated linkage map of common bean. 4. Development of a core linkage map and alignment of RF LP maps. Theoretical and Applied Genetics 97:847-856. Koinange, E.M.K., S.P. Singh, and P. Gepts. 1996. Genetic control of the domestication syndrome in common bean. Crop Science 36:1037-1045. Li, C.B., A.L. Zhou, and T. Sang. 2006. Rice domestication by reducing shattering. Science 31 1 :1936-1939. Liao, H., X.L. Yan, G. Rubio, S.E. Beebe, M.W. Blair, and J .P. Lynch. 2004. Genetic mapping of basal root gravitropism and phosphorus acquisition efficiency in common bean. Functional Plant Biology 31:959-970. Lynch, J .P., and SE. Beebe. 1995. Adaptation of Beans (Phaseolus-Vulgaris L) to Low Phosphorus Availability. Hortscience 30:1165-1171. Metais, I., B. Harnon, R. J alouzot, and D. Peltier. 2002. Structure and level of genetic diversity in various bean types evidenced with microsatellite markers isolated from a genomic enriched library. Theoretical and Applied Genetics 104:1346-1352. Miklas, P.N., J .D..Kelly, S.E. Beebe, and M.W. Blair. 2006. Common bean breeding for resistance against biotic and abiotic stresses: From classical to MAS breeding. Euphytica 147:105-131. Murphy, J ., and J .P. Riley. 1962. A modified single solution for determination of phosphate in natural waters. Anal Chem Acta 27:31-36. Murray, J .D., T.E. Michaels, C. Cardona, A.W. Schaafsma, K.P. Pauls, 2004. Quantitative trait loci for leafhopper (Empoascafabae and Empoascakraemeri) resistance and seed weight in the common bean. Plant Breeding 123:474—479. Norton, J .B. 1915. Inheritance of habit in the common bean. Am. Nat. 49:547-561. Plant Research International BV. 2002. JoinMap. Release 3.0. Plant Research International BV. Snapp, S., R. Koide, and J. Lynch. 1995. Exploitation of localized phosphorus- patches by common bean roots. Plant and Soil 177:211-218. 124 Statistical Genetics, North Carolina State University, USA. 2006. Windows QTL Cartographer. Release 2. 5. Statistical Genetics, North Carolina State University, USA. Vos, R, R. Hogers, M. Bleeker, M. Reijans, T. Vandelee, M. Homes, A. Frijters, J. Pot, J. Peleman, M. Kuiper, and M. Zabeau. 1995. Aflp - a New Technique for DNA-Fingerprinting. Nucleic Acids Research 23:4407-4414. Wortmann, C.S., R.A. Kirkby, C.A. Eledu, and DJ. Allen. 1998. Atlas of common bean (Phaseolus vulgaris L.) production in Africa CIAT, Cali, Colombia. Yan, X.L., J .P. Lynch, and SE. Beebe. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .1. Vegetative Response. Crop Science 35:1086-1093. Yan, X.L., S.E. Beebe, and J .P. Lynch. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .2. Yield Response. Crop Science 35:1094-1099. Yan, X. L., H. Liao, S. E. Beebe, M. W. Blair, and J. P. Lynch. 2004. QTL mapping of root hair and acid exudation traits and their relationship to phosphorus uptake in common bean. Plant and Soil 265: 17- 29. Yu, K., 8.]. Park, V. Poysa, and P. Gepts. 2000. Integration of simple sequence repeat (SSR) markers into a molecular linkage map of common bean (Phaseolus vulgaris L.). Journal of Heredity 91:429-434. 125 Table 1. Putative QTL for seed traits identified from 75 recombinant inbred lines developed from AND696/G19833 cross grown under high (HP) and low (LP) soil phosphorus conditions in Darien, Colombia in 2000 and 2005. Traits Linkage Nearest LOD 1?.2 AdditivityF group markerT scoreI CIM¥ Root length density LP 2005 B3 AGl 3.1* 0.14 -0.17 HP—2005 B1 AGT A01 4.7 0.18 -0.15 D AGTA02 4.5 0.20 -0.15 Root surface area HP 2005 B1 fin 5.6 0.21 -12.1 D AGTA02 4.2 0.24 -12.0 Shoot P LP 2005 B8 CTT A04 3.5 0.17 0.08 P uLtake . LP 2005 . B8 012.2200A 2.9* 0.12 1.54 HP 2005 B11 ACACO4 3.4 0.15 4.8 P use efficiency HP 2005 B11 R4.1400A 3.1* 0.12 -0.03 Seed yield HP 2005 B1 AGTAOI 6.4 0.27 155 LP 2000 B1 ATA4 3.5* 0.17 -83 HP 2000 B6 BM170 4.3 . 0.45 -301 Seed P content LP 2005 B1 . fin 8.0 0.28 -9.6 B2 BM164 4.9 0.19 -7.2 B6 GCTC03 4.5 0.13 -6.0 HP 2005 B1 fin 9.0 0.27 -l3.9 B6 GCTC03 6.8 0.19 -10.9 LP 2000 B2 BM164 3.7 ‘ 0.13 -5.0 HP 2000 - B1 ATA4 6.0 0.18 -12.6 B2 ' BM164 , 5.5 0.15 -10.8 B6 GCTC03 3.4 0.09 -8.4 1' Primer information for each marker can be found in Tables A4-1, A4-2, and A4-3 of the Appendix. I LOD: Log ofodds ¥ Proportion of the phenotypic variance explained by QTL at test site using CIM (composite interval mapping). F Effects of substituting a single allele from one parent to another. Positive values indicate that allelic contribution is from AND696 and negative from G19833. * Indicates that the LOD score fell below the cutoff range of 1000 permutations at p = 0.05. 126 #00593. 05 .00 ntv< 0:0 $-30 . 70¢. 00308 E 0:00.“. 00 :00 0:03 0000 00 0:00.00E :00 8.008003 00% 000.0% 0mg N: :0 2020 .00 038 0 00.0 00:08 0:30:00: 0:0 0:0 0:00:08 00300—05 Mmm .0302 fig 0:00:00 :08 00:. 0000300: 30 20 0e 800 85: .850882 00 :80 0002300 00020 .3 00002.... 0e .08 0000:. .080 8.0an ._ 200.: 00.20 0.0.. .0020 0.00. .850 ..00 .0920 ..00 .0050 0.0: . 351-1... 0000.00.60“... 00:20 0.0 080.00. 0 00 08.5.0 .117 0.00 808.0 \ / ...00 09204. /D ~00 . (80.92 0.3 @033 . adv . . 8520/1\000 _ 00:20 /t\00.. .Sfii .. 0.0.. . 8020 \i/ 0.0.. 080.00 0.00 .0220 \ / 0.00 . (and—.2 0.8 ”min ...om ~008< 0.00 . . . .0<<0< _ H 0.0 .250 . 0.00 .3 a 00 .00<0< 0.0. 0280 0.0 8.20 0.: . 8200 ..0. 0850 00 .. 22 0 0. .350 0.0 820.. JR .. . =0 1\ 0.0. 0. .0< 08.20 0.0 09200 0 0 .0502 Url 0.0 0020 0.. . .0980E 03.20 . . 8<<0< \ 0.. 00.20 0.0 2.20 0 0 0000.: 0. 02.2 0 0 3:20 o 0 0.0.20 ml 0.0 . 08.02. 0.0. _ 0o0 05 0.00 0:002 Amoom S :3on won 003 mmwm 3% 000080005 E000b% .0025 009.3000» 300.000 00 0030.. 088 080202 0305 .08 03000800 Ed 32 000 SE 00.0 a 0000 0. 0.08200 .0005 a 0300.0 0000.0 200 30% 0.003009 0.080 0 88.”. 0000300000 00:: 00.5.: 02980000 m0 3 0:00. 88 mo nova—p.520 00000000....“ .N 0.00mi Ava 83 A 053.23 .0000. .09 00.0000 . £0000 £000.. .80 av or 3. N_. or o v MN 9N . m._. 0.0 m6 g [or \\\\\\\\\\\\\\\\\\\\‘ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ a) &\\\\\\\\\\\\\\\\\\\\\\\\\\V a: 00002.. q . _ # H .0: 0000.0 ‘ -0. n... 000nz< -0. m 0.. 00002< - 00 ‘ .u0 0 -00 a: 00002... 0: 0000.0 . - 00 a... I .-00 . -00 r00 . . -00 129 Figure 2 (cont’d). 0.40 0.45 0.50 0.35 8 o. 00:! _|L CO 83. 33 CE3 0' I l I l'-l ‘ l 'l {-‘o’ 8 (“0’ 8 2‘3 53 ‘° ° 0 $2 3 O 9.‘ n. :I: 8 8 (D- '- c: z < —’ 8 tL—> 8 .1 § . g 8 < L I I'I I I I‘ I :3 8 a a :2 2 *0 ° Aauenbau 130 Average root diameter (mm) Root Surface Area (cmz) 00.00.003.000 m .00 0w000>0 00. 000 000002 .308 00. 030% .00 003 mmweg 000000000000 00000th 00.00: 00050000m 00000000 .00 0030.0 0000.0 0000000000 030.004 ...00 0000030000 Ed 30. 00.0 ENE. .33 00. moom 3 0.0.8200 .0055 00. 0000.03 mmwmfiu .000 mace/2 0003000— 00000 0 800.0 0000—2000 00.0: 00.500 00000050000000 m0 0. 00300 0000000000 .0000 0000000000000 0000000000 00000 .00 0000055030 50000000. ”m 000mE ..-mx 9 A EmifiEV T . .0000 n. .0020 00.00000 . . 00 00 0. 00 00 00 00 00 0. 0 0.0 0.0 0.0 0.0 0.0. 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.. 0 \ \\\\00 r 0 \ \\\\\ N 0 0“““\ ““0““ . 0 ““0““ -0 . 00.00006 ““0““ . 0 x... .. xxxx . “0““ -0 n... 00002.... 0 ““\\ . \\\\ o. \.\\\ “0“ -0. \\\ . 0““ 0. 0“ -0. . .0 r .. “0 - n... 00002< I 0. mm . 0. 00.000020. “0 -0. 0“ . fl 0. I m“ -0. 0... 0000.0 00 a... .0 . 0100002... 0... § m. .-00 . 00 T00 Aouanbeu 131 cow 3N ONN 8N our Dow 01' fig Doom 0.... mmmeO 1 .\\\\ .9: .\\\\\\‘ k\\\\\\\\\\\\\\\\\\\\\\\\\V 000000000000 m .00 0m000>0 000 000 000002 .808 000 039% 000 003 000972 0000 moom 000 0300» 0000 003 300000 0000000000 00.000.00.00 00000000 000000000w 0000000000 00200, 000000 0000000000 03000< .08 0000000000 00.0 30. 000 €0.00 00.... a 0000 000 0000 0. 00080.00 00.000 00 02.000 000000 000 000072 083.00 00000 0 8000 0000005000 00000 0000000 000000080000 00. 000 200% 00000 00000 00000000 0000000000 00000 .00 000000000000 000000000000 .0 000w00 00.0000. 000.000 000.08 0 0000 ON k\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ F cow r.- on P ;\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ 0..0 000. 0.0 000 000 00. 00. 00. 00. 0..0. P. h I co co 0) P (D a $0. I . —> k\\\\\\\\ to W W E E . 0\\\\\ 0.... 00002... 0 n... 00002< mooN I l I N O l V T l (D V I Q Aouenbau Wop MN? r3 ..0. ..0. H.00 ..00 new ~ 132 2000 A‘\‘ G19833 HP —> I““\\‘ s\\\\\\\\\\\\\\\\\\\\‘ 0‘\\\\\\\\\\\\\\\\\\\\\\\\\\‘ u\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ l G19833 LP L\\\\\\\\\\\\\““\‘\\‘\\\\\\\\\\‘\V L\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ 1— s\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ s\\\\\\\\\\\\\\\\\\\\‘ m 400 600' 800 1000 1200 1400 1600 1800 2000 2200 2400 h U I I ' T T (D V T N O 16- 14- 12- 1o- 8 1800 2005 AN0696 HP 1400 1600 1200 000 1 800 AN0696 LP 600 955. ‘400 16 14 12 O 0 (D V N O AouenbeJd . Figure 4 (cont’d). 133 Seed yield (kg-ha") .8053:qu m mo ommbaw 05 Pa .382 .Amoom S :3on «on 33 @339 35880.: Sacha noun: anabonom 333m mo 83S, Qua 38298 955 48m mEoanoag. Ed 32 23 Ag :mE S moon E «58200 .559 a 353 288 93 8324 80an 3.20 a 88 833% m8: 855 “55988“ mu a 983 05 E a: :33 A no :5. Ba 20% coca mo 2525 08 8 33m £5 E @0303 homomolmuo om: aonmmoam mo nousawammu 5893.5 .m Pawwm Arm 9: 33.9 3:205» .3: n. Pd [x o <0 G ‘0. o V o :0 o N 0 ad Q \‘r-- \\\\\V .\\\\\\\\\ * \\\\\\\\\ M\\- - k\\\ n... 883 . \ mm AouanbeJd 3% § I a: 883 q 134 L .88 25 88. 5% 25 Ed 5 32 Ba Ag m :3: “magnet. >9 @328? .853 0.8 HBO .AMDB @5850 8: m 98 0M3? m dosmbnoonoo m 80% .Amoafludmv no.8 8555 82 646 £33 eme— 58 52:8 m 83 5E» Ba 5 3882 db 5? 5a owflfi 23665372 5 25mm 5.:va 50(00 moOF0< mmem r06. RLD-LP2005 flow N65 0.3 96¢ NNN 82% /1\ P. : «9200 \1/ mm m w100 ug/g and 21 to 54pg/g for seed Zn concentration (Beebe et al., 2000). Increased mineral content in the diet does not guarantee increased micronutrient status for the consumer. There are numerous compounds in plant- based diets that reduce micronutrient absorption by the body. Among these, phytate is the major determinate of Fe and Zn bioavailability in the diet (Lopez et al., 2002). Phytate (myo-inositol 1,2,3,4,5,6 hexakisphosphate) is the major form 138 of phosphorus in cereal and legume seeds (Raboy, 1990). Phytic acid has 6 ‘ phosphate groups that can covalently bind to cations, especially K+, Ca2+, Mg“, Fe”, and Zn2+ (Tsao, 1997). Under human physiological conditions, some phytate complexes are insoluble, thus decreasing the bioavailability of phosphorus and cations, especially iron and zinc, stored as a part of the complex (Schlemmer et al., 1995). Numerous studies, using a variety of experimental techniques, have shown a correlation between high phytate diets and limited Fe and Zn absorbance in the gastrointestinal tract of humans and animals (Saha et al., 1994; House et a1, 1982; Tumlund et al., 1984; Hunt et al., 1998; Zhou et al., 1992). ‘ On one hand, P stored in legume seeds (in the form of phytic acid) can reduce micronutrient absorption by humans, but on the other hand, P is an element essential to plant growth that is often limited in the environment. In fact, P deficiency is one of the most widespread abiotic stresses affecting common bean production (W ortmann et al., 1998). Soil P deficiency often occurs in the same regions of the world plagued by micronutrient deficiencies in humans. Screening of common bean germplasm for ability to grown in low soil P conditions has resulted in the discovery of bean lines tolerant to low P soils (Lynch and Beebe, 1995; Yan etal., 1995 a,b; Beebe et al., 1997). QTL studies conducted with P efficient lines resulted in the identification of regions of the bean genome that play a role in tolerance to low P soils, and plant phenotypes associated with those regions (Beebe et al., 2006; Ochoa et al., 2006; Yan et al., 2004; Liao et al., 2004). 139 Since biofortification of common bean for Fe and Zn is underway and phosphorus plays such an important role in plant and human nutrition, an increased understanding of the genetics of tolerance to low P soils together with the genetics of seed P, Fe, and Zn concentration will be beneficial in the development of stress resistant and micronutrient fortified crops. The objectives of this research were first, to measure seed P, Fe, Zn, and phytic acid levels in a population of recombinant inbred lines developed from common bean line G19833 tolerance to P soil and AND696, a line susceptible to low P soil. The second objective was to examine the relationship among these nutrients in the seed and tolerance to low P soils. The third objective was to identify QTL'for seed nutrient levels in the RIL population. Materials and Methods: Seventy seven F57 recombinant inbred lines were developed from a cross between two common bean lines from the Andean gene pool. The initial cross 619833 x AND696 was advanced to the F59 generation by single seed descent and seed of the RILs was increased for field studies. The parent 619833 is a Peruvian landrace with an indeterminate (Type III) growth habit. G19833 has large yellow and red mottled seed with an average 100 seed weight of 41 g. G19833 has also been identified as tolerant to low P soils (Y an et al., 1995a; Yan et al., 1995b). The parent AND696 is a CIAT improved line fiom the race Nueva Granada. It has a determinate growth habit (Type I) and has large red and white mottled seed with an average 100 seed weight of 51g. AND696 has been identified as susceptible to low phosphorus soils (CIAT, 2000). 140 In 2000 and 2005, the 77 RILS, G19833, AND696, and two check varieties, G4017 (Carioca) and G16140 were planted in Darien, Colombia in a 9x9 lattice design with three replications at two soil P levels, low P (45 kg/ha triple super phosphate) and high P (300 kg/ha triple superphosphate). Phenotypic data was collected on 75 of the 77 RILS. The soil of this site in an Andisol with a native soil P of 2 mg/kg based on bray II extraction method. Seed was hand- planted at 10 cm spacing in 4 row plots. Each row was 3 m in length. The middle two rows were planted to the genotypes of interest and the outer two rows were border rows of Mesoamerican cultivar BAT 477. Seeds reached maturity at 76 to 98 days after planting. At maturity, seeds were hand-harvested. Seeds were dried and seed yield was determined at 18% moisture. Weight of 100 seed was measured at 18% moisture. A sub sample of seed was analyzed for P, Fe, and Zn concentration. Five grams of seed of each treatment and replication were cleaned with distilled water and dried. They were then freeze dried to remove all moisture. Freeze-dried samples were ground to a fine powder. Total seed P analysis was conducted according to Murphy and Riley (1962). Iron and Zn were extracted from seed with a Nitric-percloric (2:1) digestion according to the method of Benton and Jones (1989). Atomic absorption spectrometry was subsequently used to measure Fe and Zn concentration of the seed samples. Phytic acid was measured in seed from the 2005 season by high performance liquid chromatography (HPLC) using a modified version of Graf and Dintis (1982). Sample preparation for PA analysis in bean seed was conducted 141 following the procedure of Lehrfeld (1989). PA was extracted from the samples by the addition of 20ml 0.65M HCl (trace element grade) to 0.500 g of sample. The acidified samples were mechanically agitated at 250rpm for 2 hours at 21°C. Samples were centrifuged at 4,000 rpm for 15 minutes. The supemant was collected and diluted 1:5 (v/v) with HPLC grade water. A 3 m1 aliquot of the diluted sample (15 m1 total) was passed through a Bond Elut strong anion exchange column (V arian, Walnut Creek, CA) for purification. The column was washed once with 5 ml 0.14 M HCl, followed by a 10 ml wash with 0.5M HCL. Phytic acid was eluted from the column with 3m12M HCl. The eluted fraction containing PA was freeze-dried and dissolved in SmM sodium acetate. The dissolved sample was filtered through a 2pm filter. PA levels in each sample were quantified by HPLC with a refractive index detector. The column used for analysis was a Waters Symmetry C18 column (3.9mm x 150mm) (Waters, Milford, MA) heated to a temperature of 40°C. Sodium acetate (SmM) was used as the solvent at a flow rate of 1.4 ml min'l. Phytic acid dodecasodium salt from corn (Z. mays L.) (Sigma, St. Louis, MO) was used as a standard to determine PA concentration. Data was converted to concentration on a dry weight basis. Statistical analysis was conducted with SAS for Windows V8 (SAS Institute, Cary, NC). The command PROC GLM was used to determine treatment, genotype and interaction effects. Data was analyzed as a split plot with soil P level as the whole plot and plant genotype as the Split plot. The command PROC CORR was used to determine Pearson correlation coefficients among variables. Effects of plant growth habit on measured variables were determinedby means 142 comparisons of determinate vs. indeterminate growth habit types and differences were established using Tukey’s test for significance. Bartlett test for variance homogeneity was conducted for growth habit because there were unequal number of determinate (56) and indeterminate (21) RILS. A linkage map of the G19833/AND696 was developed using JoinMap J oinMap 3.0 for Windows (Van Ooijen and Voorrips, 2002) as described in the Materials and Methods section of Chapter 4 in this thesis. QTL analysis was conducted with the genetic map developed with the J oinMap program and with the phenotypic means for each RIL collected from the field study. The computer software program Windows QTL cartographer version 2.5 (Wang et al., 2006) was used to identify QTL for seed weight, seed P, Fe, Zn, and phytic acid concentrations. The Composite Interval Mapping (CIM) feature of Windows QTL cartographer employing the forward and backward regression model set tolOcM window width was used to identify QTL. CIM analysis is based on maximum likelihood estimates and linear regression to identify QTL within marker intervals. Significant QTL were considered by defining the LOD score at p=0.01 after 1000 permutation tests. Results: Variability: Seed weight: Seed weight was influenced by genotype and environment in the RIL population in both 2000 and 2005. There were no significant genotype x environment interactions (GxE) (Table 1). The mean 100 seed weight of the RILS was 2-3 g higher in the high P environment than the low P environment in 143 each year (Table 2). The ranges for seed weight overlapped between the environments, in each year (Figure 1). AND696 had larger seeds than 619833 in the high P environment (51.4 and 41.4 g per 100 seed, respectively) (Figure 1). Seed P and phytic acid: Genotype and environment had a significant effect on seed P levels of the RILS in 2000 and 2005. In each year there was a significant GxE for this trait (Table 1). Concentration of seed P was higher in the high P than the low P environment each year (Table 2). The frequency distributions of seed P showed more variability for this trait in the high P environment than the low P environment (Figure 1). 619833 had higher seed P concentration in the high P environment than AND696 (Figure 1). Percent phytic acid, which was only measured in 2005, was significantly affected by soil P environment (Table 1). There was only a genotype effect for this trait observed in the low P environment (Table 2). In the high P soil environment, AND696 was at the low extreme of the RILS, at 1.46 %, whereas the range for the RILS ranged from 1.33 to 2.85 % (Figure 2). Seed Fe and Zn: The concentrations of Fe and Zn in the seeds were significantly affected by genotype and environment. There was no significant 6 x E for either of these traits (Table 1). Iron levels were higher in the high P environment, whereas Zn levels were higher in the low P environment (Table 2). The inhibition of Zn uptake by high levels of soil P has been observed previously in common bean (Singh et al., 1988) and the phenomena is known as P induced Zn deficiency. There was greater variability for Fe levels in the population that 144 for Zn levels (Figure 3). The range of Fe levels was larger in 2005 than in 2000 (Figure 3). Correlations: Correlations with seed weight and the other seed traits were not consistent across years. In 2000, seed weight was positively correlated with seed Fe and Zn in both high and low P treatments (Table 4). In 2005, seed weight . was only positively correlated with seed Zn in the high P treatment (Table 3). In each year and each P environment, seed P was positively correlated with seed Fe and Zn concentrations (Tables 3 and 4). Phytic acid % was also correlated with seed Fe and Zn in high and low P environments, but not as strongly as total seed P (Table 4). In 2005, seed Fe and Zn concentrations were weakly and significantly correlated with P uptake in the low P treatment, I = 0.17 and 0.12 respectively. In the high P treatment, only seed Zn was correlated with P uptake (r = 0.32). Growth Habit: Seed weight, P, Fe, Zn, and phytic acid were each influenced by plant growth habit (Table 5). The indeterminate RILS had larger seed weights (49.6 g in high P and 46.2 g in low P) on average than the determinate RILs (46.8 g in high P and 43.7 g in low P) and also had greater concentrations of P, Fe, and Zn. There was an exception to this trend observed in 2000, under low P conditions where the determinate lines had higher seed P concentrations (Table 5). The differences in seed P concentration among years in may relate to differences in seed yield by growth habit in 2000 and 2005. In 2000 the average yield of the indeterminate lines was 707 kg-ha'1 under low P conditions and that of the determinate lines was 614 kg-ha'l. In 2005, there were 145 no differences in yield in the low P treatment based on growth habit. (data shown in Chapter 3). Since there were unequal number of determinate and indeterminate RILs in the population, Bartlett’s test for variance homogeneity was conducted to identify bias that may have existed in trait analysis by growth habit. Most traits exhibited variance homogeneity although there were exceptions, including seed weight (high P, 2005), seed phytic acid (low P, 2005) and seed Zn (low P, 2005) (Appendix, Tables A4-4). QTL identification: Using composite interval mapping, 32 QTL were identified for 5 seed traits in 15 marker intervals on 8 linkage groups. Twenty-eight percent of the QTL were clustered on linkage groups B1 and affected more than one trait. Tweleve percent of the QTL were clustered on linkage groups B3. An additional 31% of the QTL were located on linkage group B6 and also affected multiple traits. Individual QTL explained 10 to 35% of the phenotypic variation, and total phenotypic variation explained for any one trait was 10 to 61% (Table 1). Seed weight: A QTL for seed size was identified on linkage group B6 in 8 each year and environment where the increased effect was contributed by 619833 (Table 6, Figure 4). This QTL explained 21 to 27% of the phenotypic variability in 2005, depending on environment and only 13 to 15% of the variability in 2000. An additional QTL for seed size was found in 2005 on B3 where the increased effect was contributed by AND696 (Table 6, Figure 4). QTL for seed size have previously been identified on B3 and B6 in a population developed from bean lines PC50 and XAN159 segregating for white mold resistance (Park et al., 2001). 146 A QTL for seed size, in the same region of B3 as that identified by Park et a1. (2001) has also been found in the BAT 93/Jalo EEP558 population (BIC, 2006) Seed P and phytic acid: QTL for seed P were found on B1 near the gene for plant grth habit. QTL found for seed P concentration on Bl explained 17 to 28% of the phenotypic variation for this trait depending on year and soil P level (Table 6, Figure 4). In 2000 and 2005 in the high P environment the increased effect was contributed by 619833, whereas in low P 2000 the increased effect was contributed by AND696. A QTL for phytic acid on B8 mapped to same region as a QTL for total seed P in high P, 2005, where the increased effect was contributed by AND696 (Table 6, Figure 4). The QTL for % phytic acid explained only 10% of the phenotypic variation and the QTL for total seed P explained 20% of the phenotypic variation. . Seed Fe and Zn: QTL for both seed Fe and Zn were identified in the identical regions of B1 and B6 (Table 6) across years and environments. QTL found on B1 were near the fin gene, and also overlap with QTL for seed P concentration. 619833 was the source of the QTL identified for Fe and Zn concentration in the seed on B1. Those QTL found for seed Fe and Zn concentration on linkage group B6 explained 15 to 26% of the phenotypic variation. They also co-localized with QTL for seed weight. AND696 was the source of the increased effect of each of these QTL, conversely, 619833 was the source of the QTL for seed size identified in this region of linkage group B6 (Table 6). 147 Discussion: Increasing the Fe and Zn concentration of seed crops is a major goal that bridges agriculture and human nutrition. In common bean, significant natural variability exists for seed Fe and Zn, making such a goal achievable through plant breeding (Beebe et al., 2000). Within the AND696/619833 RIL population, Fe levels ranged from 38 to 79 mg kg1 and Zn levels ranged from 16 to 33 mg kg". Understanding the genotype x environment interactions of these traits is essential to developing nutritious seeds for different environmental conditions. In this study, 6 x E was not significant for Fe or Zn seed concentration, suggesting that breeding for increased levels of these elements is possible under diverse soil P environments. - Mutational studies in plants have shown that many genes effect micronutrient accumulation and many remain to be discovered. QTL analysis will allow gene discovery to move forward for genes underlying natural variation in seed Fe and Zn levels (Ghandilyan et al., 2006). Overlapping QTL for Fe and Zn concentration were identified on two linkage groups in this study. This _ demonstrates the possibility of breeding for these traits simultaneously. QTL for increased seed Fe and Zn concentration were located on the same region as a QTL for seed P concentration. These QTL were all contributed by ‘ 619833, which was the parent exhibiting tolerance to low P soils. This region of the genome (linkage group B1) may contain genes that improve tolerance to low P soils, as well, according to yield QTL data (Chapter 4, Table 1). Further research is required to determine the consequences of simultaneously selecting 148 increased seed Fe, Zn, and P levels on the bioavailability of these micronutrients in humans upon consumption. In this study, only a single QTL for seed phytic acid was detected (Table 6). This QTL was not in the same region as QTL for seed Fe or Zn concentrations. In Arabidopsis, QTL for seed phytic acid levels have been separate from those for seed Fe and Zn levels suggesting it is possible to increase Fe and Zn levels without increasing phytate levels as well (Bentsink et al., 2003; Ghandilyan et al., 2006) Since many of the QTL detected in this study for each of the traits co- localized it is interesting to consider the physiological role of the genes underlying these QTL. Could these genes be involved in seed transport or in uptake of nutrients by the plant roots? Iron uptake by pea (Pisum sativum) has been shown to increase during seed filling as compared to earlier in plant development. During this time, a constant source of micronutrients in required in the xylem stream for xylem-'to-phloem exchange to occur, and for micronutrients to arrive in the seed (Grusak et al., 1999). QTL for seed Fe, Zn, and P on B1 co-localized with the fin gene (Figure 4). This is the single gene in bean responsible for deterrninacy (Norton, 1915). The QTL for increased levels of each of these elements was conferred by the indeterminate parent (61983 3). That indeterminate plants generally continue root growth beyond flowering whereas determinate plants do not (Huyghe, 1998), which may explain why this region of the genome influenced these traits for increased nutrients in the seed. 149 Conclusions: Significant variability for seed weight, P, Fe and Zn concentrations was present in the AND696/619833 population. Seed P concentration was correlated with seed Fe and Zn concentrations. QTL for these traits were found to co localize on linkage group B1. Seed Fe and Zn concentrations were also correlated and QTL for these traits co-localized not only to linkage group Bl, but to B6 as well. The absence of GxE interactions for seed Fe and Zn levels is promising in terms of future development of lines with increased levels of these nutrients when planted in diverse environments. 150 References: Beebe S, Gonzalez A, and J. Rengifo 2000. Research on trace minerals in the common bean. Food Nutr. Bull. 21: 387-391 Beebe, S., J. Lynch, N. Galwey, J. Tohme, and I. Ochoa. 1997. A geographical approach to identify phosphorus-efficient genotypes among landraces and wild ancestors of common bean. Euphytica 95:325-336. Beebe, S.E., M. Rojas-Pierce, X.L. Yan, M.W. Blair, F. Pedraza, F. Munoz, J. Tohme, and J .P. Lynch. 2006. Quantitative trait loci for root architecture traits correlated with phosphorus acquisition in common bean. Crop Science 46:413- 423. Bean Improvement Cooperative (BIC) 2006. Bean Core Map 2006 [Online] http://www.css.msu.edu/bic/Genetics.cfin (posted 2006; verified November 14). Benton, J .B., and J. Jones. 1989. Plant analysis techniques Benton-J ones Laboratories, Georgia. Bentsink, L., K. Yuan, M. Koornneef, and D. Vreugdenhil. 2003. The genetics of phytate and phosphate accumulation in seeds and leaves of Arabidopsis thaliana, using natural variation. Theoretical and Applied Genetics 106: 1234-1243. Bouis, HE. 2003. Micronutrient fortification of plants through plant breeding: can it improve nutrition in man at low cost? Proceedings of the Nutrition Society 62:403-41 1. CIAT. 2000. Inheritance of low phosphorus tolerance in the Andean population AND696 x 619833. Bean Improvement for the Tropics Unit Annual Report. CIAT, Cali, Colombia. Ghandilyan, A., D. Vreugdenhil, and M.G.M. Aarts. 2006. Progress in the genetic understanding of plant iron and zinc nutrition. Physiologia Plantarum 126:407- 417. ' Graf, E., and FR. Dintzis. 1982. Determination of Phytic Acid in Foods by High- Perforrnance Liquid-Chromatography. Journal of Agricultural and Food Chemistry 30: 1094-1097. Grusak, M.A., J .N. Pearson, and E. Marentes. 1999. The physiology of micronutrient homeostasis in field crops. Field Crops Research 60:41-56. Harvest Plus 2004 Bioforified Beans Available: http://www.harvestplus.org/pubs.html 151 House, W.A., R.M. Welch, and DR. Vancampen. 1982. Effect of Phytic Acid on the Absorption, Distribution, and Endogenous Excretion of Zinc in Rats. Journal of Nutrition 112:941-953. Hunt, J .R., L.A. Matthys, and L.K. Johnson. 1998. Zinc absorption, mineral balance, and blood lipids in women consuming controlled lactoovovegetarian and omnivorous diets for 8 wk. American Journal of Clinical Nutrition 67:421-430. Huyghe, C. 1998. Genetics and genetic modifications of plant architecture in grain legumes: a review. Agronomic 18:383-41 1. International Zinc Association 2000. Zinc and Human Health. International Zinc Association. Brussels, Belgium. Koinange, E.M.K., S.P. Singh, and P. Gepts. 1996. Genetic control of the domestication syndrome in common bean. Crop Science 36: 1037-1045. Lehrfeld, J. 1989. High-Performance Liquid-Chromatography Analysis of Phytic Acid on a Ph-Stable, Macroporous Polymer Column. Cereal Chemistry 66:510- 5 1'5. Liao, H., X.L. Yam, 6. Rubio, S.E. Beebe, M.W. Blair, and J.P. Lynch. 2004. Genetic mapping of basal root gravitropism and phosphorus acquisition efficiency in common bean. Functional Plant Biology 31 :959-970. Lynch, J .P., and SE. Beebe. 1995. Adaptation of Beans (Phaseolus vulgaris L) to Low Phosphorus Availability. Hortscience 30:1165-1171. Murphy, J ., and J .P. Riley. 1962. A modified single solution for determination of phosphate in natural waters. Anal Chem Acta 27:31-36. Norton, J .B. 1915. Inheritance of habit in the common bean. Am. Nat. 49:547-561. Ochoa, I.E., M.W. Blair, and J .P. Lynch. 2006. QTL analysis of adventitious root formation in common bean under contrasting phosphorus availability. Crop Science 46:1609-1621. Park, S.O., D.P. Coyne, J .R. Steadman, P.W. Skroch, 2001. Mapping of QTL for resistance to white mold diseases in common bean. Crop Sci41: 1253-1262. Raboy V. 1990. The biochemistry and genetics of phytic acid synthesis. In: Morre DJ, Boss W, Loewus FA (eds) Inositol metabolism in Plants. Alan R. Liss, New York, pp 52-73. 152 Saha, P.R., C.M. Weaver, and AC. Mason. 1994. Mineral Bioavailability in Rats fi'om Intrinsically Labeled Whole Wheat-Flour of Various Phytate Levels. Journal of Agricultural and Food Chemistry 42:2531-2535. Schlemmer, U., K. Muller, and K.D. J any. 1995. The Degradation of Phytic Acid in Legumes Prepared by Different Methods. European Journal of Clinical Nutrition 49:S207-S210. Tsao, G.T., Y.Z. Zheng, J. Lu, and CS. Gong. 1997. Adsorption of heavy metal ions by immobilized phytic acid. Applied Biochemistry and Biotechnology 63- 51731-741. Tumlund, J.R., J.C. King, W.R. Keyes, B. Gong, and MC. Michel. 1984. A Stable Isotope Study of Zinc-Absorption in Young Men - Effects of Phytate and Alpha-Cellulose. American Journal of Clinical Nutrition 40:1071-1077. Plant Research International BV. 2002. JoinMap. Release 3.0. Plant Research International BV. Singh, J. P., Kararnanos, RE, and J .W.B. Stewart, 1988. The Mechanism of Phosphorus-Induced Zinc Deficiency in Bean. Canadian Journal of Soil Science. 68: 345-358. Statistical Genetics, North Carolina State University, USA. 2006. Windows QTL Cartographer. Release 2.5. Statistical Genetics, North Carolina State University, USA. Welch, RM. 2002. The impact of mineral nutrients in food crops on global human health. Plant and Soil 247:83-90. Wortmann, C.S., R.A. Kirkby, C.A. Eledu, and DJ. Allen. 1998. Atlas of common bean(Phaseolus vulgaris L.) production in Africa CIAT, Cali, Colombia. Yan, X.L., J .P. Lynch, and SE. Beebe. 1995. Genetic-Variation for Phosphorus Efficiency of Common Bean in Contrasting Soil Types .1 .V Vegetative Response. Crop Science 35: 1086-1093. Yan, X.L., S.E. Beebe, and J .P. Lynch. 1995. Genetic-Variation for Phosphorus Efficiency of Cormnon Bean in Contrasting Soil Types .2. Yield Response. Crop Science 35: 1094-1099. Yan, X.L., H. Liao, S.E. Beebe, M.W. Blair, and J .P. Lynch. 2004. QTL mapping of root hair and acid exudation traits and their relationship to phosphorus uptake in common bean. Plant and Soil 265:17-29. 153 Zhou, J .R., E.J. Fordyce, V. Raboy, D.B. Dickinson, M.S. Wong, R.A. Burns, an J .W. Erdman. 1992. Reduction of Phytic Acid in Soybean Products lrnproves Zinc Bioavailability in Rats. Journal of Nutrition 122:2466-2473. 154 Table 1: Analysis of variance of seed traits in a population of 75 recombinant inbred lines from the AND696/619833 population field grown in Darien, Colombia in 2000 and 2005 in two environments: high and low soil phosphorus. Mean Squares Source df Seed Seed P Seed Seed Seed weight Fe Zn phytic acid 2005 Genotype (G) 74 127‘" 0.29"" 205’" 26“" 0.11 "5 Environment (E) 1 1094"" 95.13‘" 3404*" 943‘" 15.49"” G x E 74 7.3 “S 0.10‘ 36 "S 7 "S 0.10 “5 2000 Genotype (G) 74 104"" 0.26"” 133‘" 20*" ----- Environment (E) 1 540*" 72.07"" 459‘" 862"" ----- G x E 74 13 "S 0.18"” 43 "5 6 “5 ----- "3 indicates not significant * Indicates significance at P < 0.05. **Indicates significance at P <0.01. *** Indicates significance at P<0.001 155 Table 2: Seed traits in high (HP) and low (LP) phosphorus soils conditions for parents AND696 and 619833 and the means and ranges of 75 recombinant inbred lines developed from the parents. The experiment was planted in 2000 and 2005 in Darien, Colombia. Mean values presented (n=3). P value indicates level of significant genotypic differences among the RILS for each trait. Parents Recombinant Inbred Lines Traits P AND696 619833 mean range P value level 2005 Seed weight HP 51.4 41.4 47.6 39.5-59.9 <.0001 (g .100 seed") LP 44.7 ------- 44.5 33.3-56.5 <.0001 Seed P HP 3.84 4.00 4.05 3.56-4.85 <.0001 (g kg") LP 3.28 ------- 3.14 2.63-3.51 <.0001 Seed Fe HP 62.5 61.6 58.3 39.02-78.75 <.0001 (mg -kg") LP 48.6 ------- 52.8 37.92-67.66 <.0001 Seed Zn HP 20.2 21.0 22.7 18.74-30.35 0.1550 (mg kg") LP 26.3 ------- 25.6 20.18-32.83 <.0001 Seed phytic HP 1.46 ------- 1.77 1.33-2.85 0.2214 acid (%) LP 1.26 ------- 1.38 1.09-2.46 0.0272 2000 Seed weight HP ------ 34 44.94 35.5-56.75 <.0001 (g -100 seed") LP ------------ 42.78 30555 <.0001 Seed P HP ------ 4.77 3.54 2.84-4.98 <.0001 (g -kg") LP ------------ 2.70 2.36-3.04 <.0001 Seed Fe HP ------ 78.45 64.97 53.3-76.9 <.0001 (mg -kg") LP ------------ 63.01 53.1-75.6 <.0001 Seed Zn HP ------ 25.43 20.60 16.3-27.4 <.0001 (mg -kg'1) LP ------------ 23.59 18.9-29 <.0001 156 Table 3. Phenotypic correlations among seed traits in a population of 75 recombinant inbred lines from the AND696/G19833 population field grown in Darien in 2005 in high and low soil phosphorus. The values above the diagonal line are under low soil P and the values below the diagonal line are under high soil P. LP Seed P Seed Seed Seed Fe Seed Zn ------------ phytic weight HP acid Seed P ---------- 0.35'" -0.05 0.53'" 0.50'" Seed phytic 0.27‘" ---------- 0.07 0.16" 0.26"" acid Seed weight 0.04 -0.09 ----------- 0.10 0.10 Seed Fe 031‘" 0.16" 0.00 ........... 053"” Seed Zn 0.48"" 0.10 0.23"" 0.31‘” ........... * Indicates significance at P < 0.05. **Indicates significance at P <0.01. *** Indicates significance at P<0.001 157 Table 4. Phenotypic correlations among seed traits in a population of 75 recombinant inbred lines developed fromAND696/Gl9833 and field grown in Darien in 2000 in high and low soil phosphorus). The values above the diagonal line are under low soil P and the values below the diagonal line are under high soil P. LP Seed P Seed Seed Fe Seed Zn ------------ weight HP Seed P ---------- -0.13' 0.18" 0.28"" Seed weight 026'" ----------- 0,14" 0,14" Seed Fe 0.39"" 0.15" ........... 0,17‘" Seed Zn 0.65‘" 0.26‘" 0.45‘" ........... * Indicates significance at P < 0.05. **Indicates significance at P <0.01. *** Indicates significance at P<0.001 158 Table 5. Seed traits in a population of 75 recombinant inbred lines developed from an AND696/G19833 cross and grown under high and low soil phosphorus conditions in Darien, Colombia in 2000 and 2005, grouped and averaged by plant growth habit of which there were two categories, indeterminate (1nd.) and determinate (Det.). Mean values presented (n=3). Traits High P Low P Ind. Det. Ind. Bet. 2005 Seed weight 49.6 a)‘ 46.8 b 46.2 a 43.7 b (g 100 seed") Seed P 4.23 a 3.97 b 3.2 a 3.1 b (g 'kg") Seed Fe 63.3 a 56.1 b 55.7 a 51.4 b (mg 158") Seed Zn 24.22 a 21.96 b 27.4 a 24.7 b (mg 'kg") Seed phytic 1.78 a 1.76 a 1.46 a 1.34 b acid (°/o) 2000 Seed weight 46.8 a 44.2 b 43.6 a 42.4 a (g .100 seed'l) Seed P 3.7 a 3.5 b 2.6 a 2.8 b (g 458") Seed Fe 67.4 a 64.0 b 63.4 a 62.8 a (mg 158") ‘ Seed Zn 22.3 a 19.9 b 24.4 a 23.2 b (mg kg“) 1' Significant differences are based on Tukey tests and are at alpha = 0.05. Values that do not share the same letter are significantly different. Tests should be read across rows and within each phosphorus treatment level, High P or Low P. 159 Table 6. Putative QTL for seed traits identified from 75 recombinant inbred lines developed from an AND696/619833 cross grown under high (HP) and low (LP) soil phosphorus conditions in Darien, Colombia in 2000 and 2005. Seed traits Linkage Nearest LOD if AdditivityF goup marker score? CIMI Seed wt. LP 2005 B3 AGTC03 5.6 0.18 2.1 B6 GCTC03 6.7 0.21 -2.4 HP 2005 B3 AGTC03 8.4 0.31 2.7 B6 BM170 7.3 0.27 -2.6 LP 2000 B3 AGTC03 2.7* 0.16 1.9 B6 GCTC03 3.8 0.15 -1 .9 HP 2000 B2 BM164 3.9 0.35 -2.8 B6 GCTC03 3.8 0.13 -1.8 Seed P LP 2005 B3 AGTC03 4.3 0.22 0.09 HP 2005 B1 AGTAOI 4.3 0.17 -0.14 B4 CGTCOS 3.6 0.18 0.13 B8 16.550A 4.3 0.20 0.14 LP 2000 B1 AGTAOI 4.3 0.17 0.08 HP 2000 B1 fin 6.9 0.28 -0.22 B2 BM143 3.1 0.10 -O.13 A ACTC06 3.6 0.14 -0.15 Phytic acid LP 2005 B8 . 16.550A 2.3* 0.10 0.06 Seed Fe LP 2005 B1 BMle 5.4 ' 0.18 -2.4 B6 BM170 6.9 0.26 2.7 HP 2005 B1 AGTAOI 8.1 0.24 -3.9 B6 BM170 5.3 0.15 2.8 B9 ATA9 5.1 0.14 2.8 A GGATOZ 3.3 0.08 22 LP 2000 BS M12.1600A 4.0 0.15 1.86 HP 2000 B1 BM165 2.6* 0.10 -2.1 B6 BM170 4.5 0.22 3.2 Seed Zn LP 2005 B1 fin 7.5 0.18 -l .2 B6 BM170 8.2 0.24 1.3 HP 2005 B1 fin 6.8 0.31 -1.3 LP 2000 B1 BMle 28* 0.10 -0.7 B6 BM170 4.1 0.16 0.9 HP 2000 Bl fin 7.8 0.35 -1.5 160 Table 6 (cont’d). T LOD: Log of odds 1 Proportion of the phenotypic variance explained by QTL at test site using CIM (composite interval mapping). F Effects of substituting a single allele from one parent to another. Positive values indicate that allelic contribution is from AND696 and negative from G19833. . Indicates that the LOD score fell below the cutoff range of 1000 permutations at p = 0.05. 161 doom mom one 5m: 65 no 805 was meow 8m 98 a3 65 no 3&5 .maougnmfl m «o 09985 Ba Ba mos—3w, .Aooom a :3on «on an? 80% Ba moon 3 85on “on ma? «.meva 38:58.: wannabe BEE mogoonow 3:38 no menu? :38 “concave 6382 .222 636888 :8 Ed 32 28 86 SE Home. 88 ea 88 a 638260 88o a gem «366689.54 nonmfipaom 05 Bob Am: won: “.053 “mafia—noon: mu 3 m Boom was 3395 voom .8.“ mnouaflbmmu zonoswoum A .mwm b- 38 0o? wv Emmoawoom 8 8 mm . , 8 me 3 mm . om 8 mm Om mw ow mm b— vv Vesvv\\\‘\‘\.\\ o o :“ieeeew: “Scenery .. e \\ \\\\\\ \\\\\\\\\ . \\\\\\\\\ \\\\\\\\\ \\.\\ \\\\ \ \\\\\\\ ..v v m “mm““mmnxmmmmeo. “was“ . _ \.\\\\ i \\\\ o o a“: “m““ 61385 ““8“ S e :m .e e, . 61885. a“ m“ .2 . 2 “e \ \ m“ . m -e 6.. 882.4 a \K Y nil d._§\\\n -2 . z . 3888 Emu ooom 81m )0 Aouenbeid 162 9m 11 989.0 m6 06 md 0.” Wm 9. me. my m 86m ON 9. NF 3. m: 9. ON o6 m6 m1 mmmmww .QI mach—24‘ of ad _ n3. omen—24¢ A 9m md .383 2 ON 3 NF 3. 9. 3 ON ma 81mm Aouenbeig 163 .mnoamoanou m mo owns? 65 can 835» .Amoom E 850% “on we? mmwofiwv flfigmob Enact—o .893 mambonom 332mm mo mos?» SEE 68662 $65. .865 £8888 :8 So 36. an 83 fie see: 38 a 63828 85a a 52» 5298372 couflzmoq ofi 89a Gina won: 333 “afloEooB mm E Eon cumin ween .8.“ mnousfibmmo >325on .N .wmm. g .266 ease 8% Nu . oN a; m... V... N4 oé “a... woooz< S ‘cocovmo .\\\\\\\\\\\\\\\\\ NP 3. 91m )0 Aouanbeid or O.———> .— moooz< or 61' 64$ Emu moow on. 164 1G19833 HP E 0- a” (u ...I I 0 § . L\\\ g \I . .m 8 § , \\\\\\\\ n\\\\\\\\\\\ t\“\\\‘\‘\““‘ .\\\\\\\\\\\\\\w .\\\\\\\\\\\\\\\\\\i l\\‘\\\\‘\\\\\\\\\\‘\‘\‘\\\\‘\“\‘\“‘\““ L‘\\\‘\\\\\\\\‘\\\\\\\\\\\\\\‘\‘\\\‘\\‘\“\“\‘\\‘ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\v s\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ t\\\\\\\\\\\\\\\\\\\\\\‘ \\\\\\\\ . h\“\\\‘ L\\\‘\\\ m. b I I I I I I I r V N 0 w (D V N C 1- 1- ‘- 3 (B '0 IO 8 CL (0 m 8 to 2 5 o n\\ —' \\\\\\\\\\\\\\\‘ ———. n\\\\\\\\\\ u\\\\\\\\\\\\\\\\\\\\\\‘ s\\\\\\\\\\\\\\\\\\‘ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ n\\\\\\\\\\\\\\\\\\\\\\‘ A‘\\\\\‘\“\\\\“‘\‘\\“\\“‘\\‘\‘\“\““‘ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\~ 0- L\\\\\\\\\\\\\\‘ _I ——. s\\\\\\\\\\\\\\\\\\‘ 8 u\\\\\\\\\\\\\\\\\\\\\\\\\‘ (D D n\\\\\\\ . E L\‘\ n\\\ I I I I I ' I ' I ' 1 V N O Q ‘0 V N O F 1- mm )0 Aouanbeid '165 8 In N o N «‘3 8 '75 70 ID 0 8 B 8 9 3 Seed Fe (mg-kg‘l) Fig. 3. Frequency distributions for seed Fe and Zn in 75 recombinant inbred lines (RILS) from the population AND696/G19833 grown in Darien Colombia in 2000 and 2005 under high (HP) and 10w (LP) soil phosphorus levels. Arrows represent mean values of parental genotypes under different treatments (G19833 was not grown in 2005 and AND696 was not grown in 2000) Values are the average of 3 rephcations Graphs on the left are for 2005 and those on the right are for 2000 Fig. 3 (cont’d). G19833 HP 28 k\\\\\\\\ b k\\\\\\\\\\\\\\\\\V —-—> k\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ 26 N\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ 24 k\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ ;\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\V 22 k\\\\\\\\\\\\\\V k\\\\\\V ‘20 .\\‘ 18 16 32 30 28 AN0696 LP 24 1 G19833 HP 1 18 AND696 HP I 16' 20 18 4 2 0 CD? N000 P 91m )0 Aouanbeig 166 ‘ Seed Zn (mg-kg'l) .88 Ba 88 .48» Ba 66 m as 28 SE m 8.3 .6688: E 8882 865 on 4.8 .AEeeeeo Emcee 88 ea .28 88 65 a 28 23% 688 .88888 6N es... .8...“ .m $8 é 8882 H8 55 one emfié 23666872 .e 883 3:5 to: 5.0.20 . 32 F.oo<“. 1 9% 6 Em \1 / mmm 38.82 . no» 5320 new 85.00 to 8050 to m 325 m.» mu. m. .8633 8:5 6.. ., . «58.8: o.» 82n>d ~.~ 8Chi- level Square P uptake HP 1.06 0.3032 LP 0.49 0.4851 P use efficiency HP 0.01 0.9206 LP 0.36 0.5468 Days to HP 10.35 0.0013* maturity 2005 LP 0.03 0.8631 ' Days to HP 1.58 0.2081 maturity 2000 LP 0.06 0.8094 Seed P content HP 9.56 0.0020* ' 2005 LP 0.03 0.8561 Seed P content HP 4.98 0.0257* 2000 LP 2.76 0.0967 Seed yield HP 1.13 0.2888 2005 LP 1.07 0.3015 Seed yield HP 0.76 0.3847 2000 LP 6.07 0.0137* * Variance is not homogeneous between grth habit classes 174 Table A3-2. Bartlett test for variance homogeneity based on growth habit, where there were two classes, determinate and indeterminate, and 1 degree of freedom. Tests were conducted for roots traits measured under low P (LP) and high P (HP) treatments in the recombinant inbred line population developed from a cross between common bean lines AND696 and G19833. Traits P Chi-square Pr>Chi- Ievel Sguare Adventitious HP 0. 12 0.7266 roots LP 2.37 0.1234 Root length HP 8.62 0.0033* density LP 1.86 0.1723 Specific root HP 16.31 <0.0001* length LP 0.63 0.4259 Root surface HP 7.98 0.0047* area LP 1.37 0.2418 Average root HP 0.21 0.6431 diameter LP 0.1 1 0.735 8 * Variance is not homogeneous between growth habit classes 175 of md 905E902. £0 . 0005:0200 I n_I 9m 0000002000 m 00 0w000>0 05 000 00.002 00080000 3000 03:3 32 00058000003 90 0000300000 m0 0030., G008 00000500 $50.90. .030: £30% .3 000000000 0&0 000 a 300& 0000 0.355 .2.—mt 05 00 300% .3 Am 0:0 “0.— 08 :0 300% .3 030000500 .3 at? 30.0— m :00 .3 000000000 000 mucusabmfi 3000005 mm 00:: 009350.30. 30:0 Ewu 05 00 300% 00 00.0 $2 0.5 00 300% 06 :0 E9000 3 .5 3 32 .00 00003506 25. .200 030000030 80 32 08 EB :03 a 88 a 038250 .8000 a 8.53 0820 05 000072. 80303 so... a .80 802950 300 00:: 0033 0003080000 mm 00 350$». 00: m 20 .0030: m 000000000000 m 000% m0 00003506 mofiwcohm .72 000me A733 .0000 A 000.5 00 0.0 . . o - 0 u v - 0 u 0 I 2 1 NF I3. o.v md 905E000... § . 905E050 U n... fioi .\\\\\\\\\\\‘ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ - \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ .\\\\\‘ N \\\\\\\\\\\\\\\\\‘ Aouanbeid 176 Figure A3-l (cont’d). ‘ .9 3 m m CE 5 E 0- E O I m 3 H o '0 D E is? —. 0 ~ co . a .. o ,_ l f I ' I ' I ' l ' l 1 I ' I <0 0 N o co <0 0 N o 1- !- 1- ‘— 0 [ o .9 T6 " m .5 .E g - a g- S \‘ . D § 3 - s = — .\\\\\\\\\\\\\\\ 8 .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘ — _. . L .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\V L o ‘ ‘- l'l"l'l"l'f to v N o a co v N o - v- 1- 1- . Aouanbaig 177 P uptake (mg-plant'l) 000550.005 .. 0.055.900 I A... «.0 Nd to Awn— wfi . 0000. My. bun—2050 00: may. 19. T. INF r . h 0.0 . . . . -N . .4 7. .l@ a. a ‘ ... IS. low V 0005:5000... § _ .. 2 0005:5000 U . u on n: . .-NN 50.303 Tm< 0.5mm Muenbeid 178 . 00000000000 m .00 0w000>0 000 000 00002 8080.00 0000 00000.3 05 0000080000008 0000 00000800000 .00 002000 0008 000000000 0300.00. .0300 00.50% .3 000000000 0000 000 . 01mm 000% 0000 000003 .0030 000 00 00000w .3 0: 000 0.000 000 00 00000».~ .3 0080000000 00 0003 0.60. 0 :00 .3 000000000 000 0000008000000000000 .000 0000: 0000000000000 3000 Bat 05 00 00000w 00 000 .000 000 00. 0.000% 000 00 03000 00 0A 8 01:00 .00 0000300000 200 .08 08808.00 Ed .32 05 EB 0000 a 080 0a 080 a 038200 .5000 a 80003 00020 08 000078 0003000 00000 0 800.0 000000.000 A305 000.: 000000 0000380000 mu 8 0000000 0 0000 .00 000.000.00.00 0000000000 .N-m< 00080 0000 00.08... 000.005 050:8 0 080 8N OVN ONN 8N on _. cow 0!. ON w 00 _. omN OVN CNN .OON on P 09. 0.3. ON _. 8P - b — r 0 . 0 — u _ p _ p n b p . b o \ o d 0 c S c . I N I N .l V I Q l 0 I m l O u I Q T 2 g u 2 a m. r 00 I VF I 3. v . 055500000:— \\\\\\\x . 0005000000 I . . n... r / - 2 n_I :00 Aouanbaig 179 00050000000. 0%, 2055.200 I n:... 080 0.08... 02.000 050:8 0 080 CNN OON . our 09 _ 03. — - D D b L — L 00050000000. § 00.050.00.00 B n... § 8 R\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\V ——> ' I O P L _ -2 Aouanban , .5080 0.00. 200E . 180 00000000000 m .00 0w000>0 000 000 00002 0000000000 0000 000003 05 0000000000000 0000 00000000000 .00 00000., 00000 000000000 0300.0. .0000 00.50% .3 000000000 0000 00 01mm 0000» 0000 00000.5 .00w00 000 00 00000w .3 mm 000 000 000 00 00000w .3 00000000000 00 003 00.000 0 :00 .3 000000000 00 0000000000000 000000000 .00.“ 0000: 0000000000000 3000 Emu 000 00 300% 00 000 02 000 00 00000» 00.0 00 03000 00 A: 00 01mm .00 0000000000000 000. .08 30000800 00.0 32 08 E00 0000 a 0000 000 0000 00 038200 .8000 a 08020 000000 0.0. 000072 000.500 00000 0 000.0 000000.000 0.00% 00.00 000000 00000000000000 mm 00 200% 0000 .00 0000000000 000000000 .062 000mm 0000 0.-0... 00 0.000 080 w OOON OOO_. O09. OOE. OONw 82 com OOO OO? ..P._._..... 0000 0000 0000 003, 0000 0000 000 000 000 . _ q... M Io 0. \vav IO w. m m m .0. _... m 0 .0... = m m m m m , ... L ... .. T. .. . \ \ \ \ \ . ..,, .0 .., I “ “ “ “ “ 0 u 0 0 0 0 0 . .0 u . 0 0 0 0. m u g 0 N 0 “ 0 0 N t. ..u z ; m m m m .e m . E . M _. 0 0 0 0 ; w m 0 s ..v m x. 0 I0. 0... . _ 0 . N 0 m . I0 \ l0 0 g 0 IQ In 00050000000. a :00 00055000000. § . u 00 0005500000 I T 0005:0900 D a: 0.. IN—. IN? ' AouenbaJd 181 2500 0 H 3 m 2 IE 0 O o E 8 I 5%. ——> —-—-—> g E I ' I ' I ' I ' I 1 I ' I N O G) (D V N O I- I- O P 8 N 0 H D .03 to g .E o. E 9 - § .J .9 d) N 0 1: D 5 § . U § r 0 Q “G .\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ g I H l— : .\\\\\\\\\\\\\\\\\\\\’ 8 .\\\\\\\\\\\\\\\\\\\\\\\\\\\ V .\\\\\\\\\\\\\\\\\\\\' m —-’ 2' ' § L D a .3 ' T I I I I I I I ' I fl , I14 53 o co <0 0 N o Aouanber 182 Seed yield (kg 113'“) 2000 Table A4-1. Amplified fragment length polymorphisms (AF LP) names and selective primers used in developing a linkage map of AND696/G19833. AFLP name Selective Selective Polymorphism Jrimer 1T primer 21 number. ACAA E-AAC M-CAA 5 ACAC E-AAC M-CAC 4 ACAT E-AAC M-CAT 3 ACTA E-AAC M-CTA 2 ACTC E-AAC M-CT C 6 ACTG E—AAC M-CTG 3 ACTT E-AAC M-CTT 2 AGAC E-AAG M-CAC 1 AGAT E-AAG M-CAT 6 AGTA E-AAG M-CTA 2 AGTC E-AAG M-CTT 3 CAAA E-ACA M-CAA 3 CAAC E-ACA M-CAC 5 CAAG E-ACA M-CAG 1 CAAT E-ACA M-CAT 5 CATC E-ACA M-CTC 2 CATT E-ACA M-CTT 3 CCAA' E-ACC ' M-CAA 3 CCAC E-ACC M-CAC 1 CCTA E-ACC M-CTA 2 CCTT E-ACC M-CTT 1 ‘ CGAA E-ACG M-CAA 3 CGAC E-ACG M-CAC l CGTA E-ACG M-CTA 3 CGTC E-ACG M-CTC 5 CGTT E-ACG M-CTT 1 CTAA E-ACT M-CAA 4 CTAC E-ACT M-CAC 1 CTTA E-ACT M-CTA 6 CTTC E-ACT M-CTC 1 CTTT E-ACT M-CTT l GCAC B-AGC M-CAC 1 GCAT E-AGC M-CAT 5 GCTC E-AGC M-CTC 4 GGAA E-AGG M-CAA 3 GGAG E-AGG M—CAG 2 GGAT E-AGG M-CAT 3 GGTA E-AGG M4CTA 3 TEcoRl adapter (E) ligated to 3 selective nucleotides (i.e. AAC). * Msel adapter (M) ligated to 3 selective nucleotides (i.e. CAA). 'in AND696/619833 population with selective primer combination 1 and 2. 183 Table A4-2. Simple sequence repeat (SSR) primer sequence, linkage group and source information for markers polymorphic between AND696 and G19833 and used to develop a linkage map of AND696/G19833. SSR PrimerT Primer Sequence Linkage Source* group; AGl F CAT GCA GAG GAA GCA B03 2 GAG TG R GAG CGT CGT CGT TTC GAT ATA] 3 ‘F unpublished B03 5 R ATA4 F unpublished B01 5 R ATA9 F unpublished B09 5 R BM114 F AGC CTG GTG AAA TGC B09 2 TCA TAG (unlinked) R CAT GCT TGT TGC CTA ACT CTC T BM137 F CCG TAT CCG AGC ACC B06 2 GTA AC ' R CGC TTA CTC ACT GTA CGC ACG BM14O F TGC ACA ACA CAC ATT B04 2 TAG TGA C R CCT ACC AAG ATT GAT TTA T GG G BM143 F GGG AAA TGA ACA GAG B02 2 GAA A R ATG TTG GGA ACT TTT AGT GTG BM154 F TCT TGC GAC CGA GCT B09 2 TCT CC R CTG AAT CTG GGA ACG ATG ACC AG BM156 F CTT GTT CCA CCT CCC B02 2 ATC ATA GC R TGC TTG CAT CTC AGC CAG AAT C BM159 F GGT GCT GTT GCT GCT B03 2 GTT AT R GGG AGA TGT GGT AAG ATA ATG AAA 184 Table A4-2 (cont’d). SSR BM160 BM161 BM164 BM165 BM167 BM170 BMZOO BM211 BM53 BMd-l PrimerT Primer Sequence F R CGT GCT TGG CGA ATA GCT TTG CGC GGT TCT GAT CGT GAC TTC TGC AAA GGG TTG AAA GTT GAG AG TTC CAA TGC ACC AGA CAT TCC CCA CCA CAA GGA GAA GCA AC ACC ATT CAG GCC GAT ACT CC TCA AAT CCC ACA CAT GAT CG TTC TTT CAT TCA TAT TAT TCC GTT CA TCC TCA ATA CTA CAT CGT GTG ACC CCT GGT GTA ACC CTC GTA ACA G AGC CAG GTG CAA GAC CTT AG AGA TAG GGA GCT GGT GGT AGC TGG TGG TTG TTA TGG GAG AAG ATT TGT CTC TGT CTA TTC CCA C ATA CCC ACA TGC ACA AGT TTG G CCA CCA TGT GCT CAT GAA GAT AAC TAA CCT CAT ACG ACA TGA AA AAT GCT TGC ACT AGG GAG TT CAA ATC GCA ACA CCT CAC AA GTC GGA GCC ATC ATC TGT TT 185 Linkage group; BO7 B04 B02 unlinked (1301) B02 . (unlinked) B06 B01 B08 B01 B03 (unlinked) Source* 2 Table A4-2 (cont’d). SSR Primer’r Primer Sequence Linkage Source* group: BMd-lO F GCT CAC GTA CGA B01 3 GTT GAA TCT CAG R ATC TGA GAG CAG CGA CAT GGT AG BMd-3 F TGT TTC TTC CTT ATG unlinked 3 GTT AGG TTG (BO7) R GTA TCC TCC GAT CAA ATT CAC CT BMd—42 F TCA TAG AAG ATT B10 3 TGT GGA AGC A R TGA GAC ACG TAC GAG GCT GTA T Gats9l F GAG TGC GGA AGC B02 2 GAG TAG AG R TCC GTG TTC CTC TGT CTG TG MEI F unpublished B09 5 R PVct002 F TTA GAC TTT CAA B08 4 ACA TTC AC R GAT ACT ACT TAA ATG AGG AAC A PVttc002 F ATA TCT TAC AGC CAT B08 4 TAC ATT C R CTC ATC ACC CAG TCA CCT _ PVcttOOl F CCA ACC ACA TTC 'ITC B04 1 CCT ACG TC R GCG AGG CAG TTA TCT TTA GGA GTG PVagOOl F CAA TCC TCT CTC TCT B11 1 CAT TTC CAA TC R GAC CTT GAA GTC GGT GTC GTT T PVatOO3 F ACC TAG AGC CTA B04 1 ATC CTT CTG CGT R . PVgaatOOl F AAG GAT GGG TTC B04 1 CGT GCT TG R CAC GGT ACA CGA AAC CAT GCT ATC 186 Table A4—2 (cont’d). SSR PrimerT Primer Sequence Linkage Source* MP1 PVat007 F AGT TAA ATT ATA CGA B09 1 GGT TAG CCT AAA TC R CAT TCC CTT CAC ACA TTC ACC G 1' Primer orientation: forward (F); reverse (R) iBean linkage group location of SSR in the consensus map (Freyre et al., 1998), if different than where it mapped to in AND696/G19833 population, that location is noted in parenthesis. * 1 :Yu et al., 2000. Journal of Heredity 91(6):429-434. 2 : Gaitan—Solis etal., 2002. Crop Science 42:2128-2136. 3 : Blair et al. 2003. Theoretical Applied Genetics 107:1362-1374. 4 : Caixeta et al., 2005. Crop Breeding and Applied Biotechnology 5:125-133. 5: Metais et al., 2002. Theoretical Applied Genetics 104:1346-1352. (full citations available in Chapter 4, References) 187 Table A4-3. Random Amplified Polymorphism DNA (RAPD) primer sequence for markers polymorphic between AND696 and G19833 and used to develop a linkage map of AND696/G1 9833 recombinant inbred line population. RAPD name Primer sequence Polymorphism numberT AA-19 5 ’-TGAGGCGTGT-3 ’ 1 AK-6 5’-TCACGTCCCT-3’ 2 H-l2 5’-ACGCGCATGT-3’ 2 L-4 5’-GACTGCACAC-3’ I I-6 5’-AAGGCGGCAG-3’ 1 M-2 S’-ACAACGCCTC-3’ 1 N—12 S’-CACAGACACC-3’ 1 R-4 5’-CCCGTAGCAC-3’ 1 0-12 5’-CAGTGCTGTG-3’ 2 lNumber of polymorphisms with this primer in the AND696/G19833 recombinant inbred line population. 188 Table A4-4. Bartlett test for variance homogeneity based on growth habit, where there were two classes, determinate and indeterminate, and 1 degree of freedom. Tests were conducted for seed traits measured under low P (LP) and high P (HP) treatments in the recombinant inbred line population developed from a cross between common bean lines AND696 and G19833. Traits P Chi-square Pr>Chi- level Square 2005 Seed weight HP 13.25 0.0003* LP 1.02 0.3135 Seed P HP 2.66 0.1029 LP 1.17 0.2799 Seed Fe HP 1.45 0.2290 LP 1.51 0.2190 Seed Zn HP 1.47 0.2261 LP 12.60 0.0004* Seed phytic HP 0.55 0.4598 acid LP 22.95 <0.0001* 2000 Seed weight HP 1.57 0.2095 LP 0.07 0.7903 Seed P HP 6.44 0.0112* LP 0.31 0.5781 Seed Fe HP 0.06 0.8004 LP 0.88 0.3484 Seed Zn HP 0.96 0.3260 LP 1.00 0.3173 * Variance is not homogeneous between grth habit classes 189 \Ilflllllllgflllllllllfllfll111111))!