. h. r“ 8.3%! it 1 , .. .V ... LR: R r. nor L ..uuv .3. , . 3mm {$43.3}: .»a.. ,rfwwtufinw , u . I!M§p}.l.i l‘ ... J ‘ anhyfig ray w... am. I unimpvlk . zikuhr my”. 51‘! “\fa‘l‘u é . 39 ‘ v (LI: 0.) V .l. . Inlzm. . 1,. 4. 54:29 .i ...;lslxulfir) Lénéfl.x1y.m1.vt '7. f) az‘usmM-Ingtld 1.1%.... Ii.1.Un4‘.vd...d2’iiv3;ilnt~ll.n Tl . lit-I ADO... ; ..I‘ \l £71.!!le ‘1. ' wt nmNWIVIi.EL {L . l azll? , THEQS «v Illmllilllllllll‘lllllllwill (.900 C“ 3 1293 01834 1838 LIBRARY ‘ Michigan State University This is to certify that the thesis entitled EFFECTS OF COLD TREATMENT, PHOTOPERIOD, AND FORCING TEMPERATURE ON OENOTHERA FRUTICOSA 'YOUNGII-LAPSLEY' AND STOKESIA LAEVIS 'KLAUS JELITTO' presented by Emily A. Clough has been accepted towards fulfillment of the requirements for M. S . degree in Horticulture W Major professor Date 8’ /3 '?9 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 1M chlHCJD‘oOuopfi-p“ EFFECTS OF COLD TREATMENT, PHOTOPERIOD, AND FORCING TEMPERATURE ON OENOTHERA FRUTICOSA ‘YOUNGll-LAPSLEY’ AND STOKES/A LAEVIS ‘KLAUS JELITTO’ By Emily A. Clough A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Horticulture 1 999 ABSTRACT EFFECTS OF COLD TREATMENT, PHOTOPERIOD, AND FORCING TEMPERATURE ON OENOTHERA FRUTICOSA ‘YOUNGll-LAPSLEY’ AND STOKES/A LAEVIS ‘KLAUS JELITTO’ By Emily A. Clough A series of experiments were conducted to determine the effect of cold treatment, photoperiod, and forcing temperature on Oenothera fruticosa L. 'Youngii-lapsley’ and Stokesia Iaevis (Hill) Greene ‘Klaus Jelitto’. ‘Youngii-lapsley’ had a nearly obligate cold requirement and, following cold treatment, behaved as a facultative long-day plant. Only 3 weeks of cold treatment were required for 100% flowering. Time to flower decreased slightly as cold treatment increased from 3 to 15 weeks. Stokesia Iaevis ‘Klaus Jelitto’ had a facultative cold requirement and was proposed to be a facultative intermediate-day plant because flowering was most consistent and rapid under photoperiods between 12 and 13 h both before and after cold treatment. Time to flower, flower number, and plant height at flower decreased as temperature increased from 14 °C to 29 °C for both ‘Youngii-Iapsley’ and ‘Klaus Jelitto’. The flowering Characteristics of five cultivars of Oenothera fruticosa were compared under either incandescent lamps or high-pressure sodium lamps used for daylength extension. Lamp types differed in light intensity as well as spectral quality. Flower number was increased two to three times for ‘Youngii-lapsley', ssp. glauca and ‘Highlight’ when high-pressure sodium lamps were used for daylength extension compared to incandescent lamps. To my sisters Alice C. Moore and Catherine A. Clough, for their support, friendship, and love ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Art Cameron, and all the members of my committee, Dr. Royal Heins, Dr. erl Carlson, and Dr. Jim Baird, for their support, advice, and encouragement. I greatly appreciate all the opportunities that l have been given here to increase my knowledge and improve myself. I thank Dr. Bridget Behe for her mentoring and inspiration. ' This research could not have been completed without the support of Dan Tschirhart and all the undergraduate student workers and I am very grateful for all their help. I also thank the nurseries, including Greenleaf Enterprises, Walters Gardens, Skagit Gardens, Blooming Nursery, and Bluebird Nursery which, very generously, provided the plant material for these experiments. I would also like to thank my friends and colleagues for their friendship, their insightful advice, and moral support: Sarah Bruce, Joaquin Chong, Beth Fausey, Leslie Finical, Alison Frane, Priscilla Hockin, Kathy Kelly, Bin Liu, Mary- Slade Morrison, Genhua Niu, Erik Runkle, and Cathy Whitman. I give special thanks to Juan Hernandez for his love and his firm belief in me. Finally, I give thanks to my family, for their constant love and support. TABLE OF CONTENTS LIST OF TABLES ................................................................................................ vii LIST OF FIGURES ............................................................................................... ix LITERATURE REVIEW ........................................................................................ 1 Introduction ................................................................................................ 2 Characterization of Vemalization ............................................................... 4 Effective temperatures ..................................................................... 5 Duration of cold treatment ................................................................ 6 Antivemalization and devemalization ............................................... 8 Plant age. ....................................................................................... 10 Light and vemalization ................................................................... 11 Perception of Vemalization ...................................................................... 14 Mechanism of Vemalization ...................................................................... 16 Grafting studies ............................................................................. 16 Gibberellins .................................................................................... 18 Genetics of Vemalization ......................................................................... 20 Conclusion ............................................................................................... 26 Literature Cited ........................................................................................ 27 GROWTH AND DEVELOPMENT OF OENOTHERA FRUTICOSA IS INFLUENCED BY COLD TREATMENT, PHOTOPERIOD, FORCING TEMPERATURE, AND PLANT GROWTH REGULATORS ................................ 32 Abstract ................................................................................................... 33 Introduction .............................................................................................. 35 Materials and Methods ............................................................................ 38 Results and Discussion ............................................................................ 44 Summary .................................................................................................. 55 Literature Cited ......................................................................................... 57 FLOWERING CHARACTERISTICS OF SEVERAL CULTIVARS OF OENOTHERA FRUTICOSA UNDER TWO DIFFERENT LAMPS TYPES FOR DAYLENGTH EXTENSION ................................................................................. 73 Abstract .................................................................................................... 74 Introduction ............................................................................................... 75 Materials and Methods ............................................................................. 76 Results and Discussion ............................................................................ 80 Summary .................................................................................................. 84 Literature Cited ......................................................................................... 85 EFFECTS OF COLD TREATMENT, PHOTOPERIOD, AND FORCING TEMPERATURE ON GROWTH AND DEVELOPMENT OF STOKESIA LAEVIS ‘KLAUS JELI'ITO’ ............................................................................................... 91 Abstract .................................................................................................... 92 Introduction ............................................................................................... 94 Materials and Methods ............................................................................. 97 Results and Discussion .......................................................................... 101 Summary ................................................................................................ 113 Literature Cited ....................................................................................... 114 APPENDIX A: COMPARISON OF STOKES/A LAEVIS CULTIVARS GROWN UNDER TWO DIFFERENT LAMP TYPES FOR DAYLENGTH EXTENSION..128 Introduction ............................................................................................ 129 Protocol .................................................................................................. 131 Summary of Results .............................................................................. 133 Literature Cited ...................................................................................... 135 APPENDIX B: EFFECTS OF COLD TREATMENT, PHOTOPERIOD, AND FORCING TEMPERATURE IN ANEMONE HUPEHENSIS, PENSTEMON DIGITALIS ‘HUSKER RED’, AND SAXIFRAGA ARENDSII ‘TRIUMPH’ ........... 139 Introduction ............................................................................................. 140 Protocol .................................................................................................. 142 Selected data for each species .............................................................. 145 vi LIST OF TABLES Table 1. Average daily air temperatures and daily light integrals from date of transplant to average date of flowering for Oenothera fruticosa ‘Youngii-lapsley’ after each duration of cold treatment .................................................................. 59 Table 2. Average daily air temperatures from date of transplant to average date of flowering for Oenothera fmticosa ‘Youngii-lapsley’ under each photoperiod..60 Table 3. Actual temperatures of houses for the time of forcing for three developmental stages: transplant to visible bud, visible bud to flower, and transplant to flower .............................................................................................. 61 Table 4. Parameters of linear regression analysis relating forcing temperature to rate of progress from transplant to visible bud (VB), visible bud to flowering (VB to FLW) and transplant to flowering (FLW) for Oenothera fruticosa ‘Youngii- lapsley’; intercept and slope were used to calculate base temperature (T t,) and degree-days (°days) ........................................................................................... 62 Table 5. Starting plant material; dates of arrival; cold treatment initiation; forcing; average daily temperature, and average daily light integral for Oenothera fruticosa ‘Fireworks’, ‘Highlight’, ‘Youngii-Iapsley’, ‘Summer Solstice’, and O. fruticosa ssp. glauca. The labels “A” and “B” after 0. fruticosa ssp. glauca refer to different sources for plant material. Lamp type = type of lamp used to deliver daylength extension for 16-h photoperiods. HPS = high-pressure sodium lamps; INC = incandescent lamps ................................................................................. 86 Table 6. Flowering characteristics of several cultivars of Oenothera fmticosa and O. fmticosa ssp. glauca grown at 20 °C under 16-h photoperiods, delivered by day extension with either high-pressure sodium (HPS) lamps or incandescent (INC) lamps. All cultivars were treated with 15 weeks at 5 °C. Labels of significance refer to intracultivar comparisons for lamp type .............................. 87 Table 7. Average daily air temperatures and daily light integrals from date of transplant to average date of flowering of Stokesia Iaevis ‘Klaus Jelitto’ after each duration of cold treatment ........................................................................ 116 Table 8. Average air temperatures from date of transplant to average date of flowering of Stokesia Iaevis “Klaus Jelitto’ under each photoperiod .................. 117 Table 9. Actual temperatures of houses for the time of forcing for three developmental stages: Transplant to Visible Bud, Visible Bud to Flower, and Transplant to Flower. ....................................................................................... 118 vii Table 10. Parameters of linear regression analysis relating forcing temperature to rate of progress from transplant to visible bud (VB), visible bud to flowering (VB to F LW) and transplant to flowering (FLW) for Stokesia Iaevis ‘Klaus Jelitto’. Intercept and slope were used to calculate base temperature (T,) and degree- days (°days). .................................................................................................... 1 19 Table 11. The difference between day and night temperature (DIF) of plants of Stokesia Iaevis ‘Klaus Jelitto’ grown under different forcing temperatures and three developmental stages: from transplant to visible bud (VB), from VB to flowering (FLW), and from transplant to FLW .................................................. 120 Table 12. Starting plant material; number of plants per treatment; dates of arrival, cold treatment initiation, forcing, as well as average daily temperature and average daily light integral for Stokesia Iaevis ‘Alba’, ‘Blue Danube’, ‘Klaus Jelitto’, ‘Mary Gregory’, ‘Purple Parasols’ and ‘Wyoming’. Lamp type = type of lamp used to deliver daylength extension for 16 h photoperiods. HPS = high pressure sodium lamps; INC = incandescent lamps ........................................ 136 Table 13. Flowering characteristics of several cultivars of Stokesia Iaevis grown at 20 °C under 16 h photoperiods, delivered by day-extension with either high pressure sodium (HPS) lamps or incandescent (INC) lamps. All cultivars were treated with 15 weeks at 5 °C. Labels of significance refer to intracultivar comparisons for lamp type. Number of new leaves refers to the number of new leaves beneath the first flower ......................................................................... 137 viii LIST OF FIGURES Figure 1. Responses of Oenothera fruticosa ‘Youngii-lapsley’ to various durations of cold treatment at 5 °C for percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Year 1 (I) or Year 2 (I). Error bars represent 95% confidence intervals. All symbols represent the mean of 9 or 10 plants except for Year 1, 0 week cold treatment, when only one plant flowered. L = linear; Q = quadratic trends. ”5' ”-"Nonsigniflcant or significant at P 50.01 or 0.001, respectively ....................... 63 Figure 2. Effect of cold duration at 5 °C on number of new nodes produced during forcing in Year 2. Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. ”5' '-"'”'Nonsignificant or significant at P s 0.05, 0.01, or 0.001 , respectively .......................................................................................... 64 Figure 3. Effects of cold duration at 5 °C on number of flowers and buds (A), number of flowering lateral shoots (B), percent flowering lateral shoots (C), and plant height (D) for Oenothera fruticosa ‘Youngii-lapsley’ in Year 1 (O) or Year 2 (I). All symbols represent the mean of 9 or 10 plants except for Year 1, 0 week cold treatment, when only one plant flowered. Error bars represent 95% confidence intervals. L = linear, Q = quadratic trends. "3' ""-'"Nonsignificant or significant at P s 0.05, 0.01 or 0.001, respectively .............................................. 65 Figure 4. Responses of Oenothera fruticosa ‘Youngii-Iapsley’ to various photoperiods after 15 weeks of cold treatment at 5 °C for days to flower (A), days from visible bud to flower (B), and days to visible bud (C) in Year 1 (o) and Year 2 (I). Continual photoperiod treatments consisted of 9-h natural days extended with light from incandescent lamps. Nl = 4-h night intenuption. Error bars represent 95% confidence intervals. L = linear“, Q = quadratic trends. ”5- '-"-"'Nonsignificant or significant at P s 0.05, 0.01, or 0.001 , respectively ........... 66 Figure 5. Effects of various photoperiods after 15 weeks at 5 °C on number of final nodes (A) and number of nodes produced during forcing (B) for Oenothera fruticosa ‘Youngii-lapsley’ in Year 1 (I) or Year 2 (I). Continual photoperiod treatments consisted of 9-h natural days extended with light from incandescent lamps. NI = 4-h night interruption. Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. "3' ""'”'Nonsignificant or significant at P s 0.05, 0.01, or 0.001, respectively ................................................................. 67 Figure 6. Effects of various photoperiods after 15 weeks at 5 °C on number of flowers and buds (A), number of flowering lateral shoots (B), number of total lateral shoots (C), percent flowering laterals (D), and plant height (E) for Oenothera fmticosa ‘Youngii-Iapsley’ in Year 1 ( I) or Year 2 (I). Continual photoperiod treatments consisted of 9-h natural days extended with light from ix incandescent lamps. NI = 4-h night interruption. Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. ”3- "”-'" Nonsignificant or significant at P s 0.05, 0.01, or 0.001, respectively ............................................. 68 Figure 7. Effects of temperature on time to (A, B, C) and rate of progress toward flowering (D, E, F) in Oenothera fruticosa ‘Youngii-lapsley’. The parameters of linear regression lines are presented in Table 4. The quadratic regression lines in graphs A, B, and C are the reciprocals of correlated linear regression lines in graphs D, E, and F .............................................................................................. 69 Figure 8. Effects of temperature on numbers of flowers and buds (A), flower diameter (B), and plant height (C) for Oenothera fruticosa ‘Youngii-Iapsley’. Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. ”5' "’"Nonsignificant or significant at P s 0.01 or 0.001, respectively .................. 70 Figure 9. Effects of various plant growth regulators on days to flower (A), plant height (B), and flower diameter (C) for Oenothera fruticosa ‘Youngii-lapsley’. Rates of PGRs were: Ancymidol (100 mg-L“), Daminozide (5000 mg-L“), Paclobutrazol (30 mg-L"), Chlorrnequat (1500 mg-L") and Uniconazole (15 mg-L’ 1). Error bars represent 95% confidence intervals .............................................. 71 Figure 10. Postharvest responses of Oenothera fruticosa ‘Youngii-lapsley’ for 12 days at 24.6 °C under a 24-h photoperiod. Error bars represent 95% confidence intervals ............................................................................................................... 72 Figure 11. Cultivars of Oenothera fruticosa and O. fmticosa ssp. glauca grown at 20°C under a 16-h photoperiod provided by a 9-h natural day plus 7-h day extension with either high-pressure sodium lamps (labeled in the photographs as HID for high-intensity discharge) or incandescent lamps (inc). The labels “A” and “B” after ssp. glauca refer to different nursery sources ....................................... 89 Figure 12. Effects of various durations of cold treatment at 5 °C on percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Stokesia Iaevis ‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. NS""'"Nonsignificant or significant at P s 0.01 or 0.001, respectively ................. 121 Figure 13. Effects of cold duration at 5 °C on number of nodes (A), number of flower buds (B), and plant height (C) in Stokesia Iaevis ‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. ”5" "’"Nonsignificant or significant at P s 0.05, 0.01, or 0.001 , respectively ............................................................................................ 122 Figure 14. Effects of various photoperiods without a cold treatment on percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Stokesia Iaevis ‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. NS-'°"-'"Nonsignificant or significant at Ps0.05, 0.01, or 0.001, respectively ....................................................................................................... 123 Figure 15. Effects of various photoperiods after a 15 week cold treatment at 5 °C on percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Stokesia Iaevis ‘Klaus Jelitto' in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. NS""'-"'Nonsignificant or significant at P s 0.05, 0.01, or 0.001, respectively ....................................................................................................... 124 Figure 16. Effects of various photoperiods with and without a 15—week cold treatment at 5°C on number of final nodes (A), number of new nodes (B), number of flower buds (C), and plant height (D) in Stokesia Iaevis ‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear; Q = quadratic trends. NS"""""Nonsignificant or significant at P s 0.05, 0.01, or 0.001, respectively ........................................................................................ 125 Figure 17. Effects of temperature on time to (A, B, C) and rate of progress toward flowering (D, E, F) in Stokesia Iaevis ‘Klaus Jelitto’. The parameters of linear regression lines are presented in Table 10. The quadratic regression lines in graphs A, B, and C are the reciprocals of correlated linear regression lines in graphs D, E, and F ............................................................................................ 126 Figure 18. Effects of temperature on number of nodes (A), number of flower buds (B), and plant height (C) in Stokesia Iaevis ‘Klaus Jelitto’. L=linear; Q=quadratic trends. ”5'" "’"Nonsignificant or significant at P s 005,001, or 0.001 , respectively ............................................................................................ 127 Figure 19. Effects of cold treatment in Anemone hupehensis in Year 1 .......... 145 Figure 20. Effects of cold treatment in Anemone hupehensis in Year 2. Error bars show 95% confidence intervals ................................................................. 146 Figure 21. Effects of various photoperiods with and without a 15 week cold treatment at 5 °C in Anemone hupehensis in Year 1 ........................................ 147 Figure 22. Effects of various photoperiods with and without a 15 week cold treatment at 5 °C in Anemone hupehensis in Year 2. Error bars show 95% confidence intervals .......................................................................................... 148 Figure 23. Influence of forcing temperature on flowering time (A,B,C) and rate of progress toward flowering (D,E,F) in Anemone hupehensis. Error bars represent 95% confidence intervals .................................................................. 149 xi Figure 24. Influence of forcing temperature on number of nodes (A), number of flower buds (B), and plant height at first flower (C) in Anemone hupehensis. Error bars represent 95% confidence intervals .......................................................... 150 Figure 25. Effects of cold treatment in Penstemon digitalis 'Husker Red' in Year 1 ........................................................................................................................ 1 51 Figure 26. Effects of cold treatment in Penstemon digitalis 'Husker Red' in Year 2. Error bars show 95% confidence intervals .................................................... 152 Figure 27. Effects of various photoperiods with and without a 15-week cold treatment at 5 °C in Penstemon digitalis ‘Husker Red’ in Year 1 ...................... 153 Figure 28. Effects of various photoperiods with and without a 15-week cold treatment at 5 °C in Penstemon digitalis ‘Husker Red’ in Year 2. Error bars show 95% confidence intervals .................................................................................. 154 Figure 29. Influence of forcing temperature on flowering time (A, B, C) and rate of progress toward flowering (D, E, F) in Penstemon digitalis ‘Husker Red’. Error bars represent 95% confidence intervals ................................................. 155 Figure 30. Influence of forcing temperature on number of nodes (A), number of flower buds (B), and plant height at first flower (C) in Penstemon digitalis ‘Husker Red’. Error bars represent 95% confidence intervals .......................... 156 Figure 31. Effects of cold treatment in Saxifraga arendsii 'Triumph' in Year 1 ............................................................................................................ 157 Figure 32. Effects of cold treatment in Saxifraga arendsii 'Triumph' in Year 2. Error bars show 95% confidence intervals ....................................... 158 Figure 33. Effects of various photoperiods with and without a cold treatment in in Saxifraga arendsii 'Triumph’ in Year 1 ........................................................... 159 Figure 34. Effects of various photoperiods with and without a cold treatment in in Saxifraga arendsii 'Triumph' in Year 2. Error bars show 95% confidence intervals ............................................................................................................. 1 60 Figure 35. Influence of forcing temperature on flowering time (A, B, C) and rate of progress toward flowering (D, E, F) in Saxifraga arendsii ‘Triumph’. Error bars represent 95% confidence intervals .......................................................... 161 Figure 36. Influence of forcing temperature on number of nodes (A), number of flower buds (B), and plant height at first flower (C) in Saxifraga arendsii ‘Triumph’. Error bars represent 95% confidence intervals ................................ 162 xii Literature Review Introduction Vemalization is defined as the promotion of flower formation by exposure to low temperatures (Lang, 1965). It is an inductive process, the effects of which generally are not observed until the chilled plant is moved to higher temperatures (Chouard, 1960; Vince-Prue, 1975). Vemalization is not required for flowering of all plant species; most that respond to vemalization originated in temperate latitudes where it is ecologically advantageous for a plant not to flower during the winter months (Roberts and Summerfield, 1987). Some species have a quantitative, or facultative, response to a cold treatment and flower without one, but flower development is hastened afler a cold treatment. Plants with either a qualitative or quantitative response to cold treatment are usually long-day plants, although there is no definitive relationship between a vemalization requirement and a specific photoperiod requirement (Vince-Prue, 1975). Most species native to tropical or subtropical regions, where there is no need to avoid a period of lethally cold temperatures, are insensitive to vemalization (Roberts and Summerfield, 1987). The first scientific work on vemalization was done primarily on winter wheat and rye. Winter races are fall-planted and require the cold temperatures to hasten flowering the following spring. The spring races differ in that they do not have a requirement for cold and will flower soon after sowing. While the phenomenon that we now call vemalization was recognized as eany as the mid- nineteenth century (Klippart, 1857, cited by Whyte, 1948), scientific definition did not occur until 1918 when Gassner (cited by Whyte, 1948) showed that a great variety of plants had a cold requirement for flowering. Lysenko, working with winter rye, termed the phenomenon in his native Russian "jarovizacija" (a combination of the word for "spring" with a suffix meaning "to make") and translated it himself into English, French, and German as ”vemalization” (Whyte, 1948; Chouard, 1960). Lysenko was also the first to propose a model to explain the role of vemalization in plant development. He called it the "theory of phasic development,” in which he suggested that plants must pass through two irreversible phases in order to flower: the temperature- sensitive phase (therrnophase) and the light-sensitive phase (photophase) (Chouard, 1960; Napp-Zinn, 1984). Today, we know that only a few plants respond exactly according to this theory and it has been soundly rejected (Napp- Zinn, 1984). Throughout this century, a great deal of information concerning vemalization has been compiled and the characterization of the vemalization response is well described for many species. The genetic and physiological aspects of vemalization, however, have yet to be described in detail. Many genes from several different species are important in vemalization, but their mode of action is unknown. In this review, I will present a thorough characterization of the vemalization response and a summary of current progress on the underlying genetic and physiological mechanisms. Characterization of vemalization Gregory and Purvis defined the fundamental characteristics of vemalization in their many experiments with the rye (Secale cereale) cultivar Petkus which consists of two races- winter and spring rye (Whyte, 1948; Chouard, 1960; Vince-Prue, 1975). They observed that both races flower faster when kept in continuous light and that neither will flower if the photoperiod is shorter than 12 hours (in Vince-Prue, 1975). Without a cold treatment (also called "therrnoinduction," the actual exposure of a plant to cold temperatures), the winter race flowers in continuous light after 15 weeks when 22 leaves have been formed beneath the floral spike. The spring race flowers in about 7.5 weeks and produces six to seven leaves beneath the floral spike. After a cold treatment, the winter race behaves exactly like the spring race. The vemalization response is seen as an aftereffect of the low- temperature treatment (Chouard, 1960), and therefore the response is typically measured by the time to anthesis, the leaf number beneath the flower, or both (Purvis, 1961 ). Currently, there are no genetic or physiological markers that can be used to provide a definitive assay of the progress of vemalization during a low- temperature treatment (Brooking, 1996); hence, the classical methods of measuring the vemalization response (e.g., final leaf number at flower, time to anthesis) still are used. Effective temperatures. Effective temperatures for vemalization of many species occur between -5 °C and 15 °C, with a broad optimum between 1 °C and 7 °C (Chouard, 1960; Lang, 1965; Thomas and Vince-Prue, 1984). The response generally increases with temperatures between -4 °C and 1 °C and decreases with temperatures above 9 °C (Purvis, 1961). Although this range is effective for most species, there are exceptions. For example, Easter lily (Lilium Iongiflorum Thunb.) 'Ace' is effectively vemalized at temperatures up to 16 °C and partially vemalized at 18 °C. No vemalization occurs at 21 °C (Weiler and Langhans,1968) The optimum temperature for vemalization may depend on the length of the cold treatment (Purvis, 1961). For example, experiments with black henbane (Hyoscyamus niger L.) and ‘Petkus’ rye showed a higher optimum temperature when the duration was shorter. However, when the cold treatment was extended, any temperature within the vemalization range produced the same effect (Purvis, 1961; Vince-Pme, 1975). In contrast, experiments with the Cineran'a L. ’Cindy Blue' showed that optimum temperature was little affected by duration of cold treatment (Yeh et al., 1997a). The optimum temperature also may depend on cultivar or the method by which the vemalization response is assessed. Rawson et al. (1998) studied several cultivars of wheat differing in their vemalization requirements at several temperatures. The moderate winter cultivar Oxley had an optimum vemalization temperature of 8 to 10 °C when the response was measured as days from sowing to heading, 6 °C when measured as summed thermal time (°Cd) above a base of 0 °C, and 3 °C when measured as minimum final leaf number at heading. Different optima can be obtained when the vemalization response is measured as days from end of cold treatment to heading or days from sowing to heading. The winter cultivar JF87%014 had an optimum temperature of 8 °C when the response was measured as days from the end of cold treatment to heading; 6 °C, from sowing to heading (Rawson et al., 1998). Robertson et al. (1996) point out that there is a problem in finding an optimum temperature for vemalization because the effectiveness of any temperature in lowering final leaf number is confounded with its effect on vegetative development. They found that the saturataion rate of vemalization in various winter-wheat cultivars (they considered vemalization saturated in treatments in which the number of leaves initiated at the end of the treatment equaled the final number of leaves produced beneath the flower) responded linearly to temperatures between 5 and 14 °C. Vemalization’s effectiveness in decreasing the number of leaves below the flower decreased with temperature, and there was no optimum temperature. They suggested that higher temperatures were less effective because the rate of leaf primordium initiation responds more strongly to temperature than does that of vemalization (Robertson et al., 1996). Duration of cold treatment. Vemalization is a quantitative process whose intensity increases with the length of the exposure until a saturation response (here defined as no further reduction in time to flower or final leaf number) occurs (Vince-Prue, 1975). The duration of time to saturation is species- and cultivar- specific, but usually will occur between 10 and 100 days of low-temperature treatment (Thomas and Vince-Prue, 1984). Increasing durations of cold treatment influence other plant characteristics such as height, flower number, and days to flower. Inflorescence number of Lavandula angustifolia Mill. grown at 20 °C under a 9-hour day plus a 4-hour night interruption (NI) increased from 2 to 11 as cold duration (at 5 °C) increased from 0 to 15 weeks (Whitman et al., 1996). The increase in flower number with cold treatment was not as dramatic when L. angustifolia was grown under 9-hour short days. Flowering did not occur until plants had received a minimum of 10 weeks of cold, and flower number increased only slightly (from approximately 5 to 6) when cold duration was increased from 10 to 15 weeks. Inflorescence number of Leucanthemum xsuperbum Bergman ex J. Ingram ‘Snowcap‘ grown at 20 °C under NI increased from 5 to 13 as cold treatment (at 5 °C) increased from 0 to 15 weeks (Runkle et al., 1998). Plant height of 'Snowcap‘ increased from 14 to 22 cm as cold duration increased from 0 to 12 weeks. Stability of the vemalization response, or the length of time that a chilled plant maintains the inductive effects of the cold treatment, increases with duration of cold treatment. Picard (1965) found that for Oenothera biennis L. var. Sulfurea, the vemalized state was lost after exposure to short days (8 hours) at 18 to 22 °C for 12 to 20 days. Antivemalization and devemalization. High temperatures (generally above 30 °C) given either before or after the cold treatment can influence the vemalization response. Several researchers observed that when high temperatures were given prior to the cold treatment, there was a nullification of, or a decrease in, the vemalization response (Lang, 1965). This effect was termed "antivernalization." A recent study of the effect of high temperature before cold treatment on subsequent flowering of Brassica napus L. var. annua (a plant that has a quantitative response to cold) showed that flowering of several genotypes was delayed by the heat treatment (30 °C) (Dahanayake and Galwey, 1998). For example, plants of genotype TB 14 exposed to a 30 °C treatment for 2 days before an 8-week 4 °C treatment flowered 8 days later than unexposed cold- treated control plants. Extending the heat treatment from 2 to 4 days further delayed flowering of this genotype; these plants flowered 13 days later than the control plants (Dahanayake and Galwey, 1998). Devemalization is the elimination or reduction of the vemalization response when a heat treatment is given immediately after the cold treatment. Temperatures from 20 to 40 °C generally cause devemalization (Napp-Zinn, 1961). Temperatures between 12 and 15 °C cause neither vemalization nor devemalization in many winter cereals (Lang, 1965). Devemalization may occur only when the duration of cold temperatures is relatively short. Purvis and Gregory (1952, Cited by Purvis, 1961) found that after 8 or more weeks of cold, winter rye could not be devemalized. They also noted that even partially vemalized plants could not be devemalized if they were kept at a neutral temperature for a few days before the heat treatment was applied. The vemalization response then was deemed to be ”fixed" or "stabilized" (Purvis, 1961; Lang, 1965). Recent work with the cineraria cultivars Cindy Blue and Cindy Dark Red supports the findings of Purvis and Lang (Yeh et al., 1997b). Temperatures of 20 °C or less promoted devemalization when the cold duration (at 6 °C) was 1 or 2 weeks. Devemalization response increased with temperature and length of heat treatment. Cultivars stabilized after 3 to 9 days at 15 °C, even plants that were only partially vemalized. In contrast, Brassica napus var. annua did not stabilize rapidly even when cooled plants were maintained at 15 °C for one week prior to heat treatment at 30 °C (Dahanayake and Galwey, 1998). In fact, devemalizing heat treatments were more effective after 1 week of stabilization. It was only after 2 weeks at the neutral temperature (15 °C) that devemalization was not as effective. Not all species can be devemalized by a simple high temperature treatment. A cold-requiring variety of Chrysanthemum did not devemalize completely even when exposed to 35 °C for 30 days (Schwabe, 1955). However, this variety was devemalized by a low-intensity light treatment of about 25 foot-candles (given by incandescent bulbs) combined with high temperature (28 °C) for 10 to 20 days although neither treatment alone had a measurable effect (Schwabe, 1957). Napp-Zinn (1960, cited by Lang, 1965) found that, for Arabidopsis thaliana (L.) Heynh. , devemalization did not occur if plants were kept in light during the high temperature treatment. Lang (1965) cautions that the inhibitory action of high light with respect to devemalization indicated by the results with Arabidopsis and Chrysanthemum actually may be a general response, since many devemalization experiments were conducted in darkness or weak light. Plant age. The age at which plants are able to perceive cold temperatures for effective vemalization is very species specific. Some species such as ‘Petkus’ rye and winter wheat are capable of perceiving low temperatures soon after fertilization, while others such as Hyoscyamus niger must be at least 10 days old (Purvis, 1961). In many cases, younger plants require a longer duration of cold to achieve vemalization saturation than do older plants. When Gott (1950, cited by Purvis, 1961) chilled rye plants of different ages, she found a decline in response when plants were older than 2 weeks. Picard (1965) found that young plants of O. biennis (with a minimum of 10 leaves) required 17 weeks of cold treatment (at 11 °C/3 °C D/N) for 100% flowering, while older plants (minimum 50 leaves) required 13 weeks for 100% flowering. Recent work with two cultivars of winter wheat (Triticum aestivum L. Pioneer 2548 and Augusta) showed that the estimated minimum days of cold treatment required to reach saturation of the vemalization response decreased linearly with increasing plant age at the beginning of the cold treatment (Wang et al., 1995). These authors suggest an 10 interchangeability between days of cold treatment and plant age. Robertson et al. (1996) proposed and tested a model to explain the interaction between plant age and vemalization. The model predicted the final leaf number below the flower based on the effect of temperature on the rate of vemalization saturation and rate of leaf initiation for wheat under continuous and intermittent low temperature. Plants grown under intermittent low temperatures progress toward flowering through a combination of rates, depending on the temperature at any given time. When Robertson et al. (1996) fitted the data of Wang et al. (1995) to their model, they found that the model satisfactorily predicted a range of final leaf numbers between 8 and 21, with an r2 of 88%. For the authors this model explained to a reasonable extent the interchangeability between plant age and vemalization. Plant age when the cold treatment is given can influence the stability of vemalization. Younger plants of O. biennis had lower vemalization stability than older plants when both were maintained under short days after cold treatment before being moved to inductive long days (Picard, 1965). Light and vemalization. The interactions between photoperiod, light quality, light quantity, and vemalization are complex and inconclusive. Napp-Zinn (1984) describes the often conflicting results in this area of research as discouraging and points out that there still is too little information. He also indicates that vemalization likely means different things in different species or even in different genotypes of the same species. 11 Vemalization can be substituted by an appropriate photoperiod in certain cases. Short-day vemalization, or the replacement of a cold treatment by short days, first was studied in the 19305 for ‘Petkus’ rye (Napp-Zinn, 1984). Wellensiek (1960) was the first to discover short-long day induction in a dicotyledon, Campanula medium L., a discovery Chouard (1960) refers to as "Wellensiek's phenomenon." Wellensiek (1960) observed that if a Campanula rosette was grown to three months of age and then exposed to at least 4 weeks of warm (20 °C) short days, it would flower when transferred to 16-hour long days. Flowering under short days only was never observed, even after 2 years. Increasing the number of short days did increase the percentage of plants that bolted, but did not regularly decrease the days to bolting. Wellensiek ( 1960) also noted that ”the reaction to warm short days is not as regular as to cold." An experiment using a winter race of barley (Hordeum vulgare L. 'Arabi Abiad') showed unequivocally that short days can substitute for vemalization in this species (Roberts et al., 1988). Photoperiods of 8 or 10 hours given after germination were sufficient to decrease the number of leaves beneath the flower by a factor of 3.3 when compared to that of plants given a 13-hour photoperiod. The minimum number of leaves formed with short-day vemalization is 9 or 10. Photoperiods longer than 13 hours did not increase the number of leaves beneath the flower. However, substitution by short days does not appear to be equal to vemalization by cold treatment since flowers can emerge after formation of 6 or 7 leaves if the plant is cold-treated (Roberts et al., 1988). 12 Substitution of a cold treatment by long days has been documented for several species. Pisum sativum L. genotype If e Sn hr kept in continuous light did not respond to vemalization; in fact, vemalized plants grown under continuous light flower at a node slightly higher than that of unvemalized plants (Murfet and Reid, 1974). When plants are grown under short days (8 hours), vemalization promotes flowering of genotype If e Sn hr 6 nodes lower than that of unvemalized plants. Runkle et al. (1998) found that, for L. xsuperbum 'Snowcap', 80% of plants grown at 20 °C under NI flowered without a cold treatment, while 80% of plants grown under a 9 hour short day required a minimum of 3 weeks at 5 °C. When Campanula Iongestyla Fomin. was grown under 16 hour long days, plants flowered without vemalization, but when the photoperiod was reduced to 14 hour, plants required vemalization (Mathon, 1960). Substitution of a cold treatment with light of particular radiation ratios or of high irradiances also has been demonstrated. Bagnall (1993) showed that for three late-flowering ecotypes of Arabidopsis (’Eifel’, ‘lnnsbruck', ‘Pitztal’), time to flower and leaf number at flower were halved when nonvemalized plants were grown under fluorescent and incandescent lights (R : FR = 1.0) compared to those grown strictly under fluorescent lighting (R :FR = 5.8). Bagnall terms this phenomenon of flower induction by different ratios of radiation "photoinduction." Work by Chouard during the 1960s revealed that several species, including Geum urbanum L. and Dactylis glomerata L., flowered without cold when they 13 were kept under long days or continual light at high irradiances, usually above 14,000 Ix. Napp-Zinn (1984) suggests that this effect may be due to photosynthesis, since it often is accompanied by a need for high mineral nutrition. In certain cases, the duration of light before the cold treatment can have an effect on subsequent flowering. Long days before cold treatment were beneficial for Oenothera biennis var. sulfurea because they caused a shorter juvenile period (Picard, 1965). Under long days (16 hours) 0. biennis perceived vemalizing cold temperatures 50 to 60 days after germination, while under short days (8 hours), the species could not perceive cold temperatures for 5 to 6 months at 18 to 22 °C. Experiments comparing different photoperiods during cold treatment yield inconsistent results. Long days, short days plus a night break, or continual light during cold treatment reduced subsequent flowering of celery, while they enhanced flowering of Arabidopsis 'Stockholm' (Napp-Zinn, 1984). For Campanula medium, there was no difference between plants exposed to long or short days during the 5 °C cold treatment (Wellensiek, 1960). The photoperiod during the cold treatment also had no influence on subsequent flowering of O. biennis (Picard, 1965). Perception of Vemalization Early experiments demonstrated that vemalization is perceived in the shoot tip. Evidence for this conclusion came from localized cooling experiments 14 in which either the roots or shoots were cooled separately while the rest of the plant was kept In noninductive temperatures (Lang, 1965). Gregory and Purvis (1938, Cited by Lang, 1965) showed that even fragments of embryos, which consisted of only the shoot tip, could respond to the low temperatures. Grafting studies showed that only when the shoot tip itself had been exposed to low temperatures would flowering occur (Lang, 1965). Metzger (1988) found that maintaining roots of Thlaspi arvense L. at 21 °C (while shoots were kept at 4 °C) reduced subsequent stem elongation, and he suggested that cold treatment of tissue other than the apex may be important for integration of reproductive development. Since the site of perception appears to be in the shoot tip, dividing cells may be necessary for vemalization to occur (Wellensiek, 1964; Lang, 1965). If this supposition is true, then any dividing cell in the plant may be vemalizable. Wellensiek (1964) provided support for this theory when he found that excised leaves of Lunaria biennis L. given a cold treatment could regenerate plants that immediately flowered. When he cut off the lowest 0.5 cm of the petiole, where regeneration and thus cell division takes place, the regenerated plants remained vegetative. In contrast to Wellensiek, Metzger (1988) observed that when shoots were regenerated from Thlaspi arvense leaf cuttings from which the lower 1 to 2 cm had been removed (which occurred in only 10 to15% of the cuttings), all shoots immediately flowered. He suggested that many cell types, not necessarily only actively dividing cells, may become vemalized. However, since 15 the outcome of vemalization is the ability to flower and flowering occurs on structures that arise from the shoot apex, it may be coincidental that the apex is considered the site of perception (Metzger, 1988). Mechanism of Vemalization To date, very little is known about the mechanism of vemalization. The role of classical plant hormones in the vemalization process is still unclear and evidence for a flower-forming hormone produced by therrnoinduced tissue is equivocal. In 1965, Lang stated that studies into the biochemical foundations of vemalization were inconclusive. Almost thirty years later, Hazebroek (1993) wrote, "Essentially nothing is known about the biochemical or molecular processes that occur during vemalization.” While attempts to elucidate the mechanism of vemalization have been frustrating, recent genetic studies have provided some new insight and may renew interest in the study of vemalization. Grafting studies. In several early experiments, the subsequent flowering of nontherrnoinduced receptor plants grafted onto therrnoinduced donor plants suggested the existence of a graft-transmissible flower-forming material (Lang, 1965). Successful grafts in which the nontherrnoinduced receptor flowered were made between many different species and response types. Lang (1965) commented that many of these experiments were performed in inductive long- day conditions for the donor, and that flowering of the receptor therefore may have been due to the transmissible material (also called "florigen") produced in 16 response to photoperiod and not necessarily just thermoinduction. Direct evidence for a transmissible material formed during therrnoinduction came from grafting nontherrnoinduced Hyoscyamus niger to the short-day plant Maryland ‘Mammoth tobacco’ (Melchers, 1939, cited by Lang, 1965). The H. niger receptor flowered under both long and short days, which indicated that a transmissible material (named 'vemalin”) was produced by the therrnoinduced tissue. Efforts to extract substances from therrnoinduced plants that induced flowering in nontherrnoinduced plants were unsuccessful (Lang, 1965). Grafting studies using Pisum sativum also have revealed evidence for a transmissible substance. Reid and Murfet (1975) grafted unvemalized or vemalized scions to unvemalized or vemalized stocks and used several different lines (both early- and late-flowering) of P. sativum. They found that chilling the stock of late-flowering lines (which contain the Sn gene) could promote flowering in unvemalized scions. In fact, vemalization of the stocks for these late-flowering lines more effectively reduced the node number below the flower than vemalization of the scions. They suggested that the Sn gene produces a flowering inhibitor and that vemalization suppresses its action. They also suggested that vemalization is active at two sites: the shoot tip and the cotyledons (the only node below the graft). While flowering of nontherrnoinduced receptor plants has been achieved in grafting studies, there were many other studies in which a nonthennoinduced receptor did not flower, regardless of the state of the donor (Lang, 1965). 17 Chouard (1960) states that, aside from the success of graft transmissibility in Hyoscyamus, flowering of a noninduced receptor on a therrnoinduced donor is not generally effective and is "only quite rarely successful." Gibberellins. The application of gibberellins (GAs) to nonthennoinduced plants produces a broad range of responses; for some species, GAs can induce flowering in nontherrnoinduced plants, while for others, they result in stem elongation but no flower formation (Lang, 1965). In general, treating seeds with GAs does not effectively induce flowering (Lang, 1965). Stem elongation precedes flower formation in GA-treated plants, while flowering and elongation occur simultaneously after a cold treatment (Lang, 1965). Applying inhibitors of GA biosynthesis can influence the flowering process. For example, uniconazole, a triazol GA biosynthesis inhibitor applied 3 days before a 20-day cold treatment, delayed both stem elongation and flowering of Raphanus sativus L. (Japanese radish), a strongly facultative cold- requiring plant (Nishijima et al., 1997). The effects of uniconazole on stem elongation and flowering reached a maximum at 2 pmollpot, at which rate flowering was delayed by 23 days and stem elongation was reduced by 48 cm. Increasing the cold treatment from 20 to 40 days moderated the effects of uniconazole only slightly. Days to stem elongation for plants cooled for 40 days were reduced by approximately 13 compared to that of plants cooled for 20 days. Days to flower and stem length were similar for plants given either duration of cold treatment. Timing of the GA inhibitor application was important: 18 if the application was made during the cold treatment, it reduced stem elongation and delaying flowering, but if it was made at or 5 days after the end of the cold treatment, flowering was not delayed (Nishijima et al., 1997). Zeevaart (1978) proposed that the response to GA may depend on using the "appropriate GA". For Myosotis alpestris F.W. Schmidt, GA7 induced both stem growth and flowering, while GA3 induced only stern growth (Michniewicz and Lang, 1962, cited by Zeevaart, 1978). Research of Wittwer and Bukovac (1962) also demonstrated a differential effect on flowering when different GAs were applied to lettuce. Both GA3 and GA1 were able to induce 100% flowering, while GA, and GA5 induced 30 to 40% flowering and GA; and GAB were unable to induce it. Zeevaart (1978) suggested that plants that do not respond to GA might not be receiving the appropriate form. Gibberellins may influence flowering of those plants with suboptimal cold treatments. In a study on vemalization and exogenous Gas’ effect on flower induction of Brassica oleracea L. var. botrytis (cauliflower), Fernandez et al. (1997) showed that GA(4+-,) accelerated flowering of plants grown at 22 °C (weak flower-inducing conditions) but not 10 °C (strong flower- inducing conditions). Plants grown at 10 °C produced a flower after 23 leaves had emerged, regardless of GA application. Plants that were not treated with GA and grown at 22 °C produced a flower after 38 leaves had emerged. Plants that were grown at 22 °C and treated with GA, however, produced a flower after 26 leaves had emerged (Fernandez et al., 1997). 19 For some species, levels of GAs increase after the cold treatment (Lang, 1965; Zanewich and Rood, 1995). Hazebroek et al. (1993) found that the activity of the enzyme kaurenoic acid hydroxylase, which converts kaurenoic acid to 7- OH kaurenoic acid (two members of the GA biosynthetic pathway), was increased after vemalization. In addition, the up-regulation of the enzyme was localized in the shoot tip and no change was observed in the leaves. For Brassica napus 'Crystal', Zanewich and Rood (1995) found that not only were endogenous levels of GAs increased (3.1, 2.3, 7.8, 12.0, and 24.5 times higher for GA1‘ GA3, GAB. GA19' and GAZO' respectively) after vemalization, but also the metabolism of GAs was changed. Vemalized plants showed an increase in the conversion of [3H]GA20 to an [3H]GA1-like metabolite and reduced conversion of [3HjGA20 or [3H]GA1 to polar [3H] metabolites, possibly representing GA glucosyl conjugates. The association between GAs and the flowering process also has been indicated by GA-deficient mutants. The rosette mutant of Brassica napus has about one-tenth of the normal level of total GAs and flowers very slowly compared to the wild type (Rood et al., 1989). Genetics of vemalization A vemalization requirement is a heritable property. In certain species, crosses between winter and spring races showed that the vemalization character seemed to follow predicted Mendelian ratios (Chouard, 1960). Over the years, genes, linkage groups, and cDNAs that either confer late flowering or are 20 associated with a vemalization requirement have been identified. Genes that influence the requirement for vemalization have been identified in several species. For such species as Tn'ticum, Arabidopsis, and Pisum, numbers of genes involved in vemalization have been deduced by the study of segregation ratios. It was postulated that there are four Tn’ticum loci, labelled Vm-1 ton-4 (Napp-Zinn, 1987). For Pisum, 20 genes initially were identified and were alleles of four individual loci. A dominant allele at either of two loci, Lf or Sn, confers late flowering. The requirement for vemalization is not always caused by the presence of a dominant allele. Recessive alleles appear to cause a vemalization requirement for Tn'ticum and Arabidopsis (Napp- Zinn, 1987). In Arabidopsis thaliana, the single FRI locus is apparently the cause for the difference between the winter ecotypes SF-2 and Le-0 and the spring ecotype Col. After the cold treatment, all ecotypes with the FRI locus, either naturally occurring or introduced, flower early (Lee et al., 1993). However, the action of FRI is not necessarily based on simply presence or absence of the locus. In the Ler background, the FRI locus causes only a delay in flowering when another locus, FLC, is present (Lee et al.,1994). Lee et al. (1994) suggest that FRI and FLC act to suppress flowering. The FRI locus also confers a strong responsiveness to light quality. Far red light accelerated flowering of plants containing the FRI locus when they received no cold treatment (Lee and Amasino, 1995). After a cold treatment, the acceleration of flowering time by far red light was eliminated. Since vemalization 21 did not influence the effect of far red light on petiole extension, the authors concluded that the effects of vemalization on flowering occur after far red light perception. In contrast to situations in which one or a few genes are responsible for the vemalization requirement, sometimes there are cases where it appears that many genes may be involved. Teutonico and Osborn (1995) mapped quantitative trait loci (QTL) for flowering time in Brassica rapa L. They found two linkage groups associated with variation for flowering time in unvemalized plants. After vemalization, neither of these QTL had a significant effect on flowering time. It is interesting to note that an RFLP locus identified with a clone from the Arabidopsis cold-induced gene COR6.6 mapped very close to one of the B. rapa QTL (T eutonico and Osborn, 1995). Osborn et al. (1997) compared the QTL for flowering time in B. rapa with QTLs in B. napus and A. thaliana. By aligning map positions they found that the two QTL from B. rapa corresponded to two major QTL for flower timing already identified in B. napus. The RFLP loci were mapped for both B. napus and A. thaliana by using Arabidopsis probes for flowering time genes and one of the Brassica QTL was collinear with the top of chromosome 5 of Arabidopsis. This area contains the locus FLC and various other flowering loci (Lee et al., 1994). The other Brassica QTL showed fractured collinearity with several regions of the Arabidopsis genome, one of which was the top of chromosome 4 where FRI is located (Lee et al., 1993). Recently, three mutants with altered vemalization responses were 22 identified at three independent loci in Arabidopsis (Chandler et al., 1996). Each of the mutants was induced by ethyl methane sulphonate (EMS) in the ice mutant in the Ler background, which itself shows late flowering and a strong response to vemalization. The three loci identified were named VRN1, VRN2, and VRN3. Before vemalization, vm1-1 fca-1 flowered at the same time as fca-1 but, after vemalization, flowered about 9 days later (a 42% reduction of the vemalization response); vm2-1 fca-1 differed from vm1-1 fca-1 because it flowered later than fca-1 both before and after vemalization. The upper arm of chromosome 3 contained VRN1 (Chandler et al., 1996). Chong et al. (1994; 1998) identified two cold-induced cDNAs (verc17 and verc203) thought to be specific for the vemalization response in wheat. The verc17 appeared in vemalized winter wheat tissue only; it was absent in nonthennoinduced and devemalized plants. The verc1? CDNA had significant homology with such diverse genes as a L. lactis Plll-type proteinase, maturation protein genes, and a major merozoite surface antigen gene. When wheat plants were transformed with the antisense RNA of the vemalization-specific CDNA ver0203 under control of the CaMV 358-promoter, it was found that flowering was greatly delayed (z160 days to flower for antisense compared to 2102 days to flower for the natural orientation of the DNA). These results suggest that the VER203 protein has a role in controlling flower development in winter wheat (Chong et al., 1998). Although progress has been made in identifying genes involved 23 associated with vemalization, little is known about how their function in creating the phenomenon known as vemalization. Burn et al. (1993) found that flowering in Thlaspi arvense and Arabidopsis thaliana (late ecotype) could be accelerated significantly if unvemalized plants were treated with 5-azacytidine, a DNA demethylating agent. In contrast, treatment with 5-azaC did not influence flowering of vemalized plants. When the treated plants were allowed to set seed, the resulting plants behaved as if unvemalized. After cold or treatment with 5-azacytidine, both Nicotiana plumbaginifolia L. and Arabidopsis show reduced levels of 5-methylcytosine in their DNA. The authors suggested that methylation status could be a control mechanism of vemalization. Brock and Davidson (1994) also found that 5-azacytidine as well as gamma rays (thought to be a demethylating agent) could substitute partially for a cold treatment. The 5- aza-C treatment induced significantly earlier flowering in both spring and winter wheats, but the response was pronounced for the winter wheat. The gamma rays effectively reduced flowering time in the winter wheat only. Recently, the hypothesis that the promotion of flowering by vemalization is mediated by DNA demethylation was tested by transforming Arabidopsis plants with a methyltransferase (MET1) antisense gene (Finnegan et al., 1998). The authors found that plants homozygous for ME T1 had 15% of normal methylation levels and flowered significantly earlier than untransformed controls. They also found that the promotion of flowering generally was correlated with the extent of demethylation. They tested plants of three methyltransferase antisense families 24 that differed in DNA demethylation magnitude; plants homozygous for four copies of the antisense gene flowered significantly earlier than those homozygous for three. Demethylation caused by the MET1 antisense transgene did not prevent a response to vemalization. Cold treatment of antisense plants accelerated flowering compared to that of untreated antisense plants, which indicates that demethylation caused by the MET1 antisense transgene did not saturate the early flowering response and acted additively with a cold treatment. Studies of the inheritance of early flowering used plants with the ME T1 antisense transgene and showed that demethylation caused by the transgene did not exactly replicate early flowering due to vemalization (Finnegan et al., 1998). For wild type plants, the requirement for vemalization is reset every generation (possibly because of remethylation during gametogenesis or embryogenesis). Like those of wild type plants, progeny of cold-treated or untreated antisense plants flowered at the same time, which shows that the MET1 antisense transgene does not inhibit the resetting process. However, progeny of plants hemizygous for the transgene (and not given a cold treatment) flowered earlier than progeny of an untransformed plant. Even progeny that did not contain the antisense transgene had reduced levels of methylation and flowered earlier than progeny of untransformed plants. Although this response does suggest that early flowering is caused by lower levels of methylation and not the transgene itself, it also shows that early flowering due to demethylation by MET1 antisense does not replicate all aspects of vemalization. Finnegan et 25 al. (1998) suggested that there may be two processes involved if demethylation of DNA is necessary for vemalizationu one involved in the actual demethylation, the other involved in resetting the vemalization requirement each generation. Alternatively, the authors suggest that there may be several methyltransferases that methylate different sites and that the MET1 antisense transgene affects only one of these. Conclusion Although a great deal of information that characterizes the phenomenon of vemalization and the genetics involved has been generated, much is still unknown. The complex interactions between temperature before, during, and after the cold treatment, photoperiod, light intensity, and plant maturity make it a difficult subject to study. Regardless, a greater understanding of the genetic mechanisms for controlling developmental processes should shed light on the mystery of how cold temperatures are perceived and how this stimulus is translated into the ability to flower. 26 Literature Cited Bagnall, DJ. 1993. Light quality and vemalization interact in controlling late flowering in Arabidopsis ecotypes and mutants. Ann. Bot. 71: 75-83. Brock, RD. and J.L Davidson. 1994. 5-azacytidine and gamma rays partially substitute for cold treatment in vemalizing winter wheat. Env. Exp. Bot. 34: 195-199. Brooking, I. 1996. Temperature response of vemalization in wheat: a developmental analysis. Ann. Bot. 78: 507-512. Burn, J.E., D.J. Bagnall, J.D. Metzger, E.S. Dennis, and W.J. Peacock. 1993. DNA methylation, vemalization, and the initiation of flowering. Proc. Natl. Acad. Sci. USA90: 287-291. Chandler J., A. Wilson, and C. Dean. 1996. Arabidopsis mutants showing an altered response to vemalization. Plant J. 10: 637-644. Chong K., L. Wang, K. Tan, H. Huang, and H. Liang. 1994. Molecular cloning and characterization of vemalization-related (ver) genes in winter wheat. Physiol. Plant. 92: 511-515. Chong, K., S. Bao, T. Xu, K. Tan, T. Liang , J. Zeng, H. Huang, J. Xu, and Z. Xu. 1998. Functional analysis of the ver gene using antisense transgenic wheat. Physiol. Plant .102: 87-92. Chouard, P. 1960. Vemalization and its relation to dormancy. Annu. Rev. Plant Physiol. 11: 191-238. Dahanayake SR. and NW. Galwey. 1998. Effects of interactions between low and high temperature treatments on flowering of Spring Rape (Brassica napus var. annua). Ann. Bot. 81: 609-617. Fernandez J.A., S. Banon, J.A. Franco, A. Gonzalez, and PF. Martinez. 1997. Effects of vemalization and carbohydrate levels in the apex of cauliflower (Brassica oleracea var. botrytis) Scientia Hort. 70: 223-230. Finnegan, E.J., R.K. Genger, K. Kovac, W.J. Peacock, and ES. Dennis. 1998. DNA methylation and the promotion of flowering by vemalization. Proc. Natl. Acad. Sci. 95: 5824-5829. 27 Hazebroek, J.P., J.D. Metzger, and ER. Mansager. 1993. Thennoinductive regulation of gibberellin metabolism in Thlaspi arvense L. Plant Physiol. 102: 547-552. Lang, A. 1965. Physiology of flower initiation. In W Ruhland [ed] Encyclopedia of Plant Physiology, Springer-Verlag, 15 (1) 1380-1536. Lee, I., A. Bleecker, and RM. Amasino. 1993. Analysis of naturally occurring late flowering in Arabidopsis thaliana Mol. Gen. Genet. 237: 171- 176. Lee, l., S.D. Michaels, A.S. Masshardt, and RM. Amasino 1994. The late- flowering phenotype of FRIGIDA and mutations in LUMINADEPENDENS is suppressed in the Landsberg erecta strain of Arabidopsis. Plant J. 903- 909. Lee, I. and RM. Amasino 1995. Effect of Vemalization, photoperiod, and light quality on the flowering phenotype of Arabidopsis plants containing the FRIGIDA gene. Plant Physiol. 108: 157-162. Mathon, CC. 1960. Nouvelles recherches sur la vemalisation et le photoperiodisme Chez Ies Campanulacees. Phyton 14(2):167-174. Metzger, JD. 1988. Localization of the site of perception of therrnoinductive temperatures in Thlaspi arvense L. Plant Physiol. 88: 424-428. Murfet, LC. and J.B. Reid. 1974. Flowering in Pisum: The influence of photoperiod and vemalising temperatures on the expression of genes Lf and Sn. 2. Pflanzenphysiol. 71: 323-331. Napp-Zinn, K. 1961. Vemalisation und verwandte Erscheinungen, p. 24-75. In: W. Ruhland (ed.). Encyclopedia of Plant Physiology. Springer-Verlag, Berlin. Napp-Zinn, K. 1984. Light and Vemalization, p.75-88. In: D. Vince-Prue, B. Thomas, K.E. Cockshull (eds.). Light and the Flowering Process. Academic Press, London. Napp-Zinn, K. 1987. Vernalization—environmental and genetic regulation, p.123- 132. In: J.G. Atherton (ed.). Manipulation of Flowering. ButtenNorths, London. Nishijima, T., N. Katsura, M. Koshioka, H. Yamazaki, and L. Mander. 1997. 28 Effects of uniconazole and GA3 on cold-induced stem elongation and flowering of Raphanus sativus L. Plant Growth Reg. 21: 207-214. Osborn, T.C., C. Kole, I.A.P. Parkin, A.G. Sharpe, M. Kuiper, D.J. Lydiate, and M. Trick. 1997. Comparison of flowering time genes in Brassica rapa, B. napus and Arabidopsis thaliana. Genetics 146: 1123-1129. Picard, C. 1965. Contribution a la connaissance de la vemalisation, de ses particularites et de sa signification Chez Oenothera biennis L. var. Sulfurea De Vries. Ann. Sci. Naturelles Bot. 6 (12): 197-314. Purvis, ON. 1961. The physiological analysis of vemalization, p. 76-122. In: W. Ruhland (ed.). Encyclopedia of Plant Physiology. Springer-Verlag, Berlin. Rawson, H.M., M. Zajac, and L.D.J. Penrose. 1998. Effect of seedling temperature and its duration on development of wheat cultivars differing in vemalization response. Field Crops Res. 57: 289-300. Reid, J.B. and LC. Murfet. 1975. Flowering in Pisum: The sites and possible mechanisms of the vemalization response. J. Exp. Bot. 26: 860-867. Roberts, E.H., R.J. Summerfield, J.P. Cooper, and RH. Ellis. 1988. Environmental control of flowering in barley (Hordeum vulgare L.): l. Photoperiod limits to long-day responses, photoperiod -insensitive phases and effects of low-temperature and short-day vemalization. Ann. Bot. 62:127-144. Roberts, E.H. and R.J. Summerfield. 1987. Measurement and prediction of flowring in annual crops, p. 17-50. In: J.G. Atherton (ed.). Manipulation of Flowering. Butterworths, London. Robertson, M.J., I.R. Brooking, and J.T. Ritchie. 1996. Temperature response of vemalization in wheat: modelling the effect on the final number of mainstem leaves Ann. Bot. 78: 371-381. Rood, 8.8., D. Pearce, P.H. Williams, and RP. Pharis. 1989. A gibberellin- deflcient Brassica mutant-rosette. Plant Physiol. 89: 482-487. Runkle, E.S., R.D. Heins, A.C. Cameron, and W.H. Carlson. 1998. Flowering of Leucanthemum xsuperbum 'Snowcap' in response to photoperiod and cold treatment. HortScience 33(6): 1003-1006. 29 Schwabe, W.W. 1955. Devemalization in relation to high temperature and low light intensity treatments. J. Exp. Bot. 6: 435-450. Schwabe, WW 1957. Devemalization by low light intensity in relation to temperature and carbohydrate supply. J. Exp. Bot. 8: 220-234. Teutonico, R.A., T.C. Osborn. 1995. Mapping loci controlling vemalization requirement in Brassica rapa Theor. Appl. Genet . 91: 1279-1283. Thomas, B. and D. Vince-Prue 1984. Juvenility, photoperiodism, and vemalization, p.408-439. In M.B. Wilkins (ed.). Advanced Plant Physiol. Pitman, London. Vince-Prue, D. 1975. Vemalization, p.263-291. Photoperiodism in Plants. McGraw —Hill, London. Wang, S.Y., R.W. Ward, J.T. Ritchie, R.A. Fischer, and U. Schulthess. 1995. Vemalization in Wheat: I. A model based on the interchangeability of plant age and vemalization duration. Field Crops Res. 41: 91-100. Weller, TC and R.W. Langhans. 1968. Determination of vemalizing temperatures in the vemalization requirement of LiIium Iongiflorum (T hunb.) cv 'Ace'. J. Amer. Soc. Hort. Sci. 93: 623-629. Wellensiek, S.J . 1960. Flower formation in Campanula medium. Meded. Landbouwhogeschool Wageningen 60: 1-18. Wellensiek, SJ. 1964. Dividing cells as the prerequisite for vemalization. Plant Physiol. 39: 832-835. Whitman, C.M., R.D. Heins, A.C. Cameron, W.H. Carlson. 1996. Cold treatments, photoperiod, and forcing temperature influence flowering of Lavandula angustifolia. HortScience 31 (7): 1150-1153. Wittwer, S.H., M.J. Bukovac. 1962. Quantitative and qualitative differences in plant response to the gibberellins. Amer. J. Bot. 49: 524-529. Whyte, RD. 1948. History of research in vemalization, p.1-38. In: A.E. Mumeek and RC. Whyte (eds.). Vemalization and Photoperiodism: A symposium, Chronica Botanica, Waltham, MA. Yeh, D.M., J.G. Atherton, and J. Craigon. 1997a. Manipulation of flowering in 30 cineraria: lIl. Cardinal temperatures and thermal times for vemalization. J. Hort. Sci. 72: 379-387. Yeh, D.M., J.G. Atherton, J. Craigon. 1997b. Manipulation of flowering in cineraria: IV. Devemalization. J. Hort. Sci. 72: 545-551. Zanewich, K.P. and SB. Rood. 1995. Vemalization and gibberellin physiology of winter canola. Plant Physiol. 108: 615—621. Zeevaart, J.A.D. 1978. Phytohormones and flower formation, p.291- 327. In: 0.8. Letham, P.B. Goodwin, and T.J.V. Higgins (eds.). Phytohormones and related compounds—A comprehensive treatise, vol.2. EIsevier/North Holland Biomedical Press, Amsterdam. 31 Growth and Development of Oenothera fruticosa Is Influenced by Cold Treatment, Photoperiod, Forcing Temperature, and Plant Growth Regulators 32 Abstract The influences of cold treatment, photoperiod, forcing temperature, and plant growth regulators on growth and development of Oenothera fruticosa L. ‘Youngii-Iapsley’ were determined. Plants were cooled at 5 °C for 0, 3, 6, 9, 12, or 15 weeks under a 9-h photoperiod and subsequently forced in a 20 °C greenhouse under a 16-h photoperiod. Only one plant in two years flowered without cold treatment. Time to visible bud and flower decreased about 1 week as cold duration increased from 3 to 15 weeks. Response of plants to 10-, 12-, 13-, 14-, 16-, or 24-h photoperiods or a 4-h night interruption (NI) were determined in a 20 °C greenhouse following 0 or 15 weeks of cold treatment at 5 °C. Flowering occurred under all photoperiods only after cold treatment. Time to flower decreased about 2 weeks, flower number decreased, and plant height increased as photoperiod increased from 10 to 24 h. Days to flower under NI were similar to that of plants grown under a 16- h photoperiod. To determine the effect of forcing temperature on flowering of ‘Youngii-lapsley', plants were cooled for 15 weeks and then forced under a 16-h photoperiod at 14, 17, 20, 23, 26, or 29 °C. Plants flowered more quickly at higher temperatures but were taller and produced fewer and much smaller flowers. Days to visible bud and flowering were converted to rates, and base temperature (Tb) and thermal time to flowering (degree-days) were calculated as 4.4 °C and 606 °day. The effects of spraying ancymidol, Chlorrnequat, paclobutrazol, daminozide, and uniconazole were tested on plants cooled for 19 weeks and forced at 20 °C under a 16-h 33 photoperiod. Final plant height was reduced 31% and flower diameter was reduced 36% for plants sprayed with uniconazole. Plants sprayed with all other plant growth regulators were similar to control plants. 34 Introduction Many species of Oenothera have been cultivated since the early 18008 when they were brought from North and South America to England where they became popular garden plants (Wiesner,1991). Oenothera fruticosa L., or sundrops, is native to eastern North America and hardy in USDA zones 4-8. Oenothera fmticosa is a day-flowering species that typically flowers in June (Griffiths, 1994; Nau, 1996). In a preliminary screen, the vegetatively propagated cultivar Youngii-Iapsley showed great ornamental flowering potted-plant potential with its long natural flowering time (=4 weeks) and numerous large bright-yellow flowers (Frane, 1999). To our knowledge, no species of Oenothera is grown currently as a pot plant; for potted plant production, detailed information on flowering requirements is needed. Vemalization and photoperiod are two important factors regulating flowering of vegetative propagules. Vemalization can be defined as the promotion of flower formation by a period of low temperatures, generally between -5 and 15 °C, with a broad optimum between 1 and 7 °C (Lang, 1965). A vemalization requirement often is linked to a specific photoperiod requirement, which for many species is usually, but not always, a long-day requirement (Vince-Prue, 1975). Flowering of many herbaceous perennials is influenced by vemalization and photoperiod (lversen and Weiler, 1994; Whitman et al., 1996,1997; Runkle et al., 1998). 35 Many early experiments investigating the role and mechanism of vemalization and photoperiod in flowering were performed on various species of Oenothera and several different response types were identified. Picard (1965) observed flowering of O. biennis var. sulfurea De Vries only after a minimum cold treatment (at 11 °C /3 °C day/night [D/N]) of 10 weeks followed by long days 212 h. Other species of Oenothera, including 0. suaveolens, O. Iongiflora, and O. stricta, had facultative requirements for vemalization followed by an obligate requirement for long days (Chouard, 1960). Oenothera rosea L’Herit. ex Ait. is a facultative long-day plant with no response to vemalization (Chouard, 1960). In preliminary experiments, ‘Youngii-lapsley’ appeared to be a facultative long-day plant with an obligate vemalization requirement (Frane, 1999) Knowledge of the effect of temperature on plant development is important to accurately time crop production. Without photoperiod or vemalization effects, the rate of development increases linearly with temperature within a limited range of growing temperatures (Roberts and Summerfield, 1987). Such responses are also evident in photoperiod- and vemalization-sensitive species when the separate effects of photoperiod and vemalization are removed from those of postvemalization temperatures. The relationship between temperature and rate of development toward flowering (1/DTF where DTF is days to flower) can be described as follows: 1/DTF = bo + b,*T [1] 36 where b0 and b1 are constants and T is temperature (°C) (Roberts and Summerfield, 1987). The base temperature (T.,), or the temperature at or below which the rate of progress toward flowering is zero, and thermal time, or the number of units (°days) of thermal time above a base temperature, required for flowering can be calculated from the constants in Equation 1 as follows: T.,= -b,,/b, [2] °days=1lb1 [3] Growing temperature can affect plant appearance as well as flower timing. For Antinhinum majus L. ‘Jackpot’, there was an increase in stem length, spike length, and number of florets as temperature decreased from 21 to 10 °C (Maginnes and Langhans. 1961). Flower size decreases with increasing temperature in Galinsoga parviflora Cavan. and Agrostemma githago L. (Knapp, 1957); Campanula carpatica Jacq. ‘Blue Clips’ (Whitman et al., 1997); Gail/ardia xgrandiflora Van Houtte ‘Goblin’, Leucanthemum xsuperbum Bergman ex J. Ingram ‘Snowcap’, Coreopsis grandiflora Hogg ex Sweet ‘Sunray’ and Rudbeckia fulgida Ait. ‘Goldsturrn’ (Yuan et al., 1998). A potential problem in growing 0. fruticosa ’Youngii-lapsley' as a potted plant is its relatively tall natural height (33-54 cm). Control of height by plant growth regulator (PGR) treatment may be required to produce high-quality potted plants. Although we have been unable to find literature on the effect of PGRs on O. fmticosa, Chlorrnequat (up to 3.0 pg) applications to the root zone after 37 chilling had no effect on stem elongation or flowering of O. biennis (Picard, 1967). According to Picard (1967), chlormequat in doses between 50 and 650 pg applied to the apical meristem actually promoted stem elongation, while a dose of 750 #9 decreased it. Daminozide in doses up to 550 pg applied to the root zone after chilling decreased stem elongation and delayed flowering, the effect increasing with larger doses. A dose of 650 ,ug of daminozide completely inhibited flowering (Picard, 1967). The objectives of our study were (1 ) to determine and quantify the effects of cold treatment, photoperiod, and forcing temperature on flowering of O. fruticosa ‘Youngii-Iapsley’, and (2) to determine the effect of five commercially available PGRs on flowering and final plant height of ‘Youngii-lapsley’. Materials and Methods General. Plants of Oenothera fmticosa ‘Youngii-Iapsley’ were received as rooted stern cuttings in 72-cell plug trays on 29 Oct. 1997 (Year 1) and 23 Sept. 1998 (Year 2) from the same commercial producer. Upon arrival at Michigan State University, all plants were held in plug trays and exposed to natural day lengths (lat. 43 °N;z11-12 h) for 2 or 5 days (Year 1 or 2, respectively) until the experiments began. Cold treatments were delivered to plants held in plug trays in coolers set at 5 °C where plants were illuminated for 9 h-dayr1 with cool-white fluorescent lamps (F96T12/CWNHO, Philips, Somerset, NJ.) at approximately 10 umoI-mZ-st and watered as necessary (approximately 2 times per week) using well water acidified with sulfuric acid (HZSO,) to a pH of approximately 6.0. 38 Before placement in the greenhouses (set at 20 °C day/night [D/N] unless otherwise stated), plugs were transplanted into 13-cm square plastic pots (1 .1-L volume) with a soilless medium containing sphagnum peat moss, composted pine bark, vermiculite, and perlite (Strong-lite High Porosity Mix, Strong-Lite, Pine Bluff, Ark.). Pots were spaced at approximately 22 plants/m2. Plants were fertilized at every irrigation with a nutrient solution of well water (EC of 0.70 mS-cm‘1 and 105, 35, and 23 mg-L'1 Ca, Mg, and S, respectively) acidified with H2SO, to a titratable alkalinity of 130 mg-L'1 CaCO3 and water soluble fertilizer providing 125-12-125-13 N-P-K-Ca mg-L'1 (30% ammoniacal N) plus 1.0-0.5-0.5- 0.5-0.1-0.1 mg-L'1 (Fe, Mn, Zn, Cu, B, Mo) (MSU Special, Greencare, Chicago, Ill.) The target range was 0.5-1.2 mS-cm’1 for electrical conductivity (EC) and 5.8-6.2 for pH. In the second year, it was necessary to drench all plants transplanted after 1 Jan. 1999 twice with a 750 mg N-L'1 liquid fertilizer and K003 to raise the pH and EC into the desired ranges. Cold duration (Expt. 1). Six durations of cold treatment were tested (0, 3, 6, 9, 12, or 15 weeks). Experiments were initiated on 31 Oct. 1997 (Year 1) and 28 Sept. 1998 (Year 2). Initial node counts on aerial shoots were not taken in Year 1 for any experiment. In Year 2, the initial number of nodes was counted at transplant and averaged 18.1. A 16-h photoperiod was provided by four- hundred-watt high-pressure sodium lamps (HPS) switched on at 0800 HR and off at 2400 HR. HPS lamps provided both photoperiod and supplemental lighting with a photosynthetic photon flux (PPF) of 50 umol-mZ-s-‘ at plant height starting 39 at 0800 and continuing until the outside PPF exceeded 400 pmoI-mZ-s-L If the outside PPF then dropped below 200 pmoI-mZ-s-' lamps were again turned on until 2400 HR. Ten plants were removed from the cooler after each treatment, transplanted, and placed in the greenhouse. Photoperiod (Expt. 2). Seven photoperiods (10, 12, 13, 14, 16, and 24 h and a 4-h NI) and two cold treatments (0 or 15 weeks) were tested. Seventy plants were transplanted on 2 Nov. 1997 (Year 1) and 28 Sept. 1998 (Year 2) and placed in a greenhouse. Ten plants were used per treatment and treatments were assigned randomly each year to greenhouse benches. The remaining plants were left in the plug tray and placed in the cooler. After 15 weeks of cold treatment, 70 plants were transplanted on 18 Feb. 1998 (Year 1) and 12 Jan. 1999 (Year 2) and placed in the same photoperiod treatments as the first set of 70 plants. In Year 2, plants averaged 18.0 nodes. Opaque black cloth was pulled at 1700 HR and opened at 0800 HR every day on all benches so all plants received a similar daily light integral. Continual photoperiods were delivered by day-extension lighting with incandescent lamps that were turned on at 1700 HR and turned off after each photoperiod was completed. The 4-h NI was provided by incandescent lamps that were turned on at 2200 HR and turned off at 0200 HR. Day-extension and NI lighting with incandescent lamps provided 1 to 3 ,umol-mZ-s' at canopy level. Supplemental lighting was provided by HPS lamps (which delivered approximately 50 4O pmoI-mZ-s-‘ at canopy level) between 0800 HR and 1700 HR as described for Expt.1. Forcing temperature (Expt. 3). On 28 Sept. 1998, plants left in the plug tray were placed directly into the cooler for 15 weeks until 12 Jan. 1999, when they were transplanted. At transplant, plants averaged 19.4 leaves. Ten plants were placed into each of six greenhouses set to 14, 17, 20, 23, 26, or 29 °C. Plants received a continual 16-h photoperiod provided by HPS lamps that were on from 0800 HR to 2400 HR and delivered approximately 90 umol-mZ-s-‘ at plant height. Plant growth regulators (Expt. 4). Five commercially available plant growth regulators (PGRs) were tested for their ability to control final plant height of ‘Youngii-Iapsley’. The PGRs and rates tested included daminozide (butanedioic acid mono-[2,2-dimethyl hydrazidel) at 5000 mg-L", paclobutrazol ([2R,3R+28,38]-1-[4-chlorophenyI-4,4-dimethyl-2-[1,2,4-triazoI-1-yl]]) at 30 mg-L' ‘, chorrnequat ([2-chloroethyl] trimethylammonium chloride) at 1500 mg-L", uniconazole ([E]-[S]-1-[4-chlorophenyl]-4,4-dimethyl-2-[1,2,4-triazoI-1-yI]-pent-1- ene-3-ol) at 15 mg-L'1 and ancymidol (or-cyclopropyl-or-[4-methoxyphenyl]-5- pyrimidinemethanol) at 100 mgL‘. Plants were given a 19-week cold treatment at 5 °C and on 6 Feb. 1999 were transplanted and placed in a greenhouse set at 20 °C. A 16-h photoperiod was provided by HPS lamps, as described for Expt. 1. The daily average temperature between transplant and average date of flowering was 20.6 °C and the average daily light integral was 14.9 mol'm2~day‘. Plants averaged 18.2 nodes at the beginning of the experiment and were 41 allowed to establish for 10 days before the first PGR spray. Eight plants were sprayed with each of the chemicals until the solution dripped off the leaves (approximately 3 mI'pIant-I). Following the first spray, plants were sprayed every 10 days until the first flower opened. In total, there were three applications per PGR. Eight plants were not sprayed with any chemical and served as the control. Preliminary postharvest evaluation. The 10 plants in the 15-week cold treatment of Expt. 1 (Year 1) were moved to an environmentally controlled chamber on 3 Apr. 1998, approximately 1 week after the first flower opened. The average chamber temperature (measured with a thermometer in a water bath) was 24.6 °C. Plants were given a 24-h photoperiod delivered with both cool-white fluorescent and incandescent lamps at approximately 90 ,umOI'ITIZ'S'1 (7.8 moI-mz-day‘), watered as needed (usually every day), and were monitored for 12 days for the number of open, flowers, abscised, and spent flowers. Number of new open flowers was calculated by subtracting the number of open flowers from the previous day from the number of flowers open on the current day and adding the current day’s number of spent and abscised flowers. Data collection and analysis. The data collected for all other experiments (Expts. 1-4) included date of first visible flower bud (without dissection) and date of first open flower. Days to visible bud, days to flower, and days from visible bud to flower were calculated for all experiments. At first open flower, total plant height (excluding the container), number of visible flowers and buds, and number of 42 final nodes on the main stem were recorded. In Year 1, the number of final nodes was counted to the last visible leaf. In Year 2, the number of final nodes was counted to the last visible leaf and included the initial leaf count, the number of new nodes produced during forcing was calculated by subtracting the initial number of leaves from the final number of leaves, and the number of new nodes beneath the first flowering lateral was calculated by subtracting the number of initial nodes from the number of nodes beneath the first flowering lateral. The number of new nodes beneath the first flowering lateral was calculated only for Expts. 2 and 3. Number of lateral shoots and number of flowering lateral shoots were recorded for Year 2 only. Flower diameter was recorded for Expts. 3 and 4. For all experiments, plants that did not have visible buds after 15 weeks of forcing were considered non-flowering and were discarded. Plants that died during the experiment were discarded and not included in the results or calculations. All experiments were completely randomized designs. Data were analyzed using the general linear models procedure (PROC GLM)(SAS Institute, Cary, NC.) for analysis of variance and linear regression procedure (PROC REG) for regression models. Duncan’s mean separation test was used to analyze the PGR experiment (Expt. 4) because of the discontinuous nature of the treatments. All other experiments were analyzed with linear and quadratic trend contrasts. The cold duration and photoperiod experiments (Expts. 1 and 2, respectively) were replicated in time, and each experiment was conducted once 43 per year. The forcing temperature and PGR experiments (Expts. 3 and 4, respectively) were each conducted only once in Year 2. Greenhouse temperature control and daily light integral collection. Air temperature was monitored in each greenhouse with 36-gauge (0.127-mm- diameter) thermocouples connected to a CR 10 datalogger (Campbell Scientific, Logan, Utah). The datalogger collected temperature data every 10 s and recorded the hourly average. Average daily air temperatures from transplant to average date of flowering were calculated for every treatment in each experiment and those for Expts. 1 and 2 are shown in Tables 1 and 2. Average daily temperatures for Expt. 3 are shown in Table 3. Data for Expt. 4 are recorded above. Temperatures in each greenhouse were controlled by a Priva environmental computer (model CD750; Priva, De Lier, The Netherlands). Actual average daily air temperatures were calculated and used in all analyses. Average daily light integrals were collected at the same interval as temperatures at canopy level by using a quantum sensor (LI-COR) connected to the CR 10 datalogger. Daily light integrals were only collected for Expt. 1 in Year 2. Results and Discussion Cold duration (Expt.1). Without a cold treatment, only one out of 10 plants flowered in Year 1 and no plants flowered in Year 2 (Figure 1A). This plant is the only one that has flowered without cold in any of our experiments (1 out of 180 plants tested) and its flowering behavior was quite different from all other plants 44 examined. This plant flowered 42 to 54 days slower (Figure 1B), produced 47 to 55 more leaves (data not shown), was 23 to 34 cm shorter (Figure 3D) and produced about 25% the number of flowers that other plants did for Year 1 (Figure 3A). Because flowering without cold treatment was extremely rare, it was considered an anomaly and was not included in the analysis. We conclude that ‘Youngii-lapsley' can be classified as having an obligate requirement for vemalization with the acknowledgment that the system is “leaky” and on occasion a few plants will flower without cold treatment. In the first year, days to visible bud decreased quadratically (P g 0.001) and days to flower decreased linearly (P 5 0.001) as cold duration increased from 3 to 15 weeks (Figures1B and 1C). In the second year, days to visible bud decreased linearly (P s 0.001) and days to flower decreased linearly (P 5 0.001) as cold duration increased from 3 to 15 weeks (Figures1 B and C). While statistically significant, the decrease in time to flower was slight, 7 to 10 days. While there were significant year (P s 0.01), cold duration, and year X cold duration interaction (P 5 0.001) effects on days from visible bud to flower, the differences were small (Figure 1D); the largest difference in either year was four. Flowering of ‘Youngii-Iapsley’ was very uniform. While there was a significant year effect (P 5 0.001) on days to visible bud, days to flower, and days from visible bud to flower (P s 0.05), these differences were relatively small. For example, excluding the single uncooled plant that flowered in Year 1, the 45 greatest difference between the average of any of these variables in either year was 7 days; the smallest, zero. Within each cold-duration treatment for either year, all plants flowered within 3 to 9 days. In either year, there were no significant trends for the number of final nodes measured to the last visible leaf (data not shown). In Year 2, the number of new nodes produced during forcing decreased linearly (P 3 0.001) from 36 to 28 nodes (Figure 2). For time to flower, the slight change in time to flower that occurred with increasing durations of cold treatment was only slightly correlated (r2: 0.37) with the decrease in the number of new nodes produced during forcing (Year 2). In Year 1, the number of flowers and buds counted at initial flower opening increased quadratically (P 5 0.001) with increased cold duration (Figure 3A). In Year 2 the trend was also quadratic (P s 0.05) but flower number was similar for plants cooled for 3 weeks or 15 weeks and highest flower counts were observed on plants cooled for 6 or 12 weeks (Figure 3A). Thrips infested plants cooled for 6 or 9 weeks in Year 2, which could have accounted for some of the decreased number of flowers of plants cooled for nine weeks. Plants cold- treated for 15 weeks in Year 1 were forced during February and March, while plants cooled for 15 weeks in Year 2 were forced during January and Febnrary, and so the higher flower number seen in Year 1 could be due to naturally higher light levels during forcing. Increasing light intensity has been shown to increase number of flowers. For Dendranthema xgrandiflorum Ramat. grown under a 12- 46 h day, flower number increased from 7.2 to 19.0 as daily light integral increased from 10.4 to 29.9 mol-mZ-day‘ (Warrington and Norton, 1991 ). A count of the number of flowering lateral shoots and total (vegetative and reproductive) lateral shoots in Year 2 showed that the former increased linearly (P s 0.01) with increased durations of cold treatment (Figure 3B) while the latter was similar for all durations of cold treatment (data not shown). The percentage of flowering laterals increased from 69 to 78 as cold increased from 3 to 15 weeks in Year 2 (Figure 30). Plants in Year 2 averaged 11 cm shorter over all cold duration treatments than those grown in Year 1 (Figure 3D). In both years there was a significant (P 5 0.001) quadratic trend for the influence of cold duration on final plant height. Photoperiod (Expt. 2). Without a cold treatment, no plants flowered under any photoperiod in either year. After 15 weeks at 5 °C, all plants flowered under all photoperiods. There was no significant year effect on days to visible bud. However, there was a significant (P 5. 0.001) year effect on days to flower as well as a significant (P 5 0.001) year X photoperiod interaction. The differences between days to flower over the two years were small; the greatest difference between the average was six days under the 10-h photoperiod. Days to flower and days to visible bud decreased quadratically (P 5 0.001) by 14 (Year 1) or 19 (Year 2) as photoperiod increased from 10 to 24 h (Figure 4A and C). Under 4-h NI, days to flower were 40.4 in both years. Days to flower under NI were similar 47 to that of plants grown under photoperiods of 16 h in both years and 24 h in Year 1. There was a significant (P g 0.001) year effect as well as a significant (P 5 0.001) year X photoperiod interaction for days from visible bud to flower. In Year 1, days from visible bud to flower decreased linearly (P 5 0.001) with increasing photoperiod (Figure 4B). However, in Year 2, the trend was quadratic (P 5 0.001). Actual differences were relatively small (< 7 days). From these results, we conclude that ‘Youngii-Iapsley’ is a facultative long-day plant since it flowered under all tested photoperiods, but flowering was accelerated by two weeks under the longer photoperiods. The fact that no plants flowered without cold in either year supports our conclusion that ‘Youngii-lapsley’ has a nearly obligate vemalization requirement. The final number of nodes produced under each photoperiod did in general reflect the change in flower timing. In Year 1, the final node number decreased linearly (P 5 0.001) from 38 to 30 under photoperiods $14 to 24 h, respectively (Figure 5A). However, in Year 2, the Change in final node numbers was quadratic (P s 0.05; Figure 5A). Final number of nodes in Year 2 decreased from 59 to 47 as photoperiod increased from 10 to 24 h. The number of new nodes produced during forcing in Year 2 (Figure 5B) decreased quadratically (P s 0.001) from 42 to 29 as photoperiod increased from 10 to 24 h. There were no significant trends for the number of new nodes produced beneath 48 the first flowering lateral (data not shown). Under NI, number of nodes to the final leaf was similar to that of plants grown under 16 or 24 h for either year. The number of flowers and buds at first flower decreased markedly (quadratically; P g 0.001) with increasing photoperiod (Figure 6A). For example, in Year 2, the average number of flowers decreased from 129 to 36 as photoperiod increased from 10 to 24 h. There was a significant (P 5 0.001) year effect but the year X photoperiod interaction was not significant. In general, plants grown in Year 2 produced slightly more flowers. The number of flowers appears to be influenced strongly by light levels. For example, plants cooled for 15 weeks and grown under a 16-h photoperiod in Expt. 1 provided by day extension with high-pressure sodium lamps produced 97 flowers while plants in Expt. 2 also cooled for 15 weeks and provided a 16—h photoperiod with day extension from incandescent lamps produced 47 flowers (Year 2). It was estimated that day extension with incandescent lamps reduced the daily light integral by 22 % compared to day extension with high-pressure sodium lamps. This estimation was found by subtracting the amount of light available during daylight hours (0800 to 1700 HR) (as measured by a LI-COR quantum sensor placed on the bench at canopy height) from the amount of light available during the full 16-h photoperiod of day extension with high-pressure sodium lamps, and assuming that incandescent lamps (with an output of 1-3 umoI-mz-seC") contribute essentially no light beyond the compensation point. 49 The number of flowering lateral shoots and total lateral shoots decreased quadratically (P s 0.01 for the number of flowering lateral shoots; P s 0.05 for the number of total vegetative shoots) with increasing photoperiod, while the percentage flowering lateral shoots changed little (Figure 6B and C). The change in the number of lateral shoots under the different photoperiods had a marked effect on plant appearance. Plants grown under photoperiods 514 h were very leafy and floriferous while those grown under photoperiods 214 h appeared spindly with a sparse flower display. Plants grown under NI had as many total and flowering lateral shoots as plants grown under 16 h. There was a small but significant (P s 0.05) year and photoperiod (P 3 0.001) effect on final plant height but the year X photoperiod interaction was not significant (Figure 6E). In Year 1, there were no significant trends for the effect of photoperiod on plant height. In Year 2, plant height increased quadratically (P g 0.01); differences in height were small, increasing from 42 to 48 cm as photoperiod increased from 10 to 24 h. Under NI, height was similar to that of plants grown under 24- and 16- h photoperiods for both years and a 14- h photoperiod for Year 1. Forcing temperature (Expt. 3). All plants flowered except for one plant in the 14 °C treatment. Actual temperatures for each greenhouse section during different developmental stages are shown in Table 3. Days to visible bud, days from visible bud to flower, and days to flower decreased as temperature increased from =15 to :30 °C (Figure 7A-C). Increasing temperature from 15.2 to 20.6 °C 50 accelerated flowering more than increasing it from 20.6 to 29.8 °C. For example, days to flower decreased 24 d from 60 to 36 as temperature increased 5.4 °C from 15.2 to 20.6 °C, but only 11 d from 36 to 25 as temperatures increased 9.2 °C from 20.6 to 29.8 °C. Plants grown at 20.6 °C flowered approximately 5 days faster than plants cooled for 15 weeks in Expt. 1 (Year 2) which were subsequently grown at 20.8 °C (both under a 16-h photoperiod). A possible explanation for this difference is that the plants in Expt. 3 were provided approximately 40 ,umoI-mz-S'1 more supplemental light than plants in Expt.1. Such an increase in supplemental lighting using high-pressure sodium lamps probably would increase plant temperature, potentially as much as 0.5 °C (Faust and Heins, 1997). If true, then these plants would be expected to develop more quickly, since temperature strongly influences plant development. The number of final nodes were not significant when counted to the last visible leaf. However, when the new nodes produced during forcing were counted, there was a significant (Ps0.001) linear trend (data not shown). The number of new nodes decreased from 36 to 24 as temperature increased from 15.2 to 29.8 °C. Flowering of ‘Youngii-Iapsley’ was very uniform in this experiment, although variability increased as temperature decreased. All plants grown at 29.8 °C flowered within 3 days of each other; at 15.2 °C, within 8 days. There were significant linear relationships between temperature and rate of progress toward visible bud, from visible bud to flowering, and flowering 51 (Figure 7D-F). The parameters for the equations from the linear regression analysis are given in Table 4. The reciprocal of the linear regression function described the relationship between temperature and developmental rate well (Figure 7A-C). Base temperatures for each developmental stage ranged from 1.5 to 7.3 °C (Table 4). The degree-days to flower were 606 when days-to- flower data were used alone and 619 when data for days from transplant to visible bud and days from visible bud to flower were added. The concepts of degree-days and base temperature (T.,) are useful in commercial greenhouse production because they allow prediction of flowering time under fluctuating temperature conditions when average temperature is known (Roberts and Summerfield, 1987). By using the average daily temperature (T ,), the days required to complete a developmental stage can be calculated as °days/(‘l'a - T.,). Differences between predicted and observed days for any developmental stage and temperature were no greater than 5 (data not shown). The number of flowers decreased =84% as temperature increased from 15.2 to 29.8 °C (Figure 8A). The relationship between number of flowers and temperature was linear (P < 0.001). Flower diameter decreased from 7.0 to 4.0 cm as temperature increased, which corresponds to a threefold difference in flower area (assuming that flower shape is circular) (Figure BB). There are similar effects of temperature on flower size of other herbaceous perennials such as Coreopsis grandiflora (Hogg ex Sweet.) ‘Sunray’, Leucanthemum xsuperbum (Bergman ex J. Ingram) ‘Snowcap’, Rudbeckia fulgida (Ait.) ‘Goldsturrn’ (Yuan 52 et al., 1998), Campanula carpatica ‘Blue Clips’ (Whitman et al., 1997), Coreopsis verticillata L. ‘Moonbeam’, Campanula “Birch Hybrid’ and others (Frane, 1999). Height decreased 35% in a quadratic fashion (P < 0.001) as temperature increased from 15.2 to 23.8 °C (Figure 8C). As forcing temperature increased from 23.8 to 29.8 °C, there was almost no change in plant height. Plant growth regulators (Expt.4). There were no statistical differences among the growth regulators or the control for days to visible bud, number of new leaves, or number of flowers (data not shown). Days to flower ranged from 39 to 41 (Figure 9A). Plants sprayed with ancymidol, paclobutrazol, or chlormequat flowered approximately one day earlier than the control. Days to flower were similar for control plants and plants sprayed with daminozide or uniconazole. Days from visible bud to flower were approximately 20 for non-sprayed plants and plants sprayed with daminozide, uniconazole, or chlormequat and 19 for plants sprayed with ancymidol or paclobutrazol (data not shown). Final plant height was reduced 31 % for plants sprayed with uniconazole compared to that of the control plants (Figure 9B). No other PGR had a significant effect on final plant height although higher rates of daminozide and paclobutrazol may have reduced height significantly. Uniconazole reduced flower diameter 36% compared to that of control plants (Fig QC). The flowers on plants sprayed with uniconazole were noticeably smaller: the 36% decrease in diameter resulted in a 58% decrease in flower area. Flower diameters were similar for all plants treated with one of the other PGRs and for the control. 53 While daminozide and chlormequat did not reduce final plant height, plants sprayed with these PGRs as well as uniconazole had shorter lateral stems compared to that of control plants (personal observation). Shorter lateral stems changed the plants from their natural conical shape to cylindrical. If a standard height/diameter ratio of 1.5 to 2 (height = pot plus plant; diameter = diameter at top of plant or at greatest width) suggested for a high- quality potted plant (Sachs et al., 1976) was used, spraying uniconazole at 15 mg-L'1 was only slightly successful for reducing height. The 30-cm height achieved for plants sprayed with uniconazole resulted in a height/diameter ratio of approximately 2.1 (pot height = 11 cm and diameter was approximated at 1.5 X width of pot [13 cm]). Control plants had a height/diameter ratio of 2.7. A disadvantage of using this PGR and rate is the large reduction in flower size. Further studies using different rates of uniconazole and perhaps other PGRs individually or in combination are needed to identify PGRs and rates that effectively control height without such a drastic reduction in flower size and change in plant shape. Preliminary postharvest evaluation. During 12 days of postharvest evaluation, ‘Youngii-lapsley’ maintained a good floral display. The number of open flowers reached a maximum of 53 on day 5 (Figure 10). After day 5, the number of open flowers decreased to an average of 17 on day 12. At the end of the study, plants were in adequate condition (personal observation). In another postharvest evaluation using plants of ‘Youngii—lapsley’ cooled for 6 weeks, plants had a maximum of 26 open flowers and maintained at least 20 open flowers for the first ten days of the study (data not shown), only a few more flowers than plants had at the end of the 12-day study. New flowers continued to open under low light, although at a decreasing rate (Figure 10). The number of spent flowers ranged from 4 to 21 but was relatively constant during the 12 days of the study, averaging approximately 10 per day. Over 30 flowers per plant abscised on day two; from day three onward, there were never fewer than 10 abscised flowers per plant each day. The relatively high number of flowers abscised could decrease the potted plant potential of ‘Youngii-Iapsley’ since they are unattractive and need to be removed during marketing and home display. Summam On the basis of these results, 0. fnrticosa ‘Youngii-Iapsley’ can be classified as a facultative long-day plant with an obligate vemalization requirement. This species of Oenothera is the first that we know of with this combination of vemalization and photoperiod requirements. Flowering occurred consistently after a minimum cold treatment of 3 weeks followed by any photoperiod 210 h. Increasing the duration of cold treatment up to 15 weeks resulted in slightly faster flowering and more flowering lateral shoots. Although time to flower is approximately 2 weeks faster under a 16-h or longer photoperiod than a 10-h photoperiod, flower number decreased and height increased as photoperiod increased from 10 to 24 h. Plants were most attractive when grown under photoperiods s 14 h 55 because of the high numbers of vegetative and flowering lateral shoots and thus were better suited as potted plants than those grown under longer photoperiods. Since this is a plant that responds strongly to light levels, as seen when the flower number doubled when the daily light integral doubled (Expt. 1 and 2, Year 2), supplemental lighting is necessary to produce high-quality potted plants during winter. Forcing temperatures between 17 and 20 °C are recommended for forcing plants to flower within 5-6 weeks from the start of transplant. However, since there is an increase in the number of flowers with a decrease in temperature, temperatures .<.20 °C would be best for potted plant production. Uniconazole at 15 mg-L'1 can be used to reduce final plant height =30% but also will decrease flower diameter by 36%. In preliminary postharvest trials, ‘Youngii- lapsley’ performed well indoors. A good floral display was maintained for the 12 days of the trial. 56 Literature Cited Chouard, P. 1960. Vemalization and its relation to dormancy. Annu. Rev. Plant Physiol. 11: 191-238. Faust, J.E. and RD. Heins. 1997. Quantifying the influence of high-pressure sodium lighting on shoot-tip temperature. Acta Hort. 418: 85-91. Frane, A.J. 1999. Temperature effects on timing and bud development of Coreopsis verticillata ‘Moonbeam’ and flower induction of long-day perennials under different night temperatures. MS Thesis, Michigan State Univ., East Lansing. Griffiths, M. 1994. Index of garden plants. Timber Press, Portland. Iversen, RR. and TC Weller. 1994. Strategies to force flowering of six herbaceous garden perennials. HortTechnology 4(1): 61-65. Knapp, R. 1957. Ober die Beeinflussung von Eigenschaften der Bliiten und Infloreszenzen durch die Temperatur. NatunNissenschaflen 44: 383. Lang, A. 1965. Physiology of flower initiation, p.1380-1535. In: W. Ruhland (ed.). Encyclopedia of plant physiology. Springer-Vedag, Berlin. Maginnes, EA. and R.W. Langhans. 1961. The effect of photoperiod and temperature on initiation and flowering of snapdragon (Antinhinum majus— variety Jackpot). J. Amer. Soc. Hort. Sci. 77: 600-607. Nau, J. 1996. Ball perennial manual: Propagation and production. Ball, Batavia, IL. Picard C. 1965. Contribution a la connaissance de la vemalisation, de ses particularites et de sa signfication Chez Oenothera biennis L. var. Sulfurea De Vries. Ann. Sci. Naturelles, Bot. 6 (12): 197-314. lbid., 1967. Action du CCC et du B 995 sur la mise a fleur d’une plante bisannuelle, I’Oenothera biennis. Planta 74: 302-312. Roberts, E.H. and R.J. Summerfield. 1987. Measurement and prediction of flowering in annual crops, p. 17-50. In: J.G. Atherton (ed.). Manipulation of flowering. Butterworths, London. 57 Runkle, E.S., R.D. Heins, A.C.Cameron, and W.H. Carlson. 1998. Flowering of Leucanthemum xsuperbum ‘Snowcap’ in response to photoperiod and cold treatment. HortScience 33(6): 1003-1006. Sachs, R.M., A.M. Kofranek, and WP. Hackett. 1976. Evaluating new pot plant species. Florists Rev. 159(4116): 35-36,80-84. Vince-Prue, D. 1975. Vemalization. Photoperiodism in plants. McGraw Hill, London. Warrington, I.J. and RA. Norton. 1991. An evaluation of plant growth and development under various daily quantum integrals. J. Amer. Hort. Sci. 1 16(3): 544-551. Whitman C.M., R.D. Heins, A.C . Cameron AC, and W.H. Carlson. 1996. Cold treatments, photoperiod, and forcing temperature influence flowering of Lavandula angustifolia. HortScience 31(7): 1150-1153. lbid., 1997. Cold treatment and forcing temperature influence flowering of Campanula carpatica ‘Blue Clips’. HortScience 32(5): 861-865. Wiesner, MB. 1991. Enchanted evenings: Get to know Oenothera and dusk may be your favorite time of day. Amer. Hort. 70: 16-21. Yuan, M., W.H. Carlson, R.D. Heins, and AC. Cameron. 1998. Effect of forcing temperature on time to flower of Coreopsis grandiflora, Gail/ardia xgrandiflora, Leucanthemum xsuperbum, and Rudbeckia fulgida. HortScience 33(4): 663-667. 58 .Ammcmmu 92:39 8: EoEoSmmms. e .223 0:8 “39626: 953 02 e 9: 0.0? 0.0.. N._._. o.N_. . mam—flame -- -- .. I e... - 83-53 fawn .~-E._oEv .9925 Em: >=mn omEo>< wdw mdN Vow mdN ... _.N .... mmmvrwmmv Now adv Na? Na? odN mm? wmmvrnmmr Gov @588 octau moLBmLano. Em omEo>< 8 3 e e m o as, 6. m a £85 8:98 28 Begum: 2.0.8 «a 5:823 conform .>o_mam_-__mc:o>. 639:3 865,260 .8 9526: Go Emu omEo>m 2 .5396: Go Emu EC: ......EmBE Em: 23c new moeameoqEQ Em >__mu omEm>< .F aim... 59 69950: 9.53 02 s .czv 832.35 59: :4 was 85863 .3 N New EN Row Row New mew o: 2 - - - - - - - o 8-82 EN EN 3m 3m Row Row EN 8 - - - - .. - s e 8.3% AOL 9:80.— mctau EBECQEE Em omEo>< ~_Z em or 3 9. NF 2. EV 8:892“. o. m a 382, La; notoaoaoca some Lone: .>m_mam_-__mc:o>. BREE 865860 Lo.— mccosoc Co Emu mmmuozw 9 23¢ch Co Emu EC: 35:23:52 Em 2:8 omEo>< .N min... 60 Table 3. Actual temperatures of houses for the time of forcing for three developmental stages: transplant to visible bud, visible bud to flower, and transplant to flower. Temperature = Developmental stage (days) setting (°C) Transplant to Visible bud to Transplant to visible bud flower flower 14 15.4 15.0 15.2 1 7 18.2 18.3 18.2 20 20.5 20.7 20.6 23 23.7 23.8 23.8 26 26.7 27.1 26.8 29 29.6 29.9 29.8 61 29888. .53. a .58 v a a 285853.... .Loto Emvcmum A ....Be 8e 3A 155 A $8.? .333 A em. .3. 21E 2 Eases: :33 :N S Ste A ENE. 38.? A 39$. 2:“. e m> :56 8... me tweed A 39$ hues.» A 38¢- m> 9 E395: A3 263 .5 page A. ea? 8.? 3 82m E 38.25 3558.28 Agency 3368mm“. can fit EBECaEQ omen $330.8 2 com: 995 32m new 58.95 ”50.894650? 88:3 EeEeeeo 5. 93”; 8:96... 9 E3859 25 95A. 2 m>v access... 9 use was; .EB 25 pass 2 E335: E9: mmmemoa be 2m. 9 EBECQEQ @598 9522 mazmcm c2399: Laos. Co BEDS—Ema .v cam... 62 1ool ' . T +' i +’ i E, 80- +Year1 4:: +Year2 8 60l Q) U) S c 40« GD 2 a) O. 20. A 0- A + 100. I 80. E ”‘ NS ,3 60. L0 .9 W (D 5’ 4o. , o 20. , B 0 4. g P 100. 3 so. ' .0 .92 .D :2, 604 > m ... 9 LO 0, 404 >~ 8 20. ... LQNS c '6 o A . ; 100i 0 c 9 80, U U .D 2 60. :9 .2? E 40. . . 9 L0 Year1 :% 20. o ' LO Year2 D 0 fl 1 fi‘ 1‘? A r O 3 6 9 12 15 Weeks at 5 °C Figure 1. Responses of Oenothera fruticosa ‘Youngii-lapsley’ to various durations of cold treatment at 5 °C for percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Year 1 (I) or Year 2 (I ). Error bars represent 95% confidence intervals. All symbols represent the mean of 9 or 10 plants except for Year 1, 0 week cold treatment, when only one plant flowered. L = linear; Q = quadratic trends. NS~"-"'l\lonsignificant or significant at Ps 0.01 or 0.001, respectively. 63 404 30- 20- 101 Number of new nodes 0 3 6 9 12 15 Weeks at 5 °C Figue2. Biedofmldduationat5°Connmberofra~mdesprodmedduing forcinginYearZ. ErrorbarsrqarwntQE/ooorfidemeirmnrals. L=linear; Q = qLadalic trends. "5'«“r"l\bnsigriflwrt cr sign'fimrt atPs 0.05, 0.01, a 0.001, respectively. 64 250 - r . T m 3 .D 200- t: : co {2 150- cu 3 :2 b 100- o $- 8 g 50- z ' A (I) l *5 28. o 4: m g A. .J E c: E 10 a) . 3 o c 3 51 8 '9 B E Z 0 . E .9100- t0 if: E 80‘ r ,__.’—/\‘ c» E so a) , , . _, .. , B o C: 40‘ a r: 0 § 20. ,_ 0‘ C O i t ‘A s + 60.. .. , .. ., Ns... L0 50. _ M v N5... E LO .9 30‘ ' E 201 . 10. D- 0 e 4 i t I ¢ 0 3 6 9 12 15 Weeks at 5 °C Figure 3. Effects of cold duration at 5 °C on number of flowers and buds (A), number of flowering lateral shoots (B), percent flowering lateral shoots (C), and plant height (D) for Oenothera fiuticosa ‘Youngii-lapsley in Year 1 ( I) or Year 2 (I). All symbols represent the mean of 9 or 10 plants except for Year 1, 0 week cold treatment, when only one plant flowered. Error bars represent 95% confidence intervals. L = linear, Q = quadratic trends. r‘5'"""""Nonsignilicant or significant at P s 0.05, 0.01 or 0.001, respectively. 65 654 so- 55- so- 45J 4o- Days to flower 35- \ \\ \\ \\ Days from visible bud to flower \\ \Y W \\ \‘ Days to visible bud Photoperiod (h) Figue4. Wofmfiwmsa‘YwigiHapsleymvanoispmtopaiods alter15vreelsofooldtreatmertat5°C Cortirual plntoperiodtrealrrerts oasistedon—hramraldaysexterdedvfimligttflomirmrdesoertlan'ps. N: 4-hrigttirternplion Ermrbarsrepecert95%oonfidemeirtervals.L=linear; Q=quadratictrerds. I“’3‘"v""""Norsigrificart:orsigrilimrtatPs 0.05,0.01,or 0.001,respeclively. 66 Y //Y A $3 60‘ +Year1 8 +Year2 g 50. L"Q' “6 I § 40. I 3 ... z I L QNS 0 . L . 3 27 T 57 50‘ A . .1 . . . .,,.B E +Year2 we 40. , . . 00 BB 5.5 30. . , i . _ o: ... ... j .013 . . . 88 20 73a 0 E. 10- 0 I + I t i t 4. 4.,é/s 8 10 12 14 16 18 20 22 24 Nl Photoperiod (h) FgreSETedsofvaiorspluopaiaisamr15mksat5°Cmnnberoffird nodes(A)andmnberofmdesprodmeddLnrgfordng(B)forOax1haa fnliocBa‘Yomgii-lapsley inYear1 (I) crYea2(-). Cortinual photoperiod fieatrertsmrsistedowhratualcaysena'dedvfimligtfiunirmmt Ia'rps. NI=4~hrigtirtemplion Errorbarsrepreaent %oor'lidenoe irtervds. L=lineazQ=quactalictrends "B’v’T‘hbrsigifldartdsigiflma Ps 0.05, 0.01, or 0.001, mpeclively. 67 // fi' W v Y //T 150 . A 125 < 100 4 75 « so. I - £5 ;5 L #___J 25 1 B Number of total lateral shoots Number of flowering lateral shoots Number or flowers and buds E5 0 T L VL/é/ 1004 D E E 80. 2 9 604 W a a - = C ‘3’ 40- c g 20( cf 0 A 4 4 fit 4.7;4 604 . A L’“o” E 5.. 50‘ \ ‘ fl 1 fa» 4o. -—/ g Linens g 30. c. 20y / / / i . . . - . . . .M t . . - . . e . . .,/. . 8 10 12 14 16 18 20 22 24 Nl Photoperiod (h) Figure 6. Effects of various photoperiods afler 15 weeks at 5 °C on number of flowers and buds (A), number of flowering lateral shoots (B), number of total lateral shoots (C), percent flowering laterals (D), and plant height (E) for Oenothera fruticosa ‘Youngii-lapsle in Year 1 ((2) or Year 2 I). Continual photoperiod treatments consisted of -h natural ys extende with light from incandescent lamps. NI = 4-h night intenuption. Error bars @resent 95% confidence intervals. L = linear, Q = quadratic trends. ”3"”, Nonsignificant or significant at P s 0.05, 0.01, or 0.001, respectively. 68 .A. Be A. d page E 8% c2828. boa: B5980 be 2.80th 9: 2m D Em .m .< mama E woe: c2898. 0.:ch DE .v can... 5 Ecmmma on $5 5ng Come: a £9958 at 398398? gee $38 E r .m .e 9:95: Egg 39an ho BE new. 5 .m .$ 9 85 :0 23589 ho «Dorm N chow. om mm mm VN NN om we 9 3 NF om mm mm vw NN ON 9 m— 3 N? an mm on VN mm on or we 3 NF GOV EBECQEC... cod \¢\n\b\vl A A A _ . . Noe , 3e . Be A 8e . Se Nfio o .626: o. 980 m .026: 2 use 057.; E0: QED < 25 was; a mac .9 gm Aom .8 .8 .8 ron SABP/L 3,190 69 ZOO-r 150- 100 A Number of flowers and buds ..." LQ Flower diameter (cm) Plant height (cm) 14 16 18 20 22 24 26 28 30 32 Temperature (°C) Figue 8. Effects of temperature on nurbers of flowers and buds (A), flower diameter (B), and plant height (C) for (brothera fmlr'rnsa ‘Yomgii-Iapsley. Em bars represent 95% confidence intervals. L = Iinea; Q = quadratic trends. N5 "'“Nonsigrifieant or sigrifleant at P: 0.01 or 0.(D1, respectively. 70 A 13,2, , 7 4 ,, 1., . .. ere, ZZZ, ZZZ 7 , C t// ,, . ,anZr b rrrZrWr/rZ/r/ C b a C .D a , ,// / / . /// x ,/ /// // , z , / / . , .r/M . I. ./v//, I / Vl/ ,rwr/W/VV. rv .y/V/ err/Zr // // . //// . ,, ,/,/, , // . . , I , . / / r . .../74W . y. nor/V/fl/Vw, ,I r/rV/va/r/VV/ A m m w .o 2626:0238 wrflV/Zi yZ/W/V .. . Wr/r/Zg/ZWZ VL/ wV///V/r//V/ Comm Mamas! 0mm. momma Unusual-zeta w .w. w. m w o 253 £92 Ego . .7/ AA. / V/V/VIV, 727%,” // . O. rIIv/.Vv///I/ .r 5., wfir/V//./r//V/ /1 .. /./ Air, 44 . 2. . 2441 a .err/Zflflrrr/Z/err. A m m .A 5A.. 2/4». a / m w . I. 227/273. r/ . ,3; . / zx/Avx . // ///, ... a @rZ/r/Z/Mr/ZZ ZZZ» rr V. Plant Growth Regulators Control 65432 E8 .2956 .632“. 1| 0 were: Ancymdol ( 00 rrg-L-I), Daninozide (5830 rrg-L’I), Paiobttraml (30 ngL'l), CHormeqiat (1500 mgL-l) and Uicorazole (15 rrg-L-l). Error bars I388, ardflwrerdianaterSCHorOerdherafirIiam‘Y ii-lapsley. thesof represert95%corfidenceirtervals. Frgue9. B‘feclsofvariousplantgrwvlh regulatorsondaystoflower(A), plartheight 71 60 5° ‘ ' ' ‘ '_o_ Open‘flowers ‘ . + Abscised flowers + Spentflowers 2 40 ~ - - A ...4>.. New open flowers a) 5 a: 30. r. . r _ r . H— O O .8 20 -1 \0‘. , t . . g \°" 2 104. o - O 2 4 6 8 10 12 Days Figure 10. Postharvest responses of Oenothera fi'uticosa 'Youngii-lapsley’ for 12 days at 24.6 °C under a 24~h photoperiod. Error bars represent 95% confidence intervals. 72 4 /_r ‘- Flowering Characteristics of Several Cultivars of Oenothera fruticosa under Two Different Lamp Types for Daylength Extension 73 Abstract Flowering characteristics of Oenothera fruticosa L. ‘Fireworks’, ‘Highlight’, ‘Summer Solstice’, ‘Youngii-lapsley,’ and ssp. glauca were determined. Two nursery sources were used for ssp. glauca and are indicated by a label (“A” or “3"). All plants were cooled for 15 weeks at 5 °C and subsequently grown under a 16-h photoperiod provided by either high-pressure sodium (HPS) or incandescent (INC) lamps to deliver daylength extension in 20 °C greenhouses. Light quantity was increased approximately 27% when day extension was delivered by HPS lamps rather than INC lamps. Data collected included days to flower, number of flowers and buds at first flower, number of nodes on the main stem and final plant height. Lamp type had little to no influence on days to flower. Day extension with HPS rather than INC lamps increased flower number approximately two times for ‘Youngii—lapsley’ and ssp. glauca A, three times for ssp. glauca B, and 3.5 times for ‘Highlight’. Day extension with INC lamps increased plant height at flower in ssp. glauca B and ‘Youngii-lapsley’. Oenothera fruticosa ssp. glauca shows particular promise as a potted flowering plant. 74 Introduction Oenothera fruticosa L. ‘Youngii-lapsley’ is a facultative long-day plant with an obligate vemalization requirement and potential as a flowering potted plant (Frane, 1999; chapter 2). Several other choice selections from this species include 0. fmticosa ‘Fireworks’, 0. fruticosa ‘Highlight’, 0. fruticosa ‘Summer Solstice’ and O. fmticosa ssp. glauca. Because ‘Youngii-lapsley’ manifests flowering potted plant potential we were interested in investigating that of other 0. fruticosa cultivars and O. fruticosa ssp. glauca. Different lamp types used for day extension profoundly affect flowering characteristics of O. fruticosa ‘Youngii-lapsley’. However, a 16-h photoperiod provided by day extension lighting with incandescent (INC) lamps reduced the daily light integral approximately 22% compared to day extension lighting with high-pressure sodium (HPS) lamps. 'Youngii—lapsley’ plants grown under the HPS lamps produced two times the number of flowers produced on plants grown under the INC lamps. Warrington and Norton (1991) found that increasing the daily light integral increased the rate of floral development and number of flower buds in Dendranthema xgrandiflorum (Ramat.) Kitamura, Raphanus sativus L., Zea mays L., and Cucumis sativus L. The two light types differ in photosynthetic photon flux (PPF) and in spectral quality. Incandescent lamps emit a mixture of red and far-red light that is richer in far-red than red (ratio of red 600-700 nm to far-red 700-800 nm <1 75 [Lane et al., 1965]). High- pressure sodium lamps emit a mixture also but with much more red than far-red light (Thimijan and Heins, 1983). For Campanula carpatica Jacq. ‘Blue Clips’, Coreopsis grandiflora Hogg ex Sweet ‘Early Sunrise’, and Coreopsis verticillata L. ‘Moonbeam’, time to flower did not differ for plants grown under a 16-h day above the irradiance saturation point, with day-extension delivered by HPS, INC, metal halide, or cool-white fluorescent lamps (Whitman et al.,1998). However, the different lamp types did influence final plant height. Campanula carpatica ‘Blue Clips’ and C. grandiflora ‘Early Sunrise’ had significantly longer stems when grown under INC lamps compared to HPS, cool-white fluorescent, or metal halide. The objectives of this study were to determine (1) flowering characteristics (e.g., time to flower, flower number, final plant height) and (2) determine the effect of using different lamp types for day-extension lighting for O. fruticosa ‘Fireworks’, ‘Highlight’ ‘Summer Solstice’, ‘Youngii-lapsley’, and ssp. glauca. Materials and Methods Plant material. Plants were obtained from various commercial producers on different dates between 28 Aug. 1998 and 28 Oct. 1998. The dates of arrival at Michigan State University, cold treatment initiation and forcing, average daily temperature, average daily light integral from transplant to average day of flowering, and form of starting plant material for each cultivar are shown in Table 5. We used two nursery sources for O. fruticosa ssp. glauca; one source labeled the plants as O. fruticosa ssp. glauca, and the other labeled them as O. 76 tetragona. According to Griffiths (1994), O. tetragona is the same as O. fruticosa. In the horticultural industry, many Oenothera cultivars including ‘Fireworks’, ‘Highlight’, and ‘Summer Solstice’ are labeled as O. tetragona but are actually cultivars of O. fruticosa (Griffiths, 1994). The flowering behavior and growth habit of the two sets of O. fruticosa ssp. glauca plants were quite different and for clarity, we have labeled them 0. fruticosa ssp. glauca A and O. fmticosa ssp. glauca B, the latter referring to the plants received as O. tetragona. Upon arrival at MSU, bare-root plants of ‘Fireworks’ and ‘Summer Solstice’ were placed into bulb trays in which the root masses were covered with soilless medium and put immediately in the cooler. O. fruticosa ‘Highlight’ were received as very small field-grown divisions and allowed to grow in 13-cm (1 .1-L) pots under natural daylengths (lat. 43 °N) for 15 days at 20 °C before transfer to the cooler. To control an existing Botrytis sp. infestation, O. fmticosa ssp. glauca A were held for 15 days in the 2.5-inch pots in which they arrived before being transferred to the cooler. ‘Youngii-Iapsley‘ and ssp. glauca B were held in plug trays under natural daylengths for five or six days, respectively. Cold treatments. Cold treatments were delivered to plants held in plug trays, bulb trays in which the root masses were covered in soilless medium (Stronglite High Porosity Mix, Stronglite, Pine Bluff, Ark.), or 13-cm pots (‘Highlight’ only) in coolers set at 5 °C. Within the coolers, plants were illuminated for 9 h-dayr1 with cool-white fluorescent lamps (F96T12/CWNHO, Philips, Somerset, N.J.) at approximately 10 “moi-mes“ and watered as necessary using well water acidified 77 with sulfuric acid (H2804) to a pH of approximately 6.0. All plants were given a 15-week cold treatment at 5 °C and subsequently grown in greenhouses set at 20 °C (day/night [D/N]). Plant culture. Before placement in the greenhouses, plugs or bare-root plants were transplanted into 13-cm square plastic pots (1 .1-L volume) with a soilless medium containing sphagnum peat moss, composted pine bark, vermiculite, and perlite (Strong-lite High Porosity Mix, Strong-Lite, Pine Bluff, Ark.) Since ‘Highlight’ already was transplanted into the pots, they were transferred from the cooler to the greenhouse. Pots were spaced at approximately 22 plants/m2. Plants were fertilized at every irrigation with a nutrient solution of well water (EC of 0.70 mS-cm" and 105, 35, and 23 mg-L'1 Ca, Mg, and 8, respectively) acidified with H280, to a titratable alkalinity of 130 mgoL'1 CaCO3 and water soluble fertilizer providing 125-12425.13 N-P-K-Ca mg-L'1 (30% ammoniacal N) plus 1.0-O.5-0.5-O.5-0.1-O.1 mg-L'1 (Fe, Mn, Zn, Cu, B, Mo) (MSU Special, Greencare, Chicago, Ill.) The target range was 0.5-1.2 mS-cm'1 for electrical conductivity (EC) and 5.8-6.2 for pH. In the second year, it was necessary to drench all plants transplanted after 1 Jan. 1999 twice with a 750 mg N-L'1 liquid fertilizer and KCO3 to raise the pH and EC into the desired ranges. Photoperiod and lamp types. All plants were provided a 16-h photoperiod delivered by continual day-extension lighting with either HPS or INC lamps. Day- extension lighting with INC lamps (which provided approximately 1-3 umol-mZ-s-1 at canopy level) were delivered under opaque blackcloth pulled at 1700 HR and 78 opened at 0800 HR every day. The INC lamps were turned on at 1700 HR and turned off at 2400 HR. Day-extension lighting with HPS lamps (which provided approximately 50 umol-mZ-s-t) turned off at 2400 HR and then turned on the following day at 0800 HR, fluctuating on and off during daylight hours using the following regimen: at 0800, HPS lamps turned on and continued until the outside PPF exceeded 400 umol-mZ-s-t If the outside PPF then dropped below 200 umol-mZ-s-1 lamps were again turned on until 2400 HR. Plants in the treatment in which daylength was extended with INC lamps received supplemental lighting from HPS lamps in the same fashion during the hours between 0800 and 1700 HR. Data collected included date of visible bud (without dissection) and first open flower. Days to visible bud and flower and visible bud to flower were calculated for all experiments. At first open flower, total plant height (excluding the container), number of visible flowers and buds, and number of final nodes on the main stem were recorded. Greenhouse temperature control and daily light integral collection. In each greenhouse air temperature was monitored with 36-gauge (0.127-mm-diameter) thermocouples connected to a CR10 datalogger (Campbell Scientific, Logan, Utah). The datalogger collected temperature data every 10 s and recorded the hourly average. Temperatures in each greenhouse were controlled by a Priva environmental computer (model CDY50; Priva, De Lier, The Netherlands). The average daily light integrals were recorded at canopy level using quantum 79 sensors (LI-COR) connected to the CR10 datalogger. Average daily air temperatures from transplant to average date of flowering were calculated for every treatment in each experiment. Average daily light integral data from transplant to average date of flowering were collected only for plants under day- extension with HPS lamps. It was estimated that day extension with INC lamps reduced the daily light integral 22 % compared to high-pressure sodium lamps. This estimation was found by subtracting the amount of light available during daylight hours (0800 to 1700 HR) (as measured by a Ll-COR quantum sensor placed on the bench at canopy height) from the amount of light available during the full 16-h photoperiod provided by day-extension with HPS lamps, and assuming that INC lamps (with an output of 1-3 umol-mZ-sec") contribute essentially no light beyond the compensation point. Data analysis. The experiment was a completely randomized design. Data were analyzed using the general linear models procedure (PROC GLM)(SAS Institute, Cary, NC.) for analysis of variance. Results and Discussion All plants flowered under each lamp type used for day extension. All cultivars flowered within a relatively small range of days (41 to 51) (Table 6). However, final plant height ranged between 24 and 45 cm and number of flowers produced between 12 and 502. Plant appearance differed markedly between cultivars grown under different lamp types (Figure 11). There were no significant effects of lamp type on final node number for any cultivar. 80 The influence of lamp type on flowering time was not great for any cultivar. Time to visible bud was one and two days faster under INC lamps for ‘Youngii-Iapsley’ and ‘Highlight’, respectively (data not shown). Time from visible bud to flower was two or seven days faster under INC lamps for ‘Fireworks’ or ssp. glauca A, respectively, and two days faster under HPS lamps for ‘Summer Solstice’. Lamp-type influence on days to visible bud and days from visible bud to flower was not significant for all other cultivars. For ‘Summer Solstice’, ‘Fireworks’, and ssp. glauca A, time to flower was three, four, and six days slower, respectively, for plants grown under HPS lamps rather than INC lamps (Table 2). Time to flower was one or two days faster for plants grown under HPS day-extension lighting compared to that under INC lamps for ‘Highlight’ or ssp. glauca B, respectively. Lamp type had no significant effect on days to flower for ‘Youngii-lapsley’. It is surprising that the extra light provided by day extension with HPS lamps did not result in consistently faster flowering times for all cultivars. Increasing PPF has been shown to decrease time to flower for many species, including Raphanus sativus L., and Zea mays L. (Warrington and Norton, 1991) and is likely due to an increase in plant temperature observed with an increase in PPF (Faust and Heins, 1997). When Campanula carpatica ‘Blue Clips’, Coreopsis grandiflora ‘Early Sunrise’, and Coreopsis verticillata 'Moonbeam’ were grown under HPS, INC, metal halide, and cool-white fluorescent lamps at similar photosynthetic photon fluxes, lamp type did not influence time to flower (Whitman et al., 1998). 81 Lamp type did not influence plant height on any cultivar except that of ssp. glauca B and ‘Youngii-lapsley’ under incandescent lamps, it increased 3 or 10 cm, respectively, compared to that when HPS lamps were used. Whitman et al. (1998) observed that plant height was increased significantly for Campanula carpatica ‘Blue Clips‘ and Coreopsis grandiflora “Early Sunrise’ when INC lamps rather than when HPS, metal halide, or cool-white fluorescent lamps were used for day-extension lighting. High-pressure sodium lamps used as day-extension lighting increased flower number 2, 2. 3, and 3.5 times compared to that under INC lamps for , ‘Youngii-lapsley’, D. fruticosa ssp. glauca A, O. fruticosa ssp. glauca B, and ‘Highlight’ respectively. Both light quantity and quality have been shown to influence flower number of various species. For example, Warrington and Norton (1991) found that for Dendranthema xgrandiflorum (Ramat.) Kitamura, grown under a 12-h day, flower number increased from 7.2 to 19.0 as daily light integral increased from 10.4 to 29.9 mol-mz-day‘. Rosa xhybn'da ‘Samantha’ plants produced 27-71% more flowers when grown with supplemental light from HPS lamps (at a PPF of 70 to 75 umol-mz-s‘1 one meter above of the pots at night) compared to those grown under HPS lamps fitted with a blue gel filter which reduced the R : FR ratio from 120.95 to 1:2 (Roberts et al., 1993). It is interesting to note that two cultivars (’Fireworks’ and ’Summer Solstice’) showed no effect of lamp type on flower number. 82 Although they were received labeled as the same subspecies, O. fruticosa ssp. glauca A and B behaved very differently when grown similarly and it was not obvious that they were the same species upon inspection. Plants labeled “B” flowered two to 10 days faster, were 8 or 9 cm shorter, and produced =1.5 to 2 times more flowers than those labeled “B”. Plants labeled “8” were much more uniform than “A”, which was expected since ssp. glauca B was vegetatively propagated while ssp. glauca A was propagated by seed. Based on standard deviation, variation for days to flower for A was 5.9 when plants were grown under HPS lamps and 3.6 when they were grown under INC; for B, 0.9 and 1.3, respectively. A preexisting infestation of Botrytis sp. on O. fruticosa ssp. glauca A could have contributed to the lower flower number and slower flowering time and greater variability. Otherwise, we cannot explain their different flowering behaviors. Because of its flowering characteristics, 0. fmticosa ssp. glauca has potential as potted plant crop and may be better suited as one than ‘Youngii- lapsley’. Subspecies glauca B was naturally short and floriferous, particularly when grown under a 16-h photoperiod provided by day-extension lighting with HPS lamps. Without plant growth regulators (PGR), this subspecies has a height/diameter ratio of 2.0, which is in the desired range of 1.5 to 2.0 for a potted flowering plant, as suggested by Sachs (1976). This ratio can be compared to the natural height/diameter ratio of 2.7 for ‘Youngii-Iapsley’. 83 Most of the other cultivars showed some potential but had major drawbacks. ‘Fireworks’ was naturally short but produced few flowers. However, in a preliminary study, rooted stem cuttings of ‘Fireworks’ cooled for eight weeks and subsequently grown under a 16-h photoperiod provided by HPS for day extension attained a height of 26 cm, produced 159 flowers, and showed excellent potted plant potential. ‘Highlight’ could have potted plant potential if the height were controlled by PGR. ’Summer Solstice’ lacked potential because of its natural tallness (40 cm) and relatively low flower count (46 to 59). Summary Different lamp types used for daylength extension had only small effects on time to flower but, for some cultivars, greatly influenced flower number. High- pressure sodium lamps used for day-extension lighting drastically increased flower number of ‘Highlight’, ‘Youngii-lapsley’, and O. fruticosa ssp. glauca A and B. Because of the large variability between 0. fruticosa ssp. glauca produced by different sources, it is advisable to screen stock-plant material before wide- scale production of this species is attempted. For production of ‘Youngii-lapsley’, ‘Highlight’, and ssp. glauca, using HPS lamps for day extension is recommended for producing more attractive plants. Subspecies glauca merits attention for its pot plant potential. 84 Literature Cited Frane, A.J. 1999. Temperature effects on timing and bud development of Coreopsis verticillata ‘Moonbeam’ and flower induction of long-day perennials under different night temperatures. MS Thesis, Michigan State Univ., East Lansing. Faust, J.E. and RD. Heins. 1997. Quantifying the influence of high-pressure sodium lighting on shoot-tip temperature. Acta Hort. 418: 85-91. Griffiths, M. 1994. Index of garden plants. Timber Press, Portland. Lane, H., H. Cathey, and L. Evans. 1965. The dependence of flowering in several long-day plants on the spectral composition of light extending the photoperiod. Amer. J. Bot. 52:1006-1014. Roberts, G.L., M.J. Tsujita, and B. Dansereau. 1993. Supplemental light quality affects budbreak, yield, and vase life of cut roses. HortScience 28(6): 621-622. Sachs, R.M., A.M. Kofranek, and WP. Hackett. 1976. Evaluating new pot plant species. Florists Rev. 159 (4116):35—36, 80-84. Thimijan, R. and RD. Heins. 1983. Photometric, radiometric, and quantum light units of measure: A review of procedures for interconversion. HortScience 1 8:818—822. Warrington, I.J. and RA. Norton. 1991. An evaluation of plant growth and development under various daily quantum integrals. J. Amer. Soc. Hort. Sci. 116(3):544-551. Whitman, C.M., R.D. Heins, A.C. Cameron, and W.H. Carlson. 1998. Lamp types and irradiance level for daylength extensions influence flowering of Campanula carpatica ‘Blue Clips’, Coreopsis grandiflora ‘Early Sunrise’, and Coreopsis verticillata ‘Moonbeam’. J. Amer. Soc. Hort. Sci. 123(5):802-807. 85 .889 AdeV 8:68 9336 -29; 5? 3.8988 59.03% 595 .565 Eco—a ..2 8282 25 2o; 8% Eu... .869 .02: E8328... 5? 3.82.8 59.633 895 $65 $53 .2 388:8 mm; 5.8.5 Em: 2.8 ~ m 3: tom mm: 82 82 82 885 «.9 Now 02. do“. E 502 v .80 mm __mo-va .3.» < mdw mdw we: 83 83 mam? 8sz ed mdw 02. den. 5 .Eom 3 .92. mm coc_-m.~ dmm 0.: new we... 82 82 8% >282 od odm oz. .cmw NP .Emw mm .Eow mm __mo-mm -__mc:o> Q: :N 9%.. $0.. mam? mam? a.» How oz. .52. m: .amm on .33 9 62-98 3929: v.9 Now we: 32. camp 89 .8:me Rm mdm 02. .80 me .Eew N .58 N 82-9.3 5&an Pd? 33 we: mam? mam? mam? Na.» mdm 02. den. 3 .fiom N .53 N 69.92 mxeosmLE 8L A.->mu.~E._oEv QBanEQ comma comma $3me .9925 Em: 2:8 9:20.. E853: _m>Em .53 32:30 ommeo>< cab 9:3 9mm. 28 memo .8 9mm 9555 2:3 cameo}. ngm. Emommccmos n 02. ”8E3 Esfiow 2333 .29: n ma: .wnocmaowoca 5-3 Low c2338 59633 82.3 2 new: aEm. he one n 83 QES ..mtofiE Ema .2 82:8 EmLoEu o. .82 8:30 .amm 38:3: .0 Loco .9 new ..<.. gone. 9: .8sz d3 gouge .0 new ..oo=m_ow LoEEam. ..>m_mam_-__mc:o>. ..EQEQI. ..mvtoaoi. mmoozzt 8.35050 .2 $525 Em: 2:8 momemzw new .oLBmLmQEm. 2:3 momem>m 6588 ”cosmEE EmEumoz 28 ”...;th 5 $38 ”.mtmfiE Ema 9.355 .m aim... 86 8:85:35 5 mm 3 or mo... 2. m... 3. 2 oz. 8.83.653 ... wz . 8:85:35 mom om av w mo... m3 8 :V A oz. £939: wz wz ... 8:85:9m mm ov wv or de mv ow mv 0.. oz. 8.7..ow ..oEE3m mz wz .... 8:85:90 mm Va 5 m mo... we mm 2» v 02. mfozo..n_ 83: ocm AEo. Ego: EmE Logo: 2 $80 $8.: 8?» 9:8 .8530 98.6: .0 .02 8.826: .0 .02 8b aEm. .8 88.8.38 8253835 9 8.2 8:85:35 .0 28m... .00 m «m 9.82. 3 .52, :98: 295 8858 __< .maEm. .02.. 28888:. .0 8E3. 5&1. E368 938995.: .655 5.2. 8.888 >8 .5 882.8 68:899.: ....9 53:3 00 cu «m :26... 83:5 .98 88.33.: .0 ocm 8853.: 3.85250 .0 9858 .988 .0 8289888 8.826.“. .m aim... 87 5.8.688: .50... .6 .56 .mod w n. 6 E8555 .0 2855.802 .....ka C .‘ 8:85:05 om F 5 03 5 vm «.3 ¢ ”2 i 8:85:95 E ov on on NV 3 :0 52 0.. or coco maI 02. THE 02. m 83%.. .Qmm < 8sz d8 3:8 o 233 88 Figure 11. Cultivars of Oenothera fmticosa and O. fruticosa ssp. glauca grown at 20°C under a 16-h photoperiod provided by a 9-h natural day plus 7-h day extension with either high-pressure sodium lamps (labeled in the photographs as HID for high-intensity discharge) or incandescent lamps (inc). The labels “A” and “B” after ssp. glauca refer to different nursery sources. 89 A 'Fireworks' B 'Summer Solstice' 16hr HID 16 hr inc C 'Highlight' m p. \ — “We 16 hr HID 16 hr inc 16 hr HID 16 hr inc 90 Effects of Cold Treatment, Photoperiod, and Forcing Temperature on Growth and Development of Stokesia Iaevis ‘Klaus Jelitto’ 91 Abstract The influences of cold treatment, photoperiod, and forcing temperature on growth and development of Stokesia Iaevis ‘Klaus Jelitto' were determined. Plants were cooled at 5 °C for 0, 3, 6, 9, 12, or 15 weeks under a 9-h photoperiod and subsequently forced in a 20 °C greenhouse under a 16—h photoperiod. Percentage flowering increased and time to visible bud and to flower decreased with increasing cold durations. Photoperiods of 10, 12, 13, 14, 16, 24 h, or a 4-h night interruption (Nl) were tested in a 20 °C greenhouse following 0 or 15 weeks of cold treatment at 5 °C. Before cold treatment, plants flowered under photoperiods between 10 h and 14 h and one plant flowered under 24 h in Year 2. Percentage flowering, time to visible bud and to flower decreased with increasing photoperiod. After cold treatment, plants flowered under all photoperiods (except 24 h in Year 1). Percentage flowering, time to visible bud and to flower, and flower number decreased with increasing photoperiod. Plant height increased with increasing photoperiod. Stokesia Iaevis ‘Klaus Jelitto’ may best be described as an intermediate—day plant since highest flowering percentage and fastest time to flower were observed under photoperiods between 12 and 13 h both before and after cold treatment. To determine the effect of forcing temperature on flowering of ‘Klaus Jelitto', plants were cooled for 15 weeks and then forced under a 16-h photoperiod at 14, 17, 20, 23, 26, or 29 °C. Time to visible bud and to flower increased and percentage 92 flowering decreased at higher forcing temperatures. Plant height and flower number decreased greatly as temperature increased. 93 Introduction Stokesia Iaevis (Hill) Greene is an herbaceous perennial native to the southeastern coastal plain of South Carolina, Louisiana, and the panhandle of Florida (Gunn and White, 1974). Plants produce large (approximately 10 cm diameter) lavender blue flowers usually between July and October, but in warmer climates, such as in the southern United States, flowering can occur year-round (Fischer,1998). ‘Klaus Jelitto’ is a vegetatively propagated cultivar of Stokesia Iaevis with large, light lavender-blue flowers (Fischer,1998). Stokesia Iaevis is used for home perennial gardens, and rarely, for commercial cut flower production (Nau, 1996). Since 1970, Stokesia Iaevis has been investigated for its potential as an oilseed crop (Gunn and White, 1974). The achenes of Stokesia Iaevis contain 40% oil, and of that, 70% is epoxyoleic acid (also known as vemolic acid) which is easily converted to epoxy oil, a material used in the manufacture of plastics (Gunn and White, 1974; Perdue et al., 1986). Flower induction in many plant species is controlled by vemalization and photoperiod. The promotion of flowering by cold treatment is defined as vemalization while photoperiodism is the control of flowering by daylength (Lang, 1965). Both vemalization and photoperiod requirements can be facultative or obligate. Vemalization and photoperiod are inductive processes and once floral induction has occurred, plants can be moved to noninductive conditions, and flowering will still occur (Lang, 1965). Modifying the length of the cold treatment or photoperiods after induction has occurred can dramatically influence plant 94 appearance. For example, plants of Leucanthemum xsuperbum Bergmans ex J. Ingram ‘Snowcap’ cooled for 15 weeks at 5 °C produced an average of 3 inflorescences under a 10-h photoperiod while they produced an average of 11 inflorescences under 16 h (Runkle et al., 1998). As cold treatment duration increased from O to 15 weeks, inflorescence number increased by approximately three times for plants of ‘Snowcap’ subsequently grown under a 4-h night interruption (NI), (Runkle et al., 1998). Such manipulations of cold treatments and photoperiod are important for producing high-quality crops. In a preliminary experiment, ‘Klaus Jelitto’ was tested either under a 9-h short day or a 4-h Nl before and after a 15-week cold treatment at 5 °C. ‘Klaus Jelitto’ was tentatively proposed to be a facultative short-day plant with an obligate requirement for cold treatment (Frane, 1999). Knowledge of the effect of temperature on plant development is important to accurately time crop production. In the absence of the effects such as photoperiod or vemalization, in general, the rate of development increases linearly with temperature within a limited range of growing temperatures (Roberts and Summerfleld, 1987). Such responses are also evident in photoperiod- and vemalization-sensitive species when the separate effects of photoperiod and vemalization are removed from those of post-vemalization temperatures. The relationship between temperature and rate of development toward flowering (1/DTF where DTF is days to flower) can be described as follows: 1/DTF = b, + b,*T [1] 95 where b0 and b1 are constants and T is temperature (°C) (Roberts and Summerfield, 1987). The base temperature (T.,) , or the temperature at or below which the rate of progress toward flowering is zero, and thermal time, or the number of units (°days) of thermal time above a base temperature required for flowering can be calculated from the constants in Equation 1 as follows: T,= -bo/b, [2] °days=1lb1 [3] Besides influencing rate of flowering, temperature influences other aspects of growth and development such as number of flowers and plant height at flower. For Antinhinum majus L. ‘Jackpot’, there was an increase in stem length, spike length and number of florets as temperature decreased from 21 to 10 °C (Maginnes and Langhans. 1961). Flower bud number, plant height, and flower diameter of Coreopsis grandiflora Hogg ex Sweet ‘Sunray’, Leucanthemum xsuperbum ‘Snowcap’, and Rudbeckia fulgida Ait. ‘Goldsturm’ decreased as forcing temperature increased from 15 to 27 °C (Yuan et al., 1998) We could find no literature on forcing temperatures for ‘Klaus Jelitto' or any cultivar of Stokesia Iaevis. The objectives of these experiments were to describe and quantify the effects of cold treatment, photoperiod, and forcing temperature on growth and development of Stokesia Iaevis ‘Klaus Jelitto’. 96 Materials and Methods General. Plants of Stokesia Iaevis “Klaus Jelitto' were received as crown divisions or rooted basal cuttings in either 2.5-inch pots or in 36-cell plug trays on 15 Oct. 1997 (Year 1) and 28 Aug. 1998 or 16 Oct. 1998 (Year 2) from two separate commercial producers. Upon arrival at Michigan State University, all plants were held in the 2.5— inch pots or plug trays and exposed to natural daylengths (lat. 43 °N;z11-12 h) until the experiments began. Cold treatments were delivered to plants held in 2.5-inch pots or plug trays in coolers set at 5 °C. Within the coolers, plants were illuminated for 9 h-day1 with cool-white fluorescent lamps (F96T12/CWNHO, Philips, Somerset, N.J.) at approximately 10 ,ufilO'Tfl’S" and watered as necessary (approximately 2 times per week) using well water acidified with sulfuric acid (H280,) to a pH of approximately 6.0. Before placement in the greenhouses (set at 20 °C day/night [DIN] unless otherwise stated), plugs were transplanted into 13-cm square plastic pots (1 .1-L volume) with a soilless medium containing sphagnum peat moss, composted pine bark, vermiculite, and perlite (Strong-lite High Porosity Mix, Strong-Lite, Pine Bluff, Ark.). Pots were spaced at approximately 22 plants/m2. Plants were fertilized at every irrigation with a nutrient solution of well water (EC of 0.70 mS-cm'1 and 105, 35, and 23 mg-L'1 Ca, Mg, and S, respectively) acidified with HZSO4 to a titratable alkalinity of 130 mg-L'1 CaCO3 and water soluble fertilizer providing 125-12-125-13 N-P-K-Ca mg-L’1 (30% ammoniacal N) plus 1.0-0.5-0.5- 0.5-0.1-0.1 mg-L‘1 (Fe, Mn, Zn, Cu, B, Mo) (MSU Special, Greencare, Chicago, 97 Ill.). The target range was 0.5-1.2 mS-cm'1 for electrical conductivity (EC) and 5.8-6.2 for pH. In the second year, it was necessary to drench all plants transplanted after 1 Jan. 1999 twice with a 750 mg N-L'1 liquid fertilizer and KCO3 to raise the pH and EC into the desired ranges. Cold duration (Expt. 1). Six durations of cold treatment were tested (0, 3, 6, 9, 12, or 15 weeks). Experiments were initiated on 18 Oct. 1997 (Year 1) and 7 Oct. 1998 (Year 2). No initial node count was taken in Year 1. In Year 2, the ; initial number of nodes was counted at transplant and averaged 9.8 nodes. Ten plants were removed from the cooler after each treatment duration, transplanted and placed in the greenhouse. A 16-h photoperiod was provided by four-hundred-watt high-pressure sodium lamps (HPS) switched on at 0800 HR and off at 2400 HR. HPS lamps provided both the photoperiod and supplemental lighting with a photosynthetic photon flux (PPF) of 50 umol-mZ-s-' at plant height starting at 0800 and continuing until the outside PPF exceeded 400 umol-mZ-s". If the outside PPF then dropped below 200 umol~m2-s~' lamps were again turned on until 2400 HR. Photoperiod (Expt.2). Seven photoperiods (10, 12, 13, 14, 16, 24 h and a 4-h night interruption [MD and two cold treatment durations (0 or 15 weeks at 5 °C) were tested. Seventy plants were transplanted on 18 Oct. 1997 (Year 1) and 28 Oct. 1998 (Year 2) and placed in a greenhouse. Ten plants were used per treatment and treatments were assigned randomly each year to greenhouse benches. The remaining plants were left in the 2.5-inch pots and placed in the 98 cooler. After 15 weeks of cold treatment, seventy plants were transplanted on 31 Jan. 1998 (Year 1) and 14 Feb. 1999 (Year 2) and placed in the same photoperiod treatments as the first set of seventy plants. Plants averaged 10.8 nodes in Year 1 and 10.3 nodes in Year 2. Opaque black cloth was pulled at 1700 HR and opened at 0800 HR every day on all benches so all plants received a similar daily light integral. Continual photoperiods were delivered by day-extension lighting using incandescent lamps which were turned on at 1700 HR and turned off after each photoperiod was completed. The 4-h night interruption was provided by incandescent lamps which were turned on at 2200 HR and turned off at 0200 HR. Day-extension and NI lighting using incandescent lamps provided approximately 1-3 umol-st1 at canopy level. Supplemental lighting during daylight hours was given as described for Expt. 1, except that HPS lamps were terminated at 1700 HR. Forcing Temperature (Expt. 3). On 7 Oct. 1998, plants left in plug trays were placed directly into the cooler for 15 weeks until 20 Jan. 1999, when they were transplanted. At transplant, plants averaged 10.1 leaves. Ten plants were placed into each of 6 greenhouses set to 14, 17, 20, 23, 26, and 29 °C. Plants received a continual 16-h photoperiod provided by HPS lamps that remained on continuously from 0800 HR to 2400 HR and delivered approximately 90 umol-mz-s". Data collection and analysis. The data collected for all experiments included date of first visible bud (without dissection) and date of first open flower. Days to 99 visible bud, days to flower, and days from visible bud to flower were calculated for all experiments. At first open flower, total plant height (excluding the container), number of visible flower buds, and number of final nodes on the main stem were recorded. The number of final nodes was counted to the last visible leaf and when the initial node count was taken, included the number of initial nodes. The number of new nodes produced during forcing was calculated by subtracting the initial number of leaves from the final number of leaves, and the number of new nodes beneath the first flower (in Year 2) was calculated by subtracting the number of initial nodes from the number of nodes beneath the first flower. For all experiments, plants that did not have visible bud after 15 weeks of forcing were considered non-flowering and were discarded. Plants that died during the experiment were discarded and not included in the results. All experiments were completely randomized designs. Data were analyzed using the general linear models procedure (PROC GLM) (SAS Institute, Cary, NC.) for analysis of variance and linear regression procedure (PROC REG) for regression models. All experiments were analyzed with Linear and Quadratic trend contrasts. The cold duration and photoperiod experiments (Expts.1 and 2, respectively) were replicated in time, and conducted once per year. The forcing temperature experiment (Expt. 3) was conducted only once, in Year 2. Greenhouse temperature control and daily light integral collection. Air temperature was monitored in each greenhouse with 36-gauge (0.127—mm- 100 diameter) thermocouples connected to a CR 10 datalogger (Campbell Scientific, Logan, Utah). The datalogger collected temperature data every 10 s and recorded the hourly average. Average daily air temperatures from transplant to average date of flowering for Expts. 1 and 2 are shown in Tables 7 and 8. Average daily temperatures for three developmental stages of plants in Expt. 3 are shown in Table 9. Temperatures in each greenhouse were controlled by a Priva environmental computer (model CD750; Priva, De Lier, The Netherlands). Actual average daily air temperatures were calculated and used in all analyses. Average daily light integral data were collected only for Expt. 1 in Year 2. The average daily light integrals were recorded at canopy level (at the same interval as temperature data) using quantum sensors (LI-COR) connected to the CR 10 datalogger. It was estimated that plants in Expt. 2 received approximately 78% of the daily light integral that plants received in Expt. 1. This estimation was found by subtracting the amount of light available during daylight hours (0800 to 1700 HR) from the amount of light available during the full 16-h photoperiod of day extension with high-pressure sodium lamps. Supplemental lighting was similar between the houses from 0800 to 1700 HR and it was assumed that day- extension lighting with incandescent lamps (with an output of 1-3 umol-mZ-sec") contribute essentially no light beyond the compensation point. Results and Discussion Cold duration (Expt. 1). Fifty to 60 °/o of ‘Klaus Jelitto’ plants flowered without a cold treatment (Figure 12A). In Year 1, percentage flowering reached 100 after 101 6 weeks of cold treatment (Figure 12A) while in Year 2, flowering percentage did not reach 100 until plants had received 15 weeks cold treatment. Days to visible bud and days to flower decreased quadratically (P s 0.01 for days to flower; P g 0.001 for days to visible bud) in Year 1, although there were no significant trends for either variable in Year 2 (Figures 12B, C). In Year 1, time to flower decreased greatly as cold duration increased from zero to 6 weeks (112 to 79 days). The decrease in days to flower as cold duration increased from 6 to 15 weeks was slight (79 to 74 days), thus it appears that vemalization was saturated after 6 weeks of 5 °C treatment. Increasing cold duration generally decreased variability for flower timing. For example, variability based on standard deviation decreased from 23.8 to 11.3 days as cold duration increased from zero to 15 weeks in Year 2. Based on these results, we conclude that ‘Klaus Jelitto’ has a facultative requirement for vemalization. There was a large difference between the two years for plants that did not receive any cold treatment but only small differences between the years for plants cooled 23 weeks. Plants without a cold treatment flowered slowly (112 days) in Year 1 while in Year 2, they flowered relatively quickly (76 days). Starting plant material was different between the years: 2.5-inch pots in Year 1, and 36-cell plugs in Year 2 each received from a different nursery. Although no initial node count was taken in Year 1, the initial node count of plants in 36-cell plug trays averaged 9.8, which was very similar to the node counts found for plants in 2.5-inch pots in both Year 1 and Year 2 (10.8 and 10.2, respectively), 102 which would have been similar to the node count for plants in Expt. 1, in Year 1. We can not readily explain the different flowering behavior between the noncooled plants in the two years. The experiments were initiated within 11 days of each other between the years and, although we do not have the light data for Year 1, we assume that plants in each year received relatively similar amounts of light. Temperature was carefully controlled; the greatest difference between the years was 1.2 °C, thus temperature probably did not cause the differences in flowering that we observed between the years or between the different cold duration treatments. There were small but significant differences for the effect of cold treatment on days from visible bud to flower. In Year 1, days from visible bud to flower decreased quadratically (P s 0.01), while in Year 2, there was an increasing linear (P s 0.01) trend (Figure 12D). Although significant, the differences in time from visible bud to flower between different cold durations within a year or between the years were relatively small: the greatest difference between treatments within a year and treatments between years was 11 days. While the change in flower timing with increased cold treatment was only clearly observed in Year 1, the number of nodes counted to the last visible leaf did reflect earlier floral initiation with increasing cold treatment in both years (Figure 13A). There were no significant differences for the final or new node number beneath the first flower in Year 2, which reflects the lack of significant linear or quadratric trends observed in Year 2. The greatest decrease in final 103 and new number of nodes (31 to 26 for final nodes in Year 2) occurred as cold duration increased from zero to six weeks, after which the change in number of nodes was very small (26 to 25 for final nodes in Year 2) as cold duration was increased to 15 weeks. In addition, variability of node number (both final and new nodes) decreased as cold duration increased. These observations are consistent with the hypothesis that vemalization of ‘Klaus Jelitto’ is saturated after 6 weeks of cold treatment. The number of new nodes in Year 2 was almost identical to the number of final nodes in Year 1. This was most likely due to the fact that initial leaf counts were not taken in Year 1, and the leaves present on the plant at the beginning of the experiment had probably senesced by the time the plant flowered. The number of final nodes in Year 1 most likely represented the number of new nodes produced during forcing. It is interesting to note that the faster flowering observed for plants without a cold treatment in Year 2 was not reflected by either the final or new number of nodes. There were large differences in number of flower buds between the years (Figure 13B). There were no significant trends for the effect of cold duration on the number of flower buds in Year 1, but there was a significant (P s 0.01) linear increase in Year 2. The number of flower buds per plant was variable but, increased from five for noncooled plants to 13 for plants cooled for 15 weeks. There was little effect of cold duration on plant height. In Year 1, height decreased linearly (P 3 0.001) from 59 to 49 cm as cold duration increased from 104 zero to 15 weeks (Figure 130). There were no significant differences for height in Year 2. Photoperiod (Expt. 2) Without a cold treatment, in Year 1, “Klaus Jelitto’ plants flowered only under photoperiods between 12 and 14 h; flowering percentage declined from 75 to 16 in this photoperiod range (Figure 14A). Percentage flowering was much higher for noncooled plants in the second year where all plants flowered under photoperiods between 10 and 13 h and one plant each flowered under 14 and 24 h. No plants flowered under night interruption (NI) without cold treatment in either year. While plants did not flower under 16 h without a cold treatment in this experiment, 50 and 60% flowered without a cold treatment in Expt.1 (in Year 1 and Year 2, respectively). Plants in Expt. 2 only received about 78% of the light received by the plants in Expt. 1 which could explain some of the difference. It has been observed for several cold-requiring species that high light levels can induce flowering without cold treatment. Blondon and Chouard (1965) found that a cold-requiring clone of Dactylis glomerata L. flowered under a 16-h photoperiod without a cold treatment when provided with a 16-h day at a photosynthetic photon flux of 280 umol-mz-sec1 (24.2 mol-mZ-day"). In Year 2, the plants that flowered under 10 h and the single plant that flowered under 14 h were very slow to flower (135 or 139 days, respectively) compared to plants under 12 or 13 h photoperiods, which flowered in 81 or 86 days, respectively (Figures 14B, C, D). The plant that flowered under the 24-h 105 photoperiod flowered in 74 days. The two plants that flowered under 14 h in Year 1 flowered 63 days faster than the one plant that flowered under 14 h in Year 2. Average daily temperature was only 1 °C higher in Year 2 compared to Year 1, and can not explain this difference. The experiment in Year 2 started only 10 days later than in Year 1, thus it is not likely that differing light levels caused the difference. We have no explanation for the large difference in percentage flowering or time to flower between the two years for noncooled plants. We speculate that if we had left the plants on the bench for longer than 100 days, more plants would have eventually flowered. After the cold treatment, percentage flowering was 90 or 100 under 10 h and decreased to 0 or 10 under 24 h in Year 1or 2, respectively (Figure 15A). Eighty and 90 percent of the plants flowered under Nl in Year 1 and 2, respectively. Following cold treatment, days to visible bud and days to flower changed little between photoperiod treatments except for the one group of plants under 10 h in Year 1 (Figures 15B, C). In Year 1, plants grown under 10 h flowered in 122 days, while plants grown under 10 h in Year 2 flowered in 79 days. In Year 2, plants under 10 h were only approximately 10 days slower than plants under other treatments. We have no explanation for the large differences for plants under 10 h between the years. Days from visible bud to flower changed little between photoperiods. In Year 1, there was a small tendency for increasing 106 days from visible bud to flower while in Year 2, there was a small tendency for decreasing days from visible bud to flower with increasing photoperiod (Figure 15D). However, differences in days from visible bud to flower between each photoperiod were small, as were the differences between years. The greatest difference between the years for days to visible bud to flower before or after cold treatment was 11 days. Days to visible bud, day from visible bud to flower, and days to flower for plants grown under NI were similar to plants grown under photoperiods 214h in both years. In general, cold-treated plants flowered only slightly faster than flowering non-cold-treated plants, although the effect of cold treatment on flower timing differed for plants grown under different photoperiods. For example, for plants grown under 12 h in Year 1, days to flower decreased =2 weeks after cold treatment, while days to flower decreased 8 weeks after cold treatment for plants grown under 10 h in Year 2 (Figures 14B,15B). Variability generally decreased following cold treatment. For example, the standard deviation for days to flower of plants grown under 13h before and after cold treatment in Year 2 decreased from 13.2 to 3.5. The number of final and new nodes for plants without a cold treatment decreased linearly (P g 0.001) in Year 2, and quadratically (P g 0.001) after a cold treatment in Year 1 (Figures 16A,B). There were no significant trends for either the number of final or new nodes without a cold treatment in Year 1, or 107 following a cold treatment in Year 2. The decrease observed in both final and new nodes as photoperiod increased from 10 to 24 h reflects the decrease in flower timing seen in Year 1. In Year 2, there were no significant trends for the number of final or new nodes beneath the flower following cold treatment. The lower flowering percentage observed under photoperiods 214 h together with the slower time to flower under photoperiods 310 h both before and after a cold treatment suggest that ‘Klaus Jelitto‘ does not behave similar to either a short-day plant response or a long-day plant response. On the basis of these results, we conclude that “Klaus Jelitto' is best described as a facultative intermediate-day plant. Allard (1938) described intermediate-day plants as those which flower when the day is neither too long nor too short. For example, Mikania scandens L. flowered only under photoperiods of 12 to 16 h (Allard, 1938). Only a few species, including Mikania scandens L., Phaseolus polystachios (L.) B.S.P., Eupaton'um h yssopifolium var. h yssopifolium L., a variety of Saccharum spontaneum L. (28N. G. 292) (Allard, 1938), Salsola komarovii ljin (Takeno et al., 1995), and Echinacea purpurea (L.) Moench ‘Magnus’ (E. Runkle, unpublished data) have been documented as intermediate- day plants. Thomas and Vince-Prue (1997) suggested that intermediate-day plants may be a variation of the short-day type response since short day plants also need some light in each day to flower. However, the daylength limitations for short-day plants are fairly wide, while for intermediate-day plants, the range of 108 effective photoperiods is typically quite narrow. Alternatively, Sachs (1956) suggested that intermediate-day plants may actually have dual daylength requirements for flowering but the long—day and short-day ranges overlap, so that intermediate photoperiods satisfy both photoperiodic requirements. Studies are- needed to test these hypotheses for the intermediate-day response observed for Stokesia Iaevis ‘Klaus Jelitto’ There was no consistent effect of photoperiod on flower bud number of non-cold treated plants (Figure 160). After cold treatment, the number of flower buds decreased linearly (P g 0.001 in Year 1; P s 0.01 in Year 2) with increasing photoperiod in both years. Flower bud number increased after a cold treatment only for plants grown under 12-h photoperiods. Cooled plants that flowered under Nl produced as many flowers as those grown under 214 h in both years. Plant height at first flower for plants with or without a cold treatment increased markedly as photoperiod increased from 10 and 14 h (Figure 160); for cold-treated plants in Year 1, plant height increased from 27 to 47 cm. Plant height for cooled plants grown under NI was similar to cooled plants grown under 13 h in Year 1 and similar to cooled plants grown under photoperiods 214h in Year 2. The effect of photoperiod on plant height and number of flowers drastically changed the appearance of the plants grown under different photoperiods. Those grown under photoperiods 5.12 h were short, produced many flowers, and 109 could make a very acceptable potted flowering plant. However, when plants were grown under photoperiods 213 h, the plants were too tall and produced too few flowers to be considered as a potted flowering plant. Knowledge of effective plant growth regulators would be beneficial for production of ‘Klaus Jelitto’ if plants are to be produced under long photoperiods. However, if plants of ‘Klaus Jelitto’ are produced for cut-flower production, long photoperiods may be more appropriate for the longer stems that result under longer photoperiods. Flowering of ‘Klaus Jelitto’ was extremely variable. There were large differences in percent flowering between years for both Expt. 1 and 2 (Figures 12A, 14A). Percentage flowering in Expt. 1 was generally higher in Year 1 compared to that in Year 2, but for Expt. 2, percentage flowering was much higher in Year 2 compared to that in Year 1. For example, in Expt. 2, 100% of plants flowered under 10, 12, and 13 h in Year 2 while no plants flowered under 10 h and only at 75 or 30% plants flowered under 12 or 13 h, respectively, in Year 1. This result was surprising since the plant material used for this experiment came from the same nursery and had similar starting plant sizes in both years (10.8 nodes in Year 1 and 10.3 nodes in Year 2). In addition to variability for percent flowering, there were wide ranges in days to flower even within a single treatment group. For example, all plants cooled for 15 weeks in Expt. 1 (Year 2) flowered over a range of 35 days, an unacceptably large period of time for flowering of a crop in the floriculture industry. 110 Forcing Temperature (Expt. 3) Percentage flowering decreased from 100 at temperatures 320 °C, to 90 at temperatures between 23 and 26 °C, and to 50 at 29 °C (data not shown). Days to visible bud, days from visible bud to flower, and days to flower of ‘Klaus Jelitto’ decreased with increasing temperatures (Figures 17A, B, C). Plants forced at 15.9 °C flowered in 90 days while plants forced at 29.9 °C flowered in 42 days. Rates of progress to visible bud, from visible bud to flower, and to flower were linear within the studied temperature range (Figures 170, E, F). The parameters for the linear regression analysis are shown in Table 10. The reciprocal of the linear regression lines (shown in Figures 17A, B, C) described the relationship between temperature and the time required to reach each developmental stage well. Base temperatures ranged from 3.2 to 4.2. The degrees days required to reach first open flower were 1104 when calculated from the constants obtained from the linear regression analysis with day to flower data and 1035 when the degree days was calculated by adding the data from transplant to visible bud and visible bud to flower. Degree-days and base temperature are important concepts in production of flowering plants since they allow for prediction of flowering time in environments where the temperatures are fluctuating, and the mean temperature is known (Roberts and Summerfield, 1987). Actual average days to flower did not differ from predicted days to flower more than 3 days for any temperature (Figure 17C). 111 The number of final and new nodes to the last visible leaf decreased quadratically (P s 0.05) as temperature increased (Figure 18A), although the actual decrease was small (29 to 23 for final nodes). The quadratic trend of the decrease in final and new nodes relates to the quadratic pattern observed for the decrease in days to visible bud, days from visible bud to flower, and days to flower as temperatures increase (Figures 17A, B, C). There were no significant differences for the number of final or new nodes beneath the flower. Flower number decreased 282% as temperature increased from 15.9 to 29.9 °C (Figure 18B). Others have observed that flower number decreases with increasing forcing temperature; examples include Coreopsis grandiflora ‘Sunray’, Rudbeckia fulgida.’Goldsturm’, Leucanthemum xsuperbum ‘Snowcap' (Yuan et al., 1998), and Campanula carpatica Jacq. ‘Blue Clips’ (Whitman et al., 1997). Plant height decreased dramatically with increasing temperature (Figure 180). Tallest plants (average height of 59 cm) were those forced 18.4 °C and the shortest plants (average of 21 cm) were forced at 29.9 °C. Plant height of many species increases as the difference between day temperature and night temperature increases (DIF) (Enrvin and Heins, 1995). The DIF to which plants of ‘Klaus Jelitto’ were subjected during the developmental stages of transplant to visible bud, visible bud to flower, and transplant to flower are presented in Table 11. The largest values for DIF were observed in the house set at 17 °C (actual temperature averaged 18.4 °C during the time between transplant and average 112 day to flower), which corresponds to the temperature setting where tallest plant heights observed, thus DIF may have contributed to some of the differences in plant height. S_um_m_ant Stokesia Iaevis “Klaus Jelitto‘ has a facultative requirement for vemalization. Time to flower decreased as cold treatment increased. Before cold treatment, flowering generally occurred only under photoperiods between 10 and 14 h. However, when plants were grown without a cold treatment under 16 h in high light conditions, such as provided by high-pressure sodium lamps in Expt.1, flowering occurred but at relatively low percentages (50 to 60%). After cold treatment, plants flowered under all photoperiods but flowering percentage decreased as photoperiod increased from 14- to 24-h. In Year 1, no plants flowered under 24 h. Time to flower was slowest under 10 h and decreased as photoperiod increased. Based on this study, plants of “Klaus Jelitto’ flower most rapidly, completely, and uniformly after a cold treatment under photoperiods between 12 and 13 h and can best be described as a facultative intermediate- day plant. Time to flower and percentage flowering decreased with increasing forcing temperature. A forcing temperature of 18 °C and photoperiods between 12-13 h are recommended for maximizing percentage flowering and flower number. 113 Literature Cited Allard, HA 1938. Complete or partial inhibition of flowering in certain plants when the days are too short or too long. J. Agric. Res. 57: 775-789. Blondon, F. and P. Chouard. 1965. Facteurs extemes capables de provoquer la floraison du Dactylis glomerata L. Action de la temperature. CR. Acad. Sc. 260: 6966-6969. Erwin, J.E., and RD. Heins. 1995. Therrnomorphogenic responses in stem and leaf development. HortScience 30: 940-949. Fischer, T. 1998. Stoke’s Aster— Stokesia Iaevis. Horticulture 95: 34. Frane, A.J. 1999. Temperature effects on timing and bud development of Coreopsis verticillata ‘Moonbeam’ and flower induction of long-day perennials under different night temperatures. MS Thesis, Michigan State Univ., East Lansing. Gunn, CR. and GA. White. 1974. Stokesia Iaevis: Taxonomy and Economic Value. Econ Bot 28: 130-135. Lang, A. 1965. Physiology of flower initiation, p.1380-1535. In: W. Ruhland (ed.). Encyclopedia of plant physiology. Springer-Verlag, Berlin. Maginnes, EA. and R.W. Langhans. 1961. The effect of photoperiod and temperature on initiation and flowering of snapdragon (Antirrhinum majus— variety Jackpot). J. Amer. Soc. Hort. Sci. 77: 600-607. Nau, J. 1996. Ball perennial manual: Propagation and production. Ball, Batavia, IL. Perdue, R.E., K.D. Carlson, and MG. Gilbert. 1986. Vemonia galamensis, potential new crop source of epoxy acid. Econ Bot. 40: 5468. Roberts, E.H. and R.J. Summerfield. 1987. Measurement and prediction of flowering in annual crops, p. 17-50. In: J.G. Atherton (ed.). Manipulation of flowering. Buttenlvorths, London. 114 Runkle, E.S., R.D. Heins, A.C. Cameron, and W.H. Carlson. 1998. Flowering of Leucanthemum xsuperbum ‘Snowcap’ in response to photoperiod and cold treatment. HortScience 33(6): 1003-1006. Sachs, RM. 1956. Floral initiation in Cestrum noctumum l. A long-short day plant. PI. Physiol. 31: 185-192. Takeno, K., M. Takahashi, and K. Watanabe. 1995. Flowering response of an intermediate-day plant, Salsola komarovii ljin under different photoperiodic conditions. J. Pl. Phys. 145: 121-125. Thomas B., D.Vince-Prue. 1997. Photoperiodism in Plants. Academic Press, San Diego. Whitman, C.M., R.D. Heins, A.C. Cameron, and W.H. Carlson. 1997. Cold treatment and forcing temperature influence flowering of Campanula carpatica ‘Blue Clips”. HortScience 32(5): 861-865. Yuan, M., W.H. Carison, R.D. Heins, and AC. Cameron. 1998. Effect of forcing temperature on time to flower of Coreopsis grandiflora, Gaillardia xgrandiflora, Leucanthemum xsuperbum, and Rudbeckia fulgida. HortScience 33(4): 663-667. 115 :93“... .o: EoEoSmmos. N 5.3. Qwv N: 0.: 0.: ©.N_. mmm..-mmmv -- -- l I .. N: mam TE? .38. N.E._oE. 559:. Em: 2.8 8263.. mdm vdw mdw odN odw o. ..N mam ....wmmr WON 5.2 5.9. 53 $5.. 5.9 mam 7an.. AOL 0:55. 8:33 85.9an9 5m. omm.m>< m. N. m o m o .o. m a 9.8;. 5.398 28 38> E058... 3.8 .0 8.5.8 comm .83 .955... 83.x. 8.8. 6.89.05 .0 @5526: 5 98 8965 o. .8588 .o 28 E0: 228E. Em: 2.8 ncm 8.3.9an2 ..m 2.8 89624 N aim... 116 83.326: 9:3... 02 . 3.0.52.3... .5... .... 2.... 858.2: ..3 . v. 5 You You mdm New ... 5 You 3.. I m. E I wdm 3.8 Now v. E 0 33.333 0.0... .. 3.5 Dom 3.0m EON 3.5 m. I I I 3.3. mdw mdm .-- 0 33-30. .00. 350.0. 32.33 3.3.3.3923. ...3 33.3.33. .2 ..N or 3 m. up 0.. .5 8.388... 0.. m a 832. 33> 30:80.03 :03I3 .8:3 .053... 333.5. @333. 3.33.205 .0 35.326: .0 3.8 3m3.3>3 0. 23.3833 :0 3.8 E0: 03.3.3.33E3. ..3 3m3.3>< .3 3.33... 117 Table 9. Actual temperatures of houses for the time of forcing for three developmental stages: Transplant to Visible Bud, Visible Bud to Flower, and T_rsplant to Flowr Develntal Stage (day)s *— Temperature Set POW (°C) Transplant to Visible Bud to Transplant to Visible Bud Flower Flower 14 15.1 16.7 15.9 17 18.2 18.6 18.4 20 20.6 21.3 21.0 23 23.8 23.3 23.6 26 26.8 27.2 27.0 29 29.7 30.0 29.9 118 2.32.8.8. .50... .e .3... w n. .3 2855.3 ...Otm patcmuw ~ 2:...33 30: 3.3 3335.3 3 3......333 3.9.3. 3 3333.3- 3... 0. .:3.03:3.... 2.33.3 333 N... Tune. 3 333... mflmmd n 3mg..- >>..n. 0. m> 233.3 3.... 3.3 mem... H 339.3 N 333... a 3333..- m> 0. .:3.:3:3.._. 0.8:. a: .3 3. 3.330 60...... ..0. 330.3 ..3. 3383.5 333.3 .3.:3an.3>3o .3330. 3>33.33.333 3:3 35 3.2.3.3383. 3333 3.3.3.30 0. 333: 3.32, 330.3 3:3 3323.5 ..e..__m.. 3:33.. 3.333. 3.33.6.3 .6. .25.. 3:526: o. .5329. 9.3 .2... 6. m>. 35.332. a. ..3 33.3... ..m>. as. 33.33 0. .:3_a3:3.. E0.. 333.30... .0 3.3. 0. 3.0.3.30E3. 350.0. 35.3.3. 3.3>_3:3 :0.333.33. .335. .0 3.3.3E3.3a .3. 3.33... 119 Table 11. The difference between day and night temperature (DIF) of plants of Stokesia Iaevis ‘Klaus Jelitto’ grown under different forcing temperatures and three developmental stages: from transplant to visible bud (VB), from VB to Temperature Transplant to B - Trspl setting (°C) FLW 14 0.1 1 2.03 1.08 17 1 .38 1.95 1 .69 20 0.69 1.37 1 .03 23 0.71 1 .04 0.88 26 0.46 0.97 0.76 29 0.32 0.67 0.50 120 100 1 3 5 80 3 h 3 6 1:: 60 4 3 3 £9 5 40 « 3 8 3 0- 20 . 0 - 120 . ._ 100 1 ‘3’ (:3 80 1 . .9 3 60 > 3 D 40 . 20 1 B 0 T r . 1001 8 801 .0 .9 .o '(7) 60 -1 ">' .9 3 40 4 >. 8 20 . . C 0 a . _¢ 5 100 3 . 8 9 801 3 3 3 . . . 3 L o E E 40 i . g L”QNS w 20 . . .. , _ 3‘ o D 0 - ¢ . 0 3 6 9 12 15 Weeks at 5 °C Figure 12. Effects of van'ous duations of cold treatment at 5 °C on percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in WaIaGA'S‘lGaus Jelitto’ in Year 1 (O) and Yea 2 (I). Enor bars represent 95% confidence intervals. L = linear, Q = quadratic trends. he"'vaonsigniliwar't or significant at Pg 0.01 or 0.001, respectively. 121 40 ~ U) 8 3O 4 0 C “5 53 20 < .0 E A 3 I. O 2 + numberoffinal nodes Year1.L o 10 4 7 +numberoffinalnodesYear2.L '0‘ numberofnewnodesYear 2.1. A 0 a 25 4 + Year1 3 + Year2 8 20 . .D 3 c3: 15 . c “5 g 10 . E 3 Z 5 . O 60 « 15? L’“o"$Year1 : 4o . . . . .c .Q" 3 I 20 « C O 3 6 9 1 2 1 5 Weeks at 5 °C 01» OD Figue 13. Effects of cold duration at 5 °C on nunber of nodes (A), number of flower buds (B), and plart height (C) in Sdm’a lavis‘Klaus Jelitto' in Year 1 (O) and Year 2 (I). Enor bars represent 95% confidence intervals. L = linear, Q = quadratic trends. "31"?“Nonsigrificant or sigrificant at P5 0.05, 0.01, or 0.001, respectively. 122 T - 3 '// 100« U) 5 213 75. 3 O G: 8') 501 E C 3 8 25. d) a 0 150. 1254., .. . +Year1,L 3— m “3’ too. ,+—.I.,—,Year2,L. O c 2 75. I (0 >5 8 so- / 5 32/ fl - - - .//. / V/r 120~ . .. . 3 +Year1 0 +Year2,L .D 95.. 2 5 (D ‘5 70-4 2 U) 45. ., 3‘ I 0 20V . ._ / / h / L L j 1/4c / g 80- ,. .. U .2 +Year1 .9 +Year2,L D .0 2 :9 40-1 .9 > I 5 .1: 2“ 0) € D A L 1 #g l l A %//l D 0 f 1 r 1 1 l t 17/v 81012141618 20 22 24 Nl Photoperiod (h) Figure 14. Effects of various photoperiods without a cold treatment on percentage flowen'ng (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Stdres'a Iaevis ‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Enor bars represent 95% confidence intervals. L = linear, Q = quadratic trends. '-"1"Nonsignificant or significant at Pg 0.05, 0.01, or 0.001, respectively. 123 1004 ,. .‘e” '- 5 o 3 754 o c 3 g 50- C 8 25 “- 1 if A 150‘ ' ' 'mno +Year1,LQ ‘- 125, _ +Year2 ‘é’ o 100- c .9. 3 75« 6‘ i 0 50. / B / / / / --....-,,./ 1201 . , '. 4.- 3 95* ,. .+Ye.ar2. . 2 :9 U) .5 7o. 2 g 45. O 202/ 6. ,jy / C/ b / 3 Alf.— g’ 80‘ . 8 +Year1,L"'Q' o .. :5 60‘ +Year2,Q : 3 9 401 '3 W H E > 5 t 20. 0) >5 3 O O +//1 81012141618 20 22 24 Nl Photoperiod (h) Figure 15. Effects of various photoperiods after a 15 week cold treatment at 5 °C on percentage flowering (A), days to flower (B), days to visible bud (C), and days from visible bud to flower (D) in Stdres’a IaeviS‘Klaus Jelitto’ in Year 1 (I) and Year 2 (I). Error bars represent 95% confidence intervals. L = linear, Q = quadratic trends. m""'“Nombnifimm or significant at P5 0.05, 0.01, or 0.001, respectively. 124 *1 j I I I fi/J/l—fi 50 1 . ~0- Oweek cold, Year 1. We” ----- 8 -13— Oweek cold, Year 2, Co"3 E 40 1 1 +~ ~ ~15weekcold,vear1,~L°"o'” ~ a I- 15 week cold. Year 2. Ll‘so'ts g 30 , . .. . . . “5 5 20 « .. . ..D E 3 10 - J. 0 1 1 1 1 1 1 1 1 7/r’1 B 3 4° ‘ '-o— ' Ovlreekcdd,Year1,1.”‘30’NS 3 —1:1- Oweek cold, Year 2. 00"“ c 30 1A » 1 -+ 15weekcold,Year1,-L°4°707"- - ~ 5 + 15 week cold, Year1,L"sQ"s "5 201 ...... g E E 10 . . .. , . .2 1:1 0 ‘” 1 [512 m C 3 -o— 0 week cold,Year1, Rio"5 5 15 -Cl— Oweek cold,Year2,L9'“ . 3 4 " +” " 15'week'eeld.veer1. L Q” " 3 + 15 week cold,Year 2, L'Q'“ “— § 3 5 1 ....... . 4 _ z 0 fi fi 7/r/¢ 60 . E 501 . . . 3 .9 2 30. . .. ..." .. —O— Oweek cold,Year1, L Q s E“ 20 1 --C}—wOweekcold,Year2,L"lem-‘ ~ .1 0- + 15 week cold, Year1.L°"Q"s 10 J " ‘ ' " + 15we‘ekcold,'Year2.L”‘Q'” “ 0 1 1 1 1 1 1 1 1 #41 8 10 12 14 16 18 20 22 24 Nl Photoperiod (h) Figure 16. Effects of various photoperiods with and without a 15-week cold treatment at 5 °C on number offinal nodes (A), number of new nodes (8), number of flower buds (C ), and plant height (D) in Stokesia Iaevis ‘Klaus Jelitto' in Year 1 (O) and Year 2 (I). Error bar'slepresent 95% confidence intervals. L = linear; Q = quadratic trends. ”5" ' Nonsignificant or significant at Ps 0.05, 0.01, or 0.001, respectively. 125 .u. use .m .3 case 5 mos. c2898. Ems. “5259.8 ho £89569 o5 2m 0 new .m .< magma 5 mos. c0685? unfit—wad on... .o_. 2an 5 8.583 9m 3:: 5.3662 .85. co newsman 9: ..ee__e_. 85. mg when“ 5 E .w .9 852a... c539 3652a ho 39 can 8 .m .$ 9 9:: :0 239852 ho flomtm .5 9:9”. Aoev QBEoQEmh on mm mm VN NN ON 2. 0.. 3 we on mm mm VN NN ON 2. er E. N... on mm mm vm NN ow 2. we 3 NP 00.0 F r L h b P d % J1 1 _ 4 q q a 4 d u 1 1 q q 4 d d d d u d 1 No.0 1 v0.0 .626: Egon. , . m .030: 9 van 0333 Eoc £60 < 25. @3332 960 row row 1on Tow 10m 1:18 You f8 skep/i 126 $1120 30 1 25 1 2O 1 15 1 Number of nodes 10 , , . ., + Number of final nodes 5 + Number of new nodes 0 1 11 351 301 251 201 15 1 10 1 Number of flower buds 5 1 0 60 1 50 1 4o- 30 1 Plant height (cm) 20 1 10 1 0 1 1 1 1 1 1 1 1 14 16 18 20 22 24 26 28 30 32 Temperature (°C) Figure 18. Effects of temperature on number of nodes (A), number of flower buds (B), and plant heg'l‘flow) in Sdm'alm's‘Klaus Jelitto’. L=linear; Q=quadratic trends. “5" 1 nsignificant or significant at Pg 0.05.0.01, or 0.001, respectively. 127 Appendix A: Comparison of Stokesia Iaevis Cultivars Grown under Two Different Lamp Types for Daylength Extension 128 Introduction Stokesia Iaevis (Hill) Greene is an herbaceous perennial native to the southeastern United States from South Carolina to Louisiana, including the panhandle of Florida (Gunn and White, 1974). With its large (=10 cm diameter), lavender-blue flowers, Stokesia Iaevis ‘Klaus Jelitto’ demonstrated good potential as a potted flowering plant in preliminary studies and forcing experiments (Frane, 1999; chapter 4). ‘Klaus Jelitto’ has a facultative requirement for vemalization and is best described as a facultative intermediate-day plant (chapter 4). Several cultivars of S. Iaevis exist including yellow-flowered ‘Mary Gregory’, white-flowered ‘Alba’ and ‘Silver Moon’, and ‘Purple Parasols’ which has violet flowers. ‘Klaus Jelitto’, ‘Blue Danube’, ‘Wyoming’, ‘Blue Star’ and ‘Omega Skyrocket’ all have lavender-blue flowers, as does the straight species (Armitage, 1989; Fischer, 1998). Based on the potential observed for ‘Klaus Jelitto’, we were interested in investigating the flowering characteristics of other cultivars of Stokesia Iaevis in container production. Different lamp types used for day-extension were found to influence flowering characteristics of S. Iaevis ‘Klaus Jelitto’. ‘Klaus Jelitto’ plants cooled for 15 weeks and subsequently grown under a 16—h photoperiod with day extension from HPS lamps produced approximately two times the number of flowers produced on plants grown under the INC lamps. In addition, plants of ‘Klaus Jelitto’ without a cold treatment were found to flower at 50-60% under a 16-h photoperiod when HPS lamps were used for day-extension when no plants 129 ..-—4 mtg-“7 . r flowered without cold treatment under a 16-h photoperiod when INC lamps were used for day-extension. It was estimated that a 16-h photoperiod provided by day-extension lighting with incandescent (INC) lamps provided a daily light integral that was 78% of the daily light integral provided by day-extension with high-pressure sodium lamps (see chapters 1 - 3 for explanation of estimation). Warrington and Norton (1991) found that increasing the daily light integral increased the rate of floral development and number of flower buds in Dendranthema xgrandiflorum (Ramat.) Kitamura, Raphanus sativus L., Zea mays L., and Cucumis sativus L. High-pressure sodium and INC lamps also differ in spectral quality as well as intensity. Incandescent lamps emit a mixture of red and far-red light which is richer in far-red than red (ratio of red 600-700 nm to far-red 700-800 nm < one [Lane et al., 1965]). High pressure sodium lamps emit a mixture of red and far- red light with far less far-red than red light compared to incandescent lamps (T himijan and Heins, 1983). The objectives of this study were to (1) determine flowering characteristics (eg. time to flower, flower number, final plant height) and (2) determine the effect of using different lamp types for day-extension lighting for S. Iaevis ‘Alba’, ‘Blue Danube’, ‘Klaus Jelitto’, ‘Mary Gregory’, and ‘Purple Parasols’. 130 Protocol Experimental Design: completely randomized design Cultivars: ‘Alba’, ‘Blue Danube’, ‘Klaus Jelitto’, ‘Mary Gregory’, ‘Purple Parasols’, and ‘Wyoming’ See Table 12 for starting plant material, dates of arrival, cold treatment, and forcing, as well as average daily temperature and average daily light integral during forcing. Daily light integral data were only recorded for plants under day extension with HPS lamps. Cold Treatment: 15 weeks at 5°C Photoperiod and lamp types: All cultivars were treated with a 16-h photoperiod composed of 9 h of natural light plus a 7-h daylength extension. Natural sunlight was supplemented with light from high-pressure sodium lamps. The high- pressure sodium lamps were turned on every morning at 0800 when natural light (measured on outside of greenhouse roof) fell below 400 umol-mz-sec“. High- pressure sodium lamps were terminated when ambient light level was greater than 400 ,LLMOI'mZ'SeC'1 and were not re-initiated until ambient light level fell below 200 pmol-mz-sec". Daylength extension was given with either: 1) incandescent lamps (INC) (1-3 umol-mz-sec‘) 2) high pressure sodium lamps (HPS) (4O pmol-mz-seC") 131 RESEARCH PROTOCOL: Upon arrival at MSU, plants were held under natural daylengths at 20 °C until the experiment began. Bare-root plants were immediately transplanted into bulbtrays and the root masses were covered with soilless media. Plants were cooled at 5 °C either in bulbtrays, plug trays, or 2.5-inch pots. Following cold treatment, plants were potted into 13-cm pots and placed in 20 °C greenhouses under a 16-h photoperiod delivered by daylength extension from either INC or HPS lamps. Data collected included: 1) Initial node count 2) Date of visible bud 3) Date of flowering 4) Number of flower buds at first flower 5) Final node number beneath the first flower 6) Plant height at first flower 7) Number of new nodes were calculated by subtracting the number of initial nodes from the final nodes. 8) Days from visible bud to flower were calculated by subtracting the days to visible bud from days to flower 132 Summag of Results In general, the type of lamp used for daylength extension had little effect on flower timing for cultivars of Stokesia Iaevis. There were few statistical differences for flower timing, and where statistical significance did occur, the differences were small (Table 13). For example, ‘Purple Parasols' grown under INC lamps flowered approximately 5 days faster than plants grown under HPS IF 1 lamps. Lamp type influenced plant height of only two cultivars: ‘Klaus Jelitto’ ‘ I was approximately 6 cm taller under HPS lamps compared to INC lamps while , ‘Purple Parasols’ was 11 cm taller under INC lamps compared to HPS lamps. g5 Lamp type had a stronger influence on number of flower buds. Bud number increased 1.5 to 2 times for ‘Blue Danube', ‘Klaus Jelitto’ and ‘Purple Parasols’ grown under HPS lamps compared to INC lamps. Flowering percentage varied greatly (20 to 100%) between cultivars. ‘Wyoming’ and ‘Alba’ had particularly low flowering percentages 20 to 60% and 30 to 60%, respectively, for ‘Wyoming’ and ‘Alba’,. “Purple Parasols’ had a moderate flowering percentage of 70 to 80%. The lower flower percentages observed for these cultivars may have been related to small initial plant size. Plants of ‘Alba‘ and ‘Purple Parasols’ arrived in a 72-cell tray which were considerably smaller than the starting material of most of the other cultivars (2.5- inch pots). Starting material for ‘Wyoming’ was single-eye bare-root divisions with no more than 6 leaves. 133 Although most cultivars flowered in between 61 to 85 days, had 6 to 13 flower buds and were between 39 and 52 cm tall, there were several exceptions. ‘Alba’ was the slowest cultivar to flower (> 100 days). ‘Mary Gregory’ and ‘Wyoming’ attained heights at first flower of approximately 27 and 29 cm (when data for each cultivar were pooled for both lamp types due to lack of significance) and were much shorter than other cultivars. Starting plant material size did not appear to influence flower number since ‘Alba’ and ‘Purple Parasols’, which had some of the smallest initial plant sizes had considerably more flowers than other cultivars. Arrnitage (1989) observed that ‘Alba’ was less floriferous than other cultivars in garden performance, but we observed the opposite when container produced. Potential as a potted flowering plant varied widely among cultivars. “Klaus Jelitto’ and ‘Purple Parasols’ showed the greatest potential with numerous large flowers of either light lavender-blue for ‘Klaus Jelitto’ or deep violet for ‘Purple Parasols’. In addition, ‘Klaus Jelitto’ had an attractive branching habit. ‘Mary Gregory’ showed some potential with its naturally short height and small dark green leaves but produced only a few small yellow flowers. ‘Alba’, ‘Wyoming‘, and ‘Blue Danube’ lacked potential for various reasons: ‘Alba’ and ‘Wyoming’ had very low flowering percentages and for ‘Alba’ the time needed to produce a flowering plant was excessively long, while ‘Blue Danube’ was spindly, had few flowers, and tended to fall over. 134 Literature Cited An'nitage, AM. 1989. Herbaceous Perennial Plants. Varsity Press, Athens, GA. Fischer, T. 1998. Stoke’s Aster— Stokesia Iaevis. Horticulture 95:34. Frane, A.J. 1999. Temperature effects on timing and bud development of Coreopsis verticillata ‘Moonbeam’ and flower induction of long-day perennials under different night temperatures. MS Thesis, Michigan State Univ., East Lansing. Gunn, CR. and GA. White. 1974. Stokesia Iaevis: Taxonomy and Economic Value. Econ. Bot. 28: 130-135. Lane, H.H. Cathey, and L.Evans. 1965. The dependence of flowering in several long-day plants on the spectral composition of light extending the photoperiod. Amer. J. Bot. 52: 1006-1014. Thimijan, R. and RD. Heins. 1983. Photometric, radiometric, and quantum light units of measure: A review of procedures for in interconversion. HortScience.18: 818-822. Warrington, I.J. and RA. Norton. 1991. An evaluation of plant growth and development under various daily quantum integrals. J. Amer. Soc. Hort. Sci.116(3): 544-551. 135 .386. an... E268 2:80.959; 5.? 8.22.6 59.2.66 .625 $505 9cm... .0. >30 606.com. 0.83 Emu Em... .uofiEfic 6.03 RES .025 E8333... 5.? 5.32.6 5953...... .09.: .565 3cm... .0. £652... Em: ....mu mom.o><~ 136 F. F F EN we: 82 8e. 82 N... 08 oz. .80 2 .98 N .96 N m 89-28 8.59.2. v.2 98 we: 82 82 82 e685. no. 98 oz. as. o. dew NN dew N. o. ..8-NN 29:... .11.: m8 we: 8e. 82 5.8.0 N. : 8.8 oz. ...e.. .N .80 N -- o. ..mN be... 82 82 82 3; m.oN we: 5.. 8 .80 N .92 .N 2%.. m. 3 ...ON oz. .8”. 3 .80 8 .80 S o. ......N Sec. 0: N.oN we... 82 89 82.8 N. 2 08 oz. 5.. MN .80 N -- o. ..mN 2.5 Ne... 08 we: 82 82 82 ~N.N. 3N oz. 5.. 2 .98 mN den 3 o. __8-NN 8.... .38 52.8.. .o... 669:. c.3935: comma comma .5...ng .mcmfiE Em: 2.2. 2.66 8.: 9.6.0. ...oEfio: _m>...m .2. 3:63 :53 omn.o>< 69263.. 95.. 9mm. Eco 960 .o 9.5 No .02 9.2.65 62:30 maEm_ 28338:. u 02. .maEE anom c.3832. gm... u we... .mcotoaotoca ... or .0. 5.9.9.6 565:3“. 52.3. 0. com: QES .o cab .1. cab aEm. ..mc.Eo>>>. ucm .m_omm.mn_ 8.9:... $53.0 >52. 305.63 mam? ..3:ch 62m. ..m£<. MSmm. 6.69.05 .0. .653... Em: 2.2. omm.o>m ccm 6.39882 2.8 mom.m>m mm =95 mm 6582 50.63.... 2.958.. Eco ._m>...m .0 mean :5...th .oa 3cm... .0 .mnEsc ”3.58:. Ema 95.5% .N.. ofiw... vm ..v N 00 :10 mm om ma... 0.000.00 . . N0 0 :0 00 00 ON oz. 0...... 00:00:...05 wz wz wz wz wz wz 0 NN 0 ..0 00 0N 0.: 00: 280.0 N 0N N .N 00 0... 8 oz. .005. . . 02 02 . 0z mocmoECQw m .. mm N K c: on 00 P 9.1 00:00 0 00 0 :1 mm mm 00 oz. 0:02 : wz wz wz : : wOCNOECQm a N0 F .. 5 mm mm 0:: ma... 002.00 0 00 F F 00 00 on 00: 02. 0:.m mz wz . mz wz wz QOCSECQW mm mm m 00.. N0 N0 00 mm: 3 Se 3 :N: 00 mm 00 oz. 00.< 00:0 3.0. 00>00. .030: 0: 0:0 .030: .090... 30: .030: 0:0 0.0.0.> 03.03 000030.". :0 .02 0.0.0 :0 .00502 0: 0>0o E0: 0>0D 0: 0>0D €00.00 00>... 0E0.. 02:30 .030: 0.: 0.: 500000 00>00. 30.. :0 .0080: 0... 0. 0.0:0. 00>00. 30.. :0 00:52 00>. 00.0. .0: 0000000500 02.3000... 010:0. 00:00:...90 :0 0.000.. .00 m :0 03.003 9 5.3 00:00.. 0.03 0.02.30 __< .0000. .02.: 200000000... .0 0080. $0.... c.2000 05000.0 00.... .0020 5.3 00.000.00-00 >0 00.02.00 0000002000 ... 0: .000: O. on :0 030.0 0.300. 0.09.05 :0 0.02.30 .0.0>00 :0 00.000.00.000 00:030.“. .2030... 137 2.0260000. .50... .0 .50 0000 n. .0 .000:.0m.0 .0 20005090002 .........02 000005005 wz m2 wz wz wz wz 0 mm m NN mm Nm ow ma... 0 N0 N 00 mm 00 8 02_ 0559.3 000005006 CC 0:000 m: 0.00... 138 Appendix B: Effects of Cold Treatment, Phototperiod, and Forcing Temperature in Anemone hupehensis, Penstemon digitalis ‘Husker Red’, and Saxifraga arendsii ‘Triumph’ 139 ‘ . .f...‘ u. . ’ .'_A'¢t-W.'.u '7'. n—qg_ Introduction Flowering of many temperate herbaceous perennials is influenced by four main factors: juvenility, vemalization, photoperiod, and forcing temperature. All of these factors influence both the quantity and quality of flowering. Juvenility is a developmental plant state during which plants are unable or only poorly able to perceive inductive conditions such as vemalization and photoperiod. Vemalization is the promotion of flowering by cold treatment, while photoperiodism is the control of flowering by daylength. Plants can have obligate or facultative responses to vemalization and photoperiod. Forcing temperature influences flower timing as well as plant quality features such as plant height, flower number, and flower diameter. Vemalization and photoperiod also influence plant quality features such as plant height and flower number. As herbaceous perennials have gained popularity during the past 15 years, the interest in producing them has also increased. Since consumers prefer to buy plants that are in flower, and most herbaceous perennials are not in flower in May and early June when consumers are doing most of their shopping, producers are interested in the flowering requriements of these species so that they can produce flowering plants in May and June. The objective of these experiments was to determine the influence of vemalization, photoperiod, and forcing temperature on three herbaceous perennial species: Anemone hupehensis, Penstemon digitalis ‘Husker Red’, and 140 Saxifraga arendsii ‘Triumph’. Information from these experiments can be used to create production schedules. 141 Protocol (1) Cold Duration Experiment Experimental Design: completely randomized design Cold treatment: 0, 3, 6, 9, 12, and 15 weeks at 5 °C. Research protocol: 10 plants were immediately potted into 13-cm pots with soilless medium and placed in a 20 ° C greenhouse where they were provided a 16-h photoperiod delivered by a natural day supplemented and day extension from high-pressure sodium lamps. Plants were cooled in the plug tray in coolers set at 5 ° C. Upon completion of the desired cold treatment duration, 10 plants were transferred to the greenhouse, potted into the 13-cm pots and placed in the same greenhouse as those plants which did not receive a cold treatment. Data collected: 1) initial node count (Year 2 only) 2) date of visible bud 3) date of first open flower 4) number of flower buds at flower 5) plant height at flower 6) number of nodes to either last visible leaf or beneath the first flower (2) Photoperiod Experiment Experimental design: completely randomized design Cold treatments: 0 or 15 weeks at 5 °C. 142 Photoperiods: 10, 12, 13, 14, 16, 24, and a 4-h night interruption from 10:00 pm. to 2:00 am. Research Protocol: Half of the plants were immediately placed in coolers set at 5 °C for 15 weeks. The remaining half were potted into 13-cm pots and placed in a 20 °C greenhouse and ten plants were placed under each photoperiod. Plants were forced for 15 weeks after which plants not in visible bud were discarded. Photoperiods were provided by a 9-h natural day and day extension from incandescent lamps to the desired photoperiod. Between 8:00 am. and 5:00 pm. supplemental light light from high-pressure sodium lamps was given when the ambient photosynthetic photon flux (PPF) first fell below 400 umol-m‘ 2sec“. High-pressure sodium lamps were turned off once PPF rose above 400 /.cmol-m'2-seC'1 and did not turn on again until ambient PPF fell to 200 umol-m' 2sec“. After 15 weeks of cold treatment, the other half of the plants were moved to the greenhouse, potted into the 13-cm and placed in the photoperiod treatments exactly as those without a cold treatment. Data collected: same as in Cold Duration Experiment. (3) Forcing Temperature Experiment. Experimental design: completely randomized design Cold Treatment: 15 weeks at 5 ° C. Temperatures: 14, 17, 20, 23, 26, and 29 ° C (constant day and night temperature). 143 Research protocol: All plants were given a 15 week cold treatment at 5 ° C. Following cold treatment, ten plants were potted into 13-cm pots with soilless medium and placed in each greenhouse set at one of the six temperatures. Plants were provided with a continuous 16-h photoperiod using high-pressure sodium lamps. Data collected: same as for Cold Duration Experiment and Photoperiod Experiment. 144 Anemone hupehensis Percent Flowering @100 « . ., . .E / I- o 80 ~» . . , . . . 4 3 .9 50 .. . . , u. ‘0‘ 4o .. ¢ § 20 q» o. 0 : l : l . 0 6 9 12 15 Weeks at 50 Flower Timing Plant Node Development 150 J» . . . . , l i - l 3 6 ; A125T l 1, 0 0 y 0 5.100 + . . . ,. ,. . 5 'o ,‘ o 40 : 75 J." '/.\'+.\. 3 : 0 l 25 + -I— Days tovisible bud g; 3 + Days to flowe I i 0 ‘gi l l l l : 5 0 i i i i l * 03691215 03691215 Weeks at SC Weeks at 50 Number of Flower Buds Plant Height at Flower 100; , . .1 8° . 7 ’ i l . 7%. . . ..l -501... 5 l l 5 l i 50 - . . , ,_ . . . ( ... . 25 + , . . , , , . l ,3 l 0 * :l: 20 + ; l -25 '— 1 . l 4. w‘ l: 0 l: : t l : .L i ' 0 3 6 9 12 15 0 3 6 9 12 15 Weeks at so Weeks at 5C Figure 19. Effects of cold treatment in Anemone hupehensis in Year 1. Anemone hupehensis Percent Flowering 01001 .E '5 80 ~. » 3 2 60 ... ll. ‘0' 4o 1*“ . 0 § 20 4- a. 0 1 + % 1 J, , 0 3 6 9 12 15 Weeksatsc Flower Timing Plant Node Development . F j l “0257 '-.'—'Newleavesbeneaththeflower : 13 ‘ —.— Finalleavesbeneaththeflmver .... : 020i- .. g1 ,_ i I: l 0 ._ l P o . +Daystovislblebud l a 5+ .. - -- - .. .. 25 +- . , . . z . I +Daystoflower l ‘ 04 I: 4. : 4 %%J 0 r . % 4 § 11 03691215 03691215 Weeks at 50 Weeks at SC Number of Flower Buds Plant Height at Flower 100; . «- _ . .. l 80+ I i - 607 A - ~ ~ - , i .5. 40 + W‘ “E 204- . . . .. ,, . § .3 0% JL . i I '20 T + : it 0 0 O 4 4. + : : t ‘ O 3 6 9 12 15 0 3 6 9 12 15 Weeks at SC Weeks at 5C _ “a“ ‘JW‘. 0‘ We... , Figure 20. Effects of cold treatment in Anemone hupehensis in Year 2. Error bars show 95% confidence intervals. 146 Anemone hupehensis Flower Timing Flower Timing Oweekatsc 15weekat5C 175 F P .. ‘25 ‘ A 100 ‘0” 1; m . 'B > 100 1 I a , .. 8°: 3 3 75 ‘ 60 3 2 0 2 5° 1 4° 15 C l- 2 I- § . a 0 - 0 C 81012141618202224Nl n. 81012141618202224Nl l Photoperiod (h) + DW0mmm‘fi’hotoperlod (h) ... WDWM ... mum + Dantolbwer + Peroentllowemg +Peroentl'lowemg Plant Node Development Plant "“9 ”"9“"th 3 3 so 8 80 h\. .8 4o . . C '5 6° 1 H u: 30 V . g ‘0 ‘ I- 20 5 2° ‘ '5 10 3 0 3 z z 0 81012141618202224Nl 81012141618202224Nl Photoperiod (h) Photoperiod (h) Number of Flowers Plant Height at First Flower fl 3 80 I - . .g A 40 z i l E _ a 50 ‘1 l g 30 ”5 j i 3:" 20 I. 40 ' l 9 3 20 J ', ; :2 10 § 0* f o z 81012141618202224N, 81012141618202224Nl ‘ Photoperiod (h) + °W°°'d Photoperiod (h) -O- 0mm“ + lSweekoold + 15weekoold Figure 21. Effects of various photoperiods with and without a 15 week cold treatment at 5 °C in Anemone hupehensis in Year 1. 147 Anemone hupehensis Flower Timing Flower Timing Oweekatsc 15weekatsc 175 , 175 , ,3 150 , 100.;’ a 150, A 100;D 5‘ 1251 80 3 a. 125, so 3 3 m} 60 5 3 1°01 5 so :3 a 75 C O 75 . .§ so 4° ‘e .5 so. 5 ‘° 5 l. 251 20 3 l- 25. 2° 2 o I. . o 3 oi ... 0 g 31012141513202224m 0- 81012141618202224Nl + mfiwflafg‘gtoperiod (h) _._ mm Wwotoperiod (h) + Deystoflower -I- WNW + Porositllowerlng + Pmflaverlng Plant Node Development Plant Node Development 0 week cold 15 weeks cold 8 so 8 5° U ‘ + Final nodes atl’lower ‘D + F“"°"°‘"“°"' 2 50‘ +Newleavesatflower g ‘0‘ ‘ +Mmmm I. 30‘ ' h 20 _, Eli-N é . E 10 . H. . . . E 10 . be . . :1 o - - a o - - . . z 81012141618202224Nl z 81012141618202224Nl Photoperiod (h) Photoperiod (h) Number of Flowers Plant Height at First Flower g 80 l g ‘ A 60 g 60 l g 453. Q. , . § “1 in :3 0 2o 0 § 0 1 1 1o . "e + Oweekcold Photo riod h Photoperiod 0') + 15-..... P° ‘ l 3; 3:33:33, Figure 22. Effects of various photoperiods with and without a 15 week oold treatment at 5 °C in Anemone hupehensis in Year 2. Error bars show 95% confidence intervals. 148 pinch-H3105} ll- . .. .% .0.0>.0.0. 00:00:08 $00 0000.00. 0.00 .0..m. 0.05.35; 0. £0.00 000030: 0.030. 000.020 :0 0.0. 000 3.0.0.: 0E: 000030.: 00 0.0.0.00E0. 00.0.0: :0 0000::0. .00 0.00.”. .0... 0. 0.0.00E0P on mm mm 0N NN ON 0: 0w 3. N: on 0N 0N 0N NN ON 0.. 0: 3. N: on mm mm 0N NN 0m 0: 0.. 0.. NF F L P 4 0 — l1 a d T L. T. . .1 0 L. 0 . .J . cod 0. .0.,0: 0. 0, . , : ,: .. .. : ..vod rNo.0 :0: 0 0 . : moo . 00.0 0:: ...mod Anemone hupehensis 1‘ lfil 1‘ 1.! D II 5...... . . .. -8. . 00.. . 00m o .030: 0. 0.00 m .030: 0. 0:0 0.0.03 .00.: 0.00 < 000 0.0.0.> 0. 0.00 00m sKep/L 149 91190 Anemone hupehensis J .. + Number of final nodes A 70 + Number of new nodes ‘f' Number of nodes Number of flower buds 254 80 80a Plant height (cm) 8 14 {e {a 2}) £2 2} 2% 23 Temperature (°C) Figure 24. Influence of forcing temperature on number of nodes (A), number of flower buds (B), and plant height at first flower (C) in Anemone hupehensis. Error bars represent 95% confidence intervals. 150 Penstemon digitalis 'Husker Red' 151 Percent Flowering 0 80 q. .: 5 E 50 - , : ‘0' 4o .. 0 2 20 .. 0 0 l 1 4 .0 + + o 3 6 9 12 15 Weeks at SC . Plant Node Development F'°‘"°' Timing Nodes to the last visible leaf 150 + » ~ ~ l 0 ’ i 125 l + Days to visible bud 3 30 4 I a T + Days to flower ' ' 8 $1001“- -- __.._ '~ :20. . , .. w 0 ° T O ‘ g : EsoT E10... i 25 . a l T 2 . l 0 L + . g : 4‘ + 0 + 1: 54 l I: : if] 0 3 6 9 12 15 0 3 5 9 12 15 Weeks at 5c Weeks at 50 T! 3 Number of Flower Buds Plant “9'9ht at Flower .C C .4. - < 300 80 ‘ MT ~ 250 ._ l N 32001. E 60, ' " 'l a 9, . , e 150 + E 40 .L. . . = g 100 J» .9 . * a £5 20 .. . E 50 -- E T g 0 I. i i : F : 0 1* l J . . . l l g 0 3 6 9 12 15 o 3 6 9 12 15 2 W09“ at 53 Weeks at 50 Figure 25. Effects of cold treatment in Penstemon digitalis 'Husker Red' in Year 1. Penstemon digitalis 'Husker Red' Percent Flowering E100 .. h 0 80 ._ 5 E 6" ‘" E 40 T 0 2 20 .. a 0 a i .L r : . 0 3 6 9 12 15 Weeks at SC Plant Node Development How" 11me Nodes beneath the first flower 15°? ' 1 «so. +~°d~wmwwwm 125 l‘ _ + Days to Visible bud . 8 -I- WW E + Daystoflower l g o 5 E ‘é’ 10 .. 3 . z 0 L : , g , § , 0 i i J. i i If: 0 3 6 9 12 15 0 3 6 9 12 15 18 fig Number of Flower Buds Plant Height at HOW" 2 2 200. '5 150 l m < -~ " g .5. 0 100 4» .. E g .9 2 50 T- a m l g 0 % + : : : l 0 ... 1 1 + 1 1 § i E 03691215 03691215 5 Weeks at 50 Weeks at 5c Figure 26. Effects of cold treatment in Penstemon digitalis 'Husker Red' in Year 2. Error bars show 95% confidence intervals. 152 Penstemon digitalis 'Husker Red' Number of nodes o—on-O——-o. O 51012141618202224Nl Photoperiod (h) Flower Timing Flower Tlmlng OweeketSC 15weekat50 200‘ , 1002 A 701 W A 1002 m 150 1 80 3 > so - 80 3 1:. t L g a 50 1 - - » I 3 1: l reoo 3 4o. hi—‘H , ‘ 60o V 100 c a l- . 20 g l- 10 1 20 2 0 L O n- 0 0 g 81012141618202224Nl 81012141818202224NI Photoperiod (h) Photoperiod (h) -.— Days to visible bud -0- Days to visible bud + Days to flower -.- Daystoflower + Percent flowering + Percent flowering Plant Node Development 15 weeks cold Nodes to the last vlslble leaf 3o . 20 t Number of Flowers Photoperiod (h) +15weekcold g 250 1‘ 100 3 2001 a . . a. .E . O 1 I E 75 0: 1, V . A . r o 150 . s, ‘0- ‘ I 1 "’ 50 1 .1: g 1001 a . ' g g 03‘ ' J o 81012141618202224Nl WK! 1 Plant Height at First Flower 81012141618202224Nl Photoperiod (h) +-.— 15 week cold Figure 27. Effects of various photoperiods with and without a 15-week cold treatment at 5 °C in Penstemon digitalis ‘Husker Red’ in Year 1. 153 Penstemon digitalis 'Husker Red' Flower Timing 0 week at 5 C 200 ... i i100 £1501 »80° re 3 E1001 % "508 E 50 . 40E 0... e 0 a 81012141618202224Nl + Days w mlmotopenod (h) + Days to flower + Percenttlowemg rlnL Flower Timing 15 week at5 C 175 - .. 150; 5 1003 1» A t at 125. 80 it v l P 75‘ 40 5 .§ 5°) PM e F l- 251 We '20 g o“ e 1 0 o 01012141015202224111 0- Photoperlod (h) + Daystovisiblebud + Daystoflower + Percentflowenng Plant Node Development 0 week cold Nodes beneath the flrst flower 401 +Finalnodee. .. .., -Q— 3 produced during forcing Number of nodes 0 81012141618202224N| Photoperiod (h) Number of Flowers Number of flowers 0L 8 10 12 14 16 18 2022 24 Nl -.— 0 week cold ' + 15 week m Photopenod (h) Plant Node Development 15 weeks cold Nodes beneath the first flower ’ 1 30(+,.Finalnodes. . .. ‘J+ Nodesproduceddunngiorcingl 20; l ’ t 1 I o- f 1 01012141610202224N1 Phatgpgdgd (h) 101 Number of nodes Plant Height at First Flower 100 Y i I 75 i 5° ‘1 Height (cm) 25 i r 81012141618202224Nl h _._ 0mm” P otoperiod (h) + 15week cold 0 FigLre 28. Effects of various photoperiods with and without a 15-week cold treatmert at 5 °C in Penstemon dg‘talis ‘i-hsker Red’ in Year 2. Error bars show 95% confidence intervals. 154 S 'tali gl 'Husker Red' i Penstemon d 29.95 cocouccoo $3 E0852 man ..otm ..omm .cxmsz. manage 5.53on E E .w .9 9:950: 9959 @8505 ..o 99 can 8 .m .$ 9:: 9530: no 239952 @588 do 8:25;. .mm 959“. Gay oSfiEan... om mm mm VN mm on or or 3 N_. on mm mm VN mm on or or 3 Nv on mm mm VN NN ON 2 or 3 NF \FL11 1 q . . . 1 o 826: o. 960 m .530: 2 can o_n_m_> ED: 960 < can 927.; 9 £60 ood . No.0 3.0 . mod . mod 2.0 .0? .om .OM ,ov .om cm ,9. om om . cow 5599/) sAea 155 Penstemon digitalis 'Husker Red' I JP Y 50 1 "i' A 40. 30« Number of nodes 10~ 4; Number 01 final nodes ‘ + Number of new nodes 0 1 1 1 J 1 l J Y r v I Y I T 250 4 200 « 1501 100 « Number of flower buds 50‘ 701 c 60. .. 50. 40- 304 Plant height (cm) 201 10. 0 1 1 1 4 1 1 1 1 6 1 8 20 22 24 26 28 30 32 Temperature (°C) Figure 30. Influence of forcing temperature on number of nodes (A), number of flower buds (B), and plant height at first flower (C) in Penstemon digitalis ‘Husker Red'. Error bars represent 95% confidence intervals. 156 Saxifraga arendsii 'Triumph' Percent Flowering Percent Flowering 91215 0 3 Weeks at SC Plant Node Development Flower “me nodes to last visible leaf 35 «*- 3 a 30 1' 30 i 9 3 25 ‘ . TE 1 0 I :1 25 - l 5 20 l , 2:. 20 1 -~ '1 g 15 .1 E 13 1- ” ‘2' 1° 1 '- 5 i+oays to visiblebud ’ i g 5 i . .I ‘ 0 tofl ‘ ' ~ 1 a TT a%ys éower 1 1 £1 0 ' 1 1 1 1 1 14 0 3 6 9 12 15 0 3 6 9 12 15 Weeks at 5C Weeks at 5C 71 'g Number of Flower Buds Plant "9'9ht at Flower .C 3 15 ’ ‘ g 20 -;- ' - ~ 1 a A 7 1 310+ .1 ,. E1, 15,. 0A1 ‘2 V E, 10 ;_ g 5 a}. . , , ., g E 5 * . g 0 4' 1 1 1 1 1 0 . J 1 1 . 1 1 g 0 3 6 9 12 15 0 5 5 0 12 115 2 Weeks at 5C Weeks at 5C in Year 1. 157 Figure 31. Effects of cold treatment in Saxifraga arendsii 'Triumph' Saxifraga arendsii 'Triumph' Percent Flowering E100 1 g 80 1 .2 60 ._ IL ‘5 40 -— 0 g 20 .. a O - 1 1 1 1 0 2 4 6 810121416 WeeksatSC Flower Timing Plant Node Development - _—+4=1na1nedeo . a 120 7" +N'ewnodes 60 + ' 0 1001, . 1 . _ a 1 8 >1 40 __ C 80+ , fl - co- 3 i 8 60 .i- . 0 20 2 e 3 .§ ; g 40 1. .- 0 ++ Daystovisiblebud 1 g 20 T‘ DDaystoflower g 40:11....1 0‘fi11111111 0246810121416 0246810121416 Weeks at 5C Weeks at 50 w '2: Number of Flower Buds Plant Height at Flower .C "‘ 10 c ~— “ 35 ; 3 a 8 "*- A 30 '7' n . 1; W ‘1' E 25 4- .1" , ..-; 1: + . g 3 6 19' 20.; .. 1 g 4* A ~- 2., 15* + 1 ‘1 g 2 _1_ £ 10 -1- ~ ~ ~.' "' L 5* . . . .. J b 4 di- "’ I 301111111 011111Jb1‘L1‘U1' g 0246810121416 0246810121416 2 Weeks at SC Weeks at 5C Figure 32. Effects of cold treatment in Saxifraga arendsii 'Triumph' in Year 2. Error bars show 95% confidence intervals. 158 Saxifraga aren dsii 'Triumph' Flower Timing Flower Timing 0weekat5C 15weekatSC a, 100 .100 5 A 1 1 1000' 3 g 8°? Ni—i A .g . 75 g 5; 60 l » 75 g 50 E g 40 i k , 50 2 fl 1 . 25 5 p 20 i I I 25 E e A_“. : a e 0 1‘5? o ' 1 . 0 3 81012141618202224NI a. 81012141618202224Nl a Photoperiod (h) _._ daysmm‘ebghOtoPefiOd (h) a —l- daystoflower + percentflowenng Plant Node Development Nodes to last visible leaf ' a 8 8 M e h- o 20 . g . . . L 3 to E . a 2 0 81012141818202224N| Photoperiod (h) Number of Flowers Plant Height at First Flower % 6 j E 25 = I 3 20 "5 4 J . E '5 x. . b i 0 9 at 0 , ; - 10 .o 2 ._ . a E F I 5 1 a 0 1 2 ' ' 0 81012141618202224Nl 81012141618202224Nl Photoperiod (h) Photoperiod (h) Figure 33. Effects of various photoperiods with and without a cold treatment in in Saxifraga arendsii 'Triumph' in Year 1. 159 Saxifraga arendsii 'Triumph' Flower Timing Flower Timing Oweekatsc 15weekatsc a 100 C A o >. t 75 g g » 75 s so LT. g » 50 g "' IL L 25 § p 25 «E ‘M- o 3 x 0 ° 81012141618202224Nl O. 81012141618202224Ni g Photoperiod (h) ...mumw PhOtOPOTiOd ('0 °- ..— mum + mm Plant Node Development nodes to last visible leaf a 60 O -o o I: 40 .- o '5 .2 20 81012141618202224Nl Photoperiod (h) + Final nodes + New nodes Number of Flowers Plant Height at First Flower a 12 i T 1 30 g 1° " i l g 25 g 8 ' 2. 2° '6 6 4 1 E 15 0 3 4 l - 9 10 .o x o E 2 4 :I: 3 , 1 . z 0 ~ ‘ ' o . 81012141618202224N| 81012141618202224N| Photoperiod (h) Photoperiod (h) Figure 34. Effects of various photoperiods with and without a cold treatment in in Saxifraga arendsii 'Triumph' in Year 2. Error bars show 95% confidence intervals. 160 Saxifraga arendsu 'Triumph' on mm mm VN NN ON 9. or 3 N_. on mm mm ¢N NN cm 9‘ 2 3 Nw om mm mm vm NN ON 3 0r 3 Nr $3225 853:8 $3 Eowoio. Ema bum ..fieafi. Esme gnaw s E .m .9 9538 Base 8285 co 98 van 6 .m .