. .92 4 (-7.4 In. . i. xvi-“3&4. N V ( I . «mum 4.. a», RV. ..~1 v1. 15 ”1:3 .3 -U 'Illlu’ul‘lllllulll- .I ‘p PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KzlProlecc8Pres/CIRCIDateoue.indd FACTORS AFFECTING THE EFFICACY OF TWO TURFGRASS GROWTH REGULATORS, TRINEXAPAC-ETHYL AND V-10029 By Matthew James Fagemess A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Crop and Soil Sciences 1997 ABSTRACT FACTORS AFFECTING THE EFFICACY OF TWO TURFGRASS GROWTH REGULATORS, TRINEXAPAC-ETHYL AND V-10029 By Matthew James Fagerness Tn'nexapac—ethyl is a foliar absorbed turfgrass growth regulator that inhibits shoot growth in many turfgrass species. Research initiated in 1995 showed the addition of Sylgard 309° + 28% urea ammonium nitrate could significantly enhance the eficacy of trinexapac-ethyl on four cool-season turfgrass species. Ammonium sulfate compensated for reductions in trinexapac-ethyl eficacy when hard water was the carrier and increased the rainfastness of trinexapac-ethyl for perennial ryegrass. The plant base was determined to be a preferred site of absorption for "C-trinexapac—ethyl in Kentucky bluegrass while absorption by the leaf blade could be enhanced by adding Sylgard 309°. Translocation of “C-trinexapac-ethyl in Kentucky bluegrass was acropetal when the material was absorbed by the plant base and basipetal when the material was absorbed by the leaf blade. An experimental turfgrass growth regulator, V-10029, significantly suppressed seedhead formation and shoot growth in five cool-season turfgrass species but was more injurious than trinexapac-ethyl, due to its herbicidal mode of action. INTRODUCTION Trinexapac-ethyl is a foliar absorbed turfgrass growth regulator that has been commercially available since 1993. Trinexapac-ethyl inhibits the biosynthesis of gibberellic acid by inhibiting the 3B-hydroxylation of GA” to 6A,. Compared to the activity of other turfgrass growth regulators that inlu'bit gibberellin biosynthesis, this inhibition appears very late in the pathway, occurring immediately prior to biosynthesis of the primary active gibberellin that stimulates shoot elongation. Compounds such as paclobutrazol and flurprimidol inhibit gibberellin biosynthesis by blocking an earlier step in the pathway. The singularity of the mode of action seen with trinesrapac-ethyl, coupled with its relative market infancy, create a multitude of research areas that need to be addressed with this compound. Research concerning the physiological efi‘ects of applying turfgrass growth regulators, especially on a long-term basis, has been limited. Commercially available products such as trinexapac-ethyl may impact the responses of turfgras species to a variety of environmental and cultural stresses. The purpose of the research presented in this thesis was to determine the influence of several factors related to growth regulator application on the efficacy of trinexapac—ethyl and of an experimental growth regulator, V-10029, which has an herbicidal mode of action. iii ACKNOWLEDGEMENTS Therearemanypeoplewhohavebeenofgreat assistaneetomethroughoutmy career at Michigan State University. The help and guidance fi'orn them all has helped make my time here both enjoyable and, more importantly, a positive educational experience as I move on towards my career goals. I would first like to thank Dr. Donald Penner, whose sound adviceandunderstmdinghumsdemecomemappredatethenatureofreseardt Special thanks also to Frank Roggenbuck who helped teach me how to do research and to Terry Wright, who perhaps was the only reason I came to Michigan State. I would like to thank as well Novartis, especially Dr. Joseph Dipaola, for the industry support behind my thesis project. I must also acknowledge the help I have received fi'om my committee members as well as the other faculty members and graduate students in both weed science and in turfgrass science. Perhaps largely unnoticed by many but not by me was the tremendous amount of time and effort put forth by our undergraduate help, namely Renee Feldpausch, Susan Redwine, and Caren Schmidt. Thank you all so much. Finally, I must thank my family, whose love and support over the past couple of years has really helped me adjust to being so far away fiom home. iv TABLE OF CONTENTS LIST OF TABLES ......................................................................................... vii LIST OF FIGURES ........................................................................................ ix CHAPTER 1 SPRAY APPLICATION PARAMETERS THAT INFLUENCE THE GROWTH INHIBITING EFFECTS OF TRINEXAPAC-ETHYL ABSTRACT ........................................................................ 1 INTRODUCTION .............................................................................. 3 RESULTS AND DISCUSSION 12 LIST OF REFERENCES ................................................................... 18 CHAPTER2 “C-TRINEXAPAC—ETHYL ABSORPTION AND TRANSLOCATION IN KENTUCKY BLUEGRASS ABSTRACT ....................................................................... 33 INTRODUCTION ............................................................................. 35 RESULTS AND DISCUSSION W42 LIST OF REFERENCES ..................................................................... 46 CHAPTER3 EVALUATION OF V-10029 AND TRINEXAPAC-ETHYL AS TURFGRASS GROWTH REGULATORS ON FIVE COOL-SEASON SPECIES ABSTRACT ........................................................................ 57 INTRODUCTION ............................................................................. 59 LIST OF REFERENCES ..................................................................... 69 CONCLUSIONS ........................................................................................... 82 LIST OF TABLES CHAPTERI Table l - Trinexapac-ethyl emcacy on ‘Georgetown’ Kentucky bluegrass: the efi'ect of application rates and Sylgard 309' + 28% urea ammonium nitrate ...................................................... Table 2 - Trinexapac-ethyl eficacy on ‘Mondial’ perennial ryegrass: the efl‘ect of application rates and Sylgard 309' + 28% urea ammonium nitrate ...................................................... Table 3 - Trinexapac-ethyl eficacy on ‘Putter’ creeping bentgrass: the efl‘ect of application rates and Sylgard 309‘ + 28% urea ammonium nitrate ...................................................... Table 4 - Trinexapac-ethyl eficacy on ‘Triathlon’ tall fescue: the efi‘ect of application rates and Sylgard 309' + 28% urea ammonium nitrate ...................................................... Table 5 - Efl'ects of hard water cations and ammonium sulfate (AMS) on trinexapac-ethyl activity in perennial ryegrass ........................ Table 6 - Efi‘ects of hard water cations and ammonium sulfate (AMS) on sethoxydim activity in perennial ryegrass ............................. Table 7 - Rainfastness characteristics of trinexapac-ethyl applied to perennial ryegrass............ ............................................................ Table 8 - Rainfastness characteristics of sethoxydim applied to perennial ryegrass. ....................................................................... Table A1- Specifications for greenhouse spray carrier volume treatments... Table A2 - Specifications for field spray carrier volume treatments ................................................................................. vii Ease ...... 21 ...... 22 ...... 23 ...... 24 ....... 25 ....... 26 ........ 27 ........ 28 ...31 ...... 32 CHAPTER2 Baas Table 1 - Content of absorption study "C treatment solutions ........................ 49 Table 2 - Plant segmentation in translocation studies ................................... 50 Table 3- Translocation of “C trtnertapac-ethyl from three sites of uptake... 53 Table 4 - Translocation of “C trinexapac-ethyl when rhizomes are present ......... 54 Table Bl - Contents of stock nutrient solutions ......................................... 55 Table B2 - Contents of full strength I-Ioagland’s solution .............................. 56 CHAPTER 3 Table l - Seedhead suppression in tall fescue and perennial ryegrass pots, as a function ofturfgrass growth regulator treatment....... 71 Table 2 - Growth inhibiting efi‘ects of V-10029 and trinexapac-ethyl on perennial ryegrass .............................................................. 72 Table 3 - Growth inhibiting efi'ects of V-10029 and trinexapac-ethyl on Kentucky bluegrass ............................................................ 73 Table 4 - Growth inhibiting efi'ects of V-10029 and trinexapac-ethyl on tall fescue ....................................................................... 74 Table 5 - Growth inhibiting efi‘ects of V-10029 and trinexapac-ethyl on creeping bentgrass ............................................................. 75 Table 6 - Growth inhibiting efi‘ects of V-10029 and trinexapac-ethyl on creeping red fescue ............................................................ 76 LIST OF FIGURES CHAPTERI 238: Figure l - Efi‘ects of spray carrier volume and ammonium sulfate on ‘Mondial’ perermial ryegrass 14 days after treatment ................... 29 Figure 2 - Efi‘ects of spray carrier volume on clipping production 21 days afier treatment for ‘Blacltsburg’ Kentucky bluegrass mowed at S, 7.5, and 10 cm ....................................... 30 CHAPTER 2 Figure 1 - Chemical structure for 1,2,6-“C-trinexapac-ethyl ......................... 48 Figure 2 - Absorption patterns of l‘C-trinexapac-ethyl over a 24 hour period fi'om three sites of uptake. The patterns are represented by the followin regression data: Roots: Y==0.219x + 0.042 .901 LeafBlade: Y2=1043.1x°°’ + 15.9 r1=.995 Plant Base: Y==0.961nx + 84 82951 ........................................ 51 Figure 3a,b - The efi‘ect of Sylgard 309° on absorption of "C-trinerrapac-ethyl by the leaf blade (Figure 3a) and by the plant base (Figure 3b) ......................................... 52 CHAPTER3 m Figure l -Decreasedpredictabi1ityinthepatternsofgmwth inhibition over time in Kentucky bluegrass, as a function of increasing rate of V-10029. LSD”, values for 2, 4, and 6 WAT were 16, 27, and 31, respectivelywhile R2 values for the depicted curves for growth inhibition at 2, 4, and 6 WAT were 0.73, 0.47, and 0.25, respectively ............. 77 Figure 2 - Growth inhibiting efi’ects of V-10029 (0.015 kyha) across four turfgrass species. The LSD“: values for weeks 1, 2, 3, and 4 were 17, 19, 28, and 24, respectively .............. 78 Figure 3 - Growth inhibiting efi‘ects of V-10029 (0.029 kg/ha) across four turfgrass species. The LSD”: values for weeks 1, 2, 3, and 4 were 13, 17, 21, and 29, respectively .............. 79 Figure 4 - Growth inhibiting efi'ects of V-10029 (0.059 kg/ha) across four turfgrass species. The LSD“, values for weeks 1, 2, 3, and 4 were 13, 15, 21, and 25, respectively .............. 80 Figure 5 - Growth inhibiting effects of trinexapac-ethyl (0.287 kg/ha for Kentucky bluegrass; 0.382 kg/ha for perennial ryegrass, tall fescue, and creeping bentgrass) across four turfgrass species. The LSD“, values for weeks 1, 2, 3, and 4 were 34, 19, 32, and 34, respectively .............................................. 81 CHAPTERI SPRAY APPLICATION PARAMETERS THAT INFLUENCE THE GROWTH INHIBITING EFFECTS OF TRINEXAPAC-ETHYL SPRAY APPLICATION PARAMETERS THAT INFLUENCE THE GROWTH INHIBITING EFFECTS OF TRINEXAPAC-ETHYL Matthew James Fagerness ABSTRACT Trinexapac-ethyl is a foliar absorbed cyclohexanedione turfgrass growth regulator that can inhibit shoot growth in a broad range of turfgrass species. Greenhouse studies were conducted with trinexapac-ethyl to investigate the efl‘ects of hard water, rainfastness, photolability, and spray carrier volume on eficacy. Results were compared to those obtained with the cyclohexanedione herbicide, sethoxydim. Studies with Kentucky bluegrass, perennial ryegrass, creeping bentgrass, and tall fescue showed that the adjuvant combination of Sylgard 309' plus 28% urea ammonium nitrate (UAN) could significantly enhance the efficacy of trinexapac-ethyl when applied at or below 0.382 kg ha". Ammonium sulfate compensated for reductions in both trinexapac-ethyl and sethoxydirn activity on perermial ryegrass when hard water was the carrier. Ammonium sulfate increased the rainfastness of trinertapac-ethyl on perennial ryegrass to a greater extent than did Sylgard 309" while the rainfastness of sethoxydim was only increased by the latter. Photolabilty of formulated trinexapacoethyl on perennial ryegrass was not observed to be a liability. Increasing spray carrier volume did not significantly influence the efiicacy of trinexapac-ethyl on perennial ryegrass, although efficacy enhancements observed with additions of ammonium sulfate were only apparent at low volumes. The impact of ammonium salts on trinexapac—ethyl activity had a broad range of implications for applications of this growth regulator. . 2 Abbreviations: UAN - urea ammonium nitrate, E formulation - ernulsifiable concentrate formulation, AMS - ammonium sulfate, UV - ultraviolet, DAT - days afier treatment INTRODUCTION Foliar absorbed herbicides provide efi‘ective post-emergence control of a broad range of agronomic weeds. Uptake by the shoots speeds the arrival of active ingredient at the target tissue, as compared to root absorbed materials. However, eficacy of foliar absorbed herbicides is dependent on a greater number of application parameters than for herbicides applied to the soil. Examples of such parameters are adjuvants, chemical rainfastness, carrier water quality, chemical photolability, and spray carrier volume. Trinexapac-ethyl is a foliar absorbed turfgrass growth regulator that can cause growth inhibition, with maximum eficacy at 14-21 days after treatment, in numerous turfgrass species (Johnson, 1993; Johnson, 1994). Trinexapac-ethyl falls into the cyclohexanedione class of herbicide chemistry, having structural similarity to both sethoxydim and clethodim, two common grarninicides. Spray application parameter efi‘ects on trinexapaeethyl eficacy may therefore follow the same trends that exist for sethoxydim and/or clethodim The label for trinexapac-ethyl identifies a broad range of effective spray carrier volumes (187-1683 1 ha"), achievement of rainfastness within one hour after application, and no need for an adjuvant to enhance emcacy. There is little evidence available to either support or discount these stipulations. Adjuvants may enhance spray droplet coverage on plant leaves and chemical absorption (Wanarnarta et al., (1989a); Bridges et al., 1991; Bridges et al., 1992). These efi‘ects may be due to a number of possible mechanisms, including cuticle solubilization and physical interaction with the herbicide in question (Wanamarta et al., (1989a)). 3 4 Adjuvants can thus be used to increase herbicide eficacy, enhance activity of photolabile herbicides, or overcome problems with antagonism from other chemicals and/or solution salts (Campbell and Penner, 1985; Penner, 1989; Hazen and Krebs, 1992; McInnes et al., 1992). Many adjuvants have been tested for enhancement of eitha' sethoxydim or clethodim with varying results (Wanamarta et al., (1989a); Bridges et al., 1992; Hazen and Krebs, 1992; Jordan et al., 1996). Therefore, the choice of adjuvant can be very critical. Herbicide applications on grasses may show a favorable response to additions of organosilicone surfactants, ofl‘ering potential for these adjuvants to enhance herbicide eflicacy (Field and Bishop, 1988; Roggenbuclt et al., 1990; Sun et al., 1996). Chemical rainfastness is the time required after a chemical application for enough absorption to occur so that activity is not diminished by subsequent rainfall removing the chemical from the leaf surface. Postemergence herbicides with slow rates of foliar absorption are often susceptible to losses of activity via this mechanism. Thus, the influence of adjuvants in enhancing foliar absorption has been a logical area of interest. Organosilicone surfactants have the potential for increasing the rainfastness and/or eficacy of some herbicides (Jansen, 1973). However, the mechanism by which such improvements act is disputed. Field and Bishop (1988) document reduced spray solution surface tension as a result of organosilicone surfactant action. Accelerated absorption would enhance rainfastness. They assumed increased absorption was related to herbicide penetration through the stomatal pore. However, this mechanism doesn’t explain variable eficacy responses between organosilicone surfactants which similarly reduce spray solution surface tension. Enhanced cuticular penetration thus appears responsible for the activity enhancement (Roggenbuck et al., 1990). Sun et al. (1996) document rainfastness as soon 5 as 15 minutes after application of primisulfuron on velvetleaf with an organosilicone surfactant. The overall potential ofthese mrfactantstoincreaserainfastnessisclearand warrants invesfigafionwithanychemicalwhererainfamsmayneedtobeincreased. Carrier water quality can afiect absorption and/or activity of herbicides due to the presence of Na", Ca”, Mg“, and Fe“. Glyphosate activity is reduced in the presence of many soluble cations because it forms a conjugate salt with the inorganic cation (Stahlman and Phillips, 1979; Nalewaja and Matysiak, 1993; Thelen et al., 1995). This antagonism is increased with high spray carrier volumes (Sandberg et al., 1978). Antagonism of clethodim is seen in the presence of sodium bicarbonate but is not seen with a difl'erent graminicide, quizalofop (McMullan, 1994). Antagonism between sethoxydim and Na- bentazon, through a similar formation of a Na-sethoxydim conjugate salt is also well documented (Jordan and York, 1989; Wanarnarta et al., (1989b); Wanarnarta et al., 1993; Nalewaja et al., 1994; Thelen et al., 1995). These antagonisms can be overcome with adjuvants. Addition of ammonium salts efl‘ectively restores herbicide activity by replacing the metal cation portion of the inactive conjugate salt with ammonium, forming a new conjugate that has higher solubility and thus is more readily absorbed (Jordan and York, 1989; Wanarnarta et al., 1989; Nalewaja and Matysiak, 1993; Wanarnarta et al., 1993; Thelen et al., 1995). Other methods such as replacing tank mixed applications of Na-bentazon and sethoxydim with sequential applications also overcome this antagonistic efl'ect (Rhodes and Coble, 1983). The specific susceptibility of some herbicides, especially the cyclohexanediones, to antagonism with inorganic metal cations provides a basis for research with other products showing similar chemistry. 6 Chemical photolability, most commonly due to the efl‘ects of ultraviolet (UV) light, can reduce the eficacy of some herbicides. Cyclohexanedione grass herbicides are particularly susceptible to photolability (Zorner et al., 1989; Bridges et al., 1992; Hazen and Krebs, 1992; McInnes et al., 1992; McMullan, 1994; Nalewaja et al., 1994; McMullan, 1996). The same researchers suggest a slower rate of foliar uptake than photodegradation as the reason why herbicide eflicacy is reduced, a mechanism that can be counteracted with an adjuvant. Na+ in the spray solution inhibits cyclohexanedione foliar absorption and amplified photolability-induced eficacy reductions (McMullan, 1994; Nalewaja et al., 1994). Trinexapac-ethyl is also a cyclohexanedione but is labeled as having rapid foliar absorption and, as such, should be less susceptible to photolability. The objectives of this research were to test whether spray application parameters that influence cyclohexanedione herbicide eficacy also influence the eficacy of trinexapac-ethyl. These parameters include adjuvants, carrier water quality, rainfastness, chemical photolability and spray carrier volume. Successfully identifying if trinexapac- ethyl applications Should follow the same guidelines as those for other cyclohexanedione herbicides may provide valuable information necessary in maximizing eflicacy of this turfgrass growth regulator. MATERIALS AND METHODS Turfgrass studies involving spray application parameters that impact eficacy of cyclohexanedione herbicides were initiated with trinexapac-ethyl and, in some cases, sethoxydim. Studies in the greenhouse were at 25 C +/- 2 C with supplemental lighting from high-pressure sodium lights providing 1200 umol photons rn‘2 s" during 18 hours of daylight. All pots were irrigated daily or as needed and received 5 kg nitrogen ha" in the form of Peters" 20-20-20 fertilizer on a weekly basis. Unless otherwise indicated, the specifications for treatment applications were as follows. Treatments were applied with a continuous link-belt sprayer at 170 kPa and 230 1 ha‘1 spray pressure and carrier volume, respectively, with a nozzle height of 30 cm above the canopy. Applied rates for trinexapac-ethyl as the IE formulation and sethoxydim as the 1.531; formulation were 0.191 kg ha" and 0.114 kg ha“, respectively. WW Studies were initiated in September, 1995 to evaluate the impact of an activator organosilicone adjuvant on the eficacy of trinexapac-ethyl, applied to four cool-season turfgrass species. Plant material studied was ‘Georgetown’ Kentucky bluegrass (Poa pratensis L.), ‘Mondial’ pereruiial ryegrass (Lo/iron perenne L.), ‘Putter’ creeping bentgrass (Agrosris palusrris Huds.), and ‘Triathlon’, a blend of three varieties of tall fescue (Fesruca arundinacea Schreb.). Plugs of 1-year old tall fescue were imported as sodfi'omtheHancockTurfgrassResearchCenterinEastIansing, MItothegreenhouse, where they were placed into 946-ml pots containing Bacctoo potting media and allowed to acclimateoveratwoweekperiodbeforebeingsprayed. Theotherturfgrasseswere established fi'om seed at recommended seeding rates three months prior to the studies. The study was done twice in September, 1995 with Kentucky bluegrass, creeping bentgrass, and tall fescue. Perennial ryegrass was sprayed in January, 1996. Treatments for each species included a control and trinatapac-ethyl at 0.048 kg ha" (tall fescue only), 0.095 kg ha", 0.191 kg ha", 0.382 kg ha", and 0.763 kg ha" (all species but tall fescue). All treatments were sprayed with or without Sylgard 309' +28% urea ammonium nitrate (UAN) adjuvant (5.0 ml r' and 10.0 ml r‘, respectively). arrier Water 01' mi Studies were initiated in the fall of 1995 to evaluate adjuvants for enhancing both carrier water quality and chemical rainfastness, with respect to sethoxydim and trinexapac- ethyl eflicacy. Applications were made to 2-month old ‘Mondial’ perennial ryegrass (Lolium perenne L.), originally seeded into Baccto' potting media in 946-ml pots at 293 kg seed ha”. Pilot studies were conducted to determine rates for sethoxydim and trinexapac-ethyl which would be less than fully efl'ective, such that efl‘ects of adjuvants would be detectable. Carrier water quality studies were sprayed in September and October, 1995 for both sethoxydim and trinexapac-ethyl treatments. Three carrier solutions were selected: deionized water, 0.5 mg 1‘1 calcium acetate, and 0.5 mg 1’l magnesium acetate. The pH of 9 each stock solution was determined. Trinexapac-ethyl and sethoxydim treatments were applied in each carrier solution, with or without ammonium sulfate (AMS) at 5.0 gl". ForrdnfannusuudiauinexapaeahylandsahoxydimUeaunanswueappfied with each of the following adjuvant combinations: no adjuvant, AMS at 5.0 g 1‘ ‘, Sylgard 309' at 5.0 ml 1", and Sylgard 309' plus AMS. A simulated 1.25 cm ofrainfall was applied at 303 kPa, 20 cm above the canopy surface in 1.25 mirnltes. Chemical treatments weregiveneachofthefollowingfourrainfallevents: norainfallevent, arainfallevent immediately after chemical treatment, and rainfall events either 15 or 30 minutes after treatment. The experiment had t1uee replications and was repeated. Chemical Photgfliligg A study was initiated in April, 1996 to evaluate the potential for high intensity UV light to reduce the eficacy of both sethoxydim and trinexapac-ethyl. The plant material used was 8-month old ‘Mondial’ pereru'lial ryegrass, originally seeded into Baccto' potting media at 293 g seed ha". Chemical treatments included sethoxydim and trinexapac-ethyl, with or without Sylgard 309° adjuvant at 5.0 ml rl . Uv light exposure for each treatment was for 0, 20, or 40 minutes afier spray application in a Rayonet’ photochemical reactor containing 12 high intensity UV bulbs around the interior perimeter. Light intensity in the chamber was 15 w m'z. 10 W Studies were initiated in September, 1996 to determine the efl‘ects of adjuvants and spmycarfiavolumeonuinmpaeahyleficacyinthegreenhouseandmdamnmeflw effects of spray carrier volume and mowing height on trinenpac-ethyl eficacy in the field. Plant material was l-yesr old ‘Mondial’ perermial ryegrass, originally seeded into Baccto' potting media at 293 g seed ha", for the greenhouse study and 3-year old ‘Blacksburg’ Kentuckybluegrass, establishedinanativesandyclayloamsoilattheHancockTurfgrass Research Center in East Lansing, MI. Greenhouse treatments included four adjuvant combinations: no adjuvant, Sylgard 309O at 5.0 ml 1", AMS at 5.0 g1", and Sylgard 309' plus AMS; each was applied with or withouttrinexapac-ethyl. Alltnatmentsincludingtrinmpac-ethyland/oradjuvantwere sprayed at each offive spray carrier volumes (187 1 ha", 561 1 ha", 935 I ha", 1309 1 ha", and 1683 1 ha") which encompassed the range indicated on the trinexapac-ethyl label. A 7.32 m by 14.64 m field plot was staked out and split into three equal-Sized 4.88 mby 7.32 m areas. Eachoftheseareaswasmowed atadifl‘erent cutting height (5, 7.5, and 10 cm). Single treatment plot area was 1.22 m by 2.44 m. All plots were irrigated daily or as needed and received 25 kg N ha'1 in the form of urea (46-0-0) on a biweekly basis. The study was initiated in September, 1996 in the early morning. Air temperature was 19 C and wind speed was negligible. Each mowing height block received three chemical treatments plus an untreated control treatment. Trinexapac-ethyl as the IE formulation was applied with a backpack sprayer at a rate of 0.287 kg ha'1 for each of the three spray carrier volumes (187 I ha", 561 1 ha", and 1683 1 ha"). Clippings were 11 collected at 7, 14, 21, and 28 days after treatment. Clippings were oven-dried for 48 hours and then weighed. The study was a 3 by 4 randomized complete block design and all treatments had three replications. Themrfgmssesinthegreuulousepotsweremaintainedata4anmtfingheight (2 cm for creeping bentgrass) before studies were sprayed and mowed back to this height when data was collected. Evaluation of trinexapac-ethyl growth inhibition was determined by production of clipping fiesh weight. Clipping weights were determined at 7, 14, and 21 days after treatment. The multiple species study had an additional clipping harvest at 28 DAT. Eflicacy of sethoxydim treatments was based on visual injury ratings (0-10 scale: 0=a1ninjured, 10=complete bumdown) taken at 12, 16, and 20 days after treatment. All greenhouse studies were completely randomized designs, had four replications per treatment, and were repeated, unless otherwise indicated. Data reflect combined means fiom both runs of repeated studies. Statistical analyses were based on factorial analysis of variance, with significance set at the 5% level. In the carrier volume study, the percent of control data were transformed to the arcsine for analysis of variance and mean separation. RESULTS AND DISCUSSION The magnitude of growth inhibition by trinexapac-ethyl differed for the difl‘erent species in this study. Maximum growth inhibition of trinexapac-ethyl occurred at either two to three weeks after treatment with all of the species, supporting specifications given on the chemical label. All trinexapac-ethyl treatments, with or without Sylgard 3090 plus 28% UAN, significantly reduced clipping production in Kentucky bluegrass 7 days alter treatment (DAT). The adjuvant significantly enhanced the performance of tnnexapac-ethyl at 0.095 kg ha'1 7 DAT (Table 1). Trinexapac-ethyl eficacy, at the lower two rates, was enhanced by the adjuvant 14, 21, and 28 DAT. Trinerrapac-ethyl at 0.382 kg ha" and at 0.753 kg ha" performed equally well for the first two weeks of the study but trinexapac-ethyl at the higher rate caused greater growth inhibition over the last two weeks. Trinexapac-ethyl at all rates, with the adjuvant, was still inlubiting growth 28 DAT. Trinexapac-ethyl, at all rates, inhibited growth of perennial ryegrass 7 DAT. However, among trinexapac-ethyl treatments, little significance was observed (Table 2). A notable exception at 7 DAT was the enhancement of growth inhibition caused by the adjuvant at the 0.095 kg ha" rate. A similar enhancement was observed 14 DAT. High variation between treatment replicates at 28 DAT resulted in a lack of significant growth inhibition in almost all treatments. 12 13 Trinexapac-ethyl, at all rates, inhibited the growth of creeping bentgrass 7 DAT but no difl‘erences between treatments were observed. The adjuvant significantly enhanced trinexapac-ethyl at both 0095 kg ha" and at 0.191 kg ha" 14 DAT (Table 3). A similar enhancement occurred 21 DAT. Trinexapac-ethyl at 0.048 kg ha" was included exclusively with tall fescue and it did not inhibit growth throughout the study. Trinexapac-ethyl did not have a significant overall impact on tall fescue as few treatments mecessflilly inhibited shoot growth (Table 4). The adjuvant enhanced trinexapac-ethyl at 0.095 kg ha" 14 DAT but other enhancements were not observed. Growth of creeping bentgrass was the most inhibited by trinexapac-ethyl whereas grth of tall fescue was the least inhibited, due to difl‘erences in maturity level between the two species. Growth of Kentucky bluegrass and perennial ryegrass was inhibited to a similar extent. Trinexapac-ethyl at 0.763 kg ha" was generally the most efl‘ective treatment applied. However, trinexapac-ethyl at this rate was at least twice the recommended rate for all the Species and was often seen to cause both visual injury and discoloration unacceptable to turfgrass managers. The selected Sylgard 309' plus 28% UAN adjuvant combination was efl‘ective in enhancing the eficacy of trinerrapac-ethyl. Growth inhibition in tall fescue was usually unaffected by the addition of the adjuvants. However, growth of all other species was generally more inhibited by trinexapac-ethyl with the adjuvant than without it. The adjuvant oflen reduced the necessary application rate of trinexapac-ethyl for a given response level. The exact mechanism for such an enhancement was not explored in this study but probably was a function of both of the constituents in the adjuvant. l4 Perennialryegrasswassensifivetotheefl‘ectsofbothuinetapac-ethyland sethoxydim. Seven DAT, the potentially antagonistic impact of calcium and magnesium in the water carrier on trinexapac-ethyl eficacy was not yet evident and AMS did not impact eflicacy (Table 5). However, at 14 DAT, trinexapac-ethyl applied in calcium and magnesium carrier solutions without AMS had clipping production equal to that for the untreated control. Trinexapac-ethyl applied in calcium and magnesium carrier solutions with AMS, conversely, significantly decreased clipping production. This efl‘ect diminished by 21 DAT. Sethoxydirn injury was significantly enhanced by the addition of AMS to both the calcium and magnesium carrier solutions at both 12 and 16 DAT (Table 6). Twenty DAT, the level of injury in all treatments was greater than 60% and no significant differences among treatments were observed. The pH values for the calcium and magnesium canier solutions were 7.1 and 7.5, respectively. The influence of pH on the results appeared negligible. Although significant differences were observed between treatments for both chemicals, the results may have reflected the low 5.0 gl" level of AMS that was used. AMS levels of 10 g1" and 20 g 1'1 are commonly used in many herbicide applications. The potential for AMS to ofl‘set the antagonistic effects of hard water cations is clear and AMS may have had even a greater positive impact in these studies had the level included been greater. Clipping production was significantly reduced by all trinexapac-ethyl treatments in rainfastness studies. Rainfastness of trinexapac-ethyl was also significantly increased by 15 the three adjuvants selected for these studies. Seven DAT, AMS seemed to play an important role in enhancing eficacy as only treatments containing AMS produced sigrfificanflyfewercflppingsflunmnnansmadjuvunsformyofthewuhofl‘ times (Table 7). Fourteen DAT, all trinexapac-ethyl treatments containing Sylgard 309° and most containing AMS showed enhanced eficacy, as compared to treatments without an adjuvant, at allthreewashofl'times (Table 7). Lossoftrinexapac-ethylactivitydueto washofl‘ was observed with washes at both 0 and 15 minutes after application. Compared to their unwashed counterparts, only trinuapac-ethyl with Sylgard 309° plus AMS at the 0 minute washofl‘ and trinexapac-ethyl with AMS at the 15 minute washofl‘ had no loss of activity due to the washofl‘s. By 30 minutes alter treatment, washofl‘had no significant impact on trinexapac-ethyl activity. Twenty-one DAT, the positive impact of AMS on trinexapac-ethyl rainfastness had disappeared. Loss of trinexapac—ethyl activity due to washofi‘ occurred with AMS and with Sylgard 309‘ plus AMS at the 0 minute washofl‘time and with AMS at the 30 minute washofl‘ time (Table 7). Trinexapac-ethyl with Sylgard 309' sufl‘ered no loss of activity at any of the washofl‘ times. All sethoxydim treatments showed significantly greater injury than untreated controls over the duration of the study. None of the adjuvants enhanced sethoxydim activity in the unwashed treatments. However, significant losses of activity were observed with all washoff treatments. Sylgard 309' was the only adjuvant that restored activity lost due to washofl‘ (Table 8). 16 Overall, trinexapac-ethyl seemed to be more rainfast than was sethoxydim. The label for trinexapac-ethyl indicates rainfastness within one hour of application. Results fromthesestudies suggestedrainfastnesswasmorerapidthanthat, especiallywhenan adjuvant was included. AMS, with or without Sylgard 309°, significantly increased the rainfafinessoftfinexapaedhylwhflehhadmhpaflonflwrdMusofsflhoxydim. Sylgard 309° increased the rainfastness of both cyclohexanediones. chemical Ehotolabilim Neither sethoxydim nor trinexapac-ethyl treatments sufl‘ered any loss of activity fi'om the effects of UV light exposure. Sethoxydirn was expected to lose activity under such conditions while the impact on trinexapac-ethyl was unknown (Zorner et al., 1989; Hazen and Krebs, 1992; McInnes et al., 1992; Nalewaja et al., 1994; McMullan, 1996). Newer formulations of sethoxydim contain additives inhibitory to UV photodegradation. It does not appear that trinexapac-ethyl, applied as the IE formulation, is susceptible to loss of activity due to photolability. We AMS enhanced trinexapac-ethyl eficacy in the greenhouse at 187 1 ha‘1 but insignificantly or negatively impacted eficacy at higher volumes 14 DAT (Figure 1). The observed enhancement supported results seen in the carrier water quality study. However, the lack of enhancement seen at 561 l ha'l or greater was probably related to the low AMS concentration, 5.0 gl". At high spray volumes, the ratio of ammonium to Ca++ and Mg” in the hard water may have been insuficient to ofl‘set the negative impact of hard water l7 cations on uptake. AMS has the potential to enhance trinexapac-ethyl eficacy across a broad range of spray carrier volumes but it is recommended that a higher level of AMS be included at higher volumes to adequately account for hard water cations. A significant interaction between mowing height and spray carrier volume was observed at 21 DAT. Results were not conclusive at 7, 14, or 28 DAT. Carrier volume significantly impacted trinatapac-ethyl eficacy at the 5 cm and 10 cm mowing heights (PiguresZand3). Adensercanopystinmlatedbylateralgrowthandahighercanopywith more leaf tissue may necessitate higher spray carrier volumes for applications to the 5 cm and 10 cm mowing heights, respectively. Treatments mowed at 7.5 cm, with predominance of neither lateral development nor excess leaf matter, responded equally well to trinexapac-ethyl at all spray carrier volumes (Figure 4). Because a significant interaction occurred when trinexapac-ethyl exhibited maximum eficacy at 21 DAT, an interaction between cutting height and spray carrier volume may have significant implications for field applied trinexapac-ethyl. It is suggested, based on results fi'om greenhouse studies, that adjuvants have a beneficial impact on trinexapac-ethyl eficacy, rainfastness, and activity in hard water, especially at lower spray carrier volumes. The role of AMS in overcoming hard water problems with trinexapac-ethyl is likely a key factor in explaining increased rainfastness and enhanced emcacy observed with trinexapac-ethyl treatments containing AMS. Uptake of trinexapac-ethyl seemed to be inherently more rapid than absorption of sethoxydim. LIST OF REFERENCES Bridges, D.C., AE. Smith, and LN. Falb. 1991. Efl'ect of adjuvant on foliar absorption and activity of clethodim and polar degradation products of clethodim. Weed Sci. 39:543-547. Bridges, D.C., L.N. Falb, and AE. Smith 1992. Stability and activity of clethodim as influenced by pH, UV light, and adjuvant. p. 215-222. In C.L. Foy (ed.) Adjuvants for Agrichernicals. CRC Press, Boca Raton, FL. Campbell, J .R and D. Penner. 1985. Abiotic transformations of sethoxydim. Weed Sci. 33:435-439. Field, RJ. and N. G. Bishop. 1988. Promotion of stomatal infiltration of glyphosate by an organosilicone surfactant reduces the critical rainfall period. Pestic. Sci. 24:55-62. Hazen, J .L. and P.J. Krebs. 1992. Photodegradation and absorption of sethoxydim as adjuvant-influenced surface efl'ects. p. 195-203. In C.L. Foy (ed.) Adjuvants for Agrichemicals. CRC Press, Boca Raton, FL. Jansen, L.L. Enhancement of herbicides by silicone surfactants. Weed Sci. 21:130-135. Johnson, B.J. 1993. Frequency of plant growth regulator and mowing treatments: efl‘ects on injury and suppression of centipedegrass. Agron. J. 85:276-280. Johnson, B.J. 1994. Influence ofplarrt growth regulators and mowing on two bermudagrasses. Agron. J. 86:805-810. Jordan, D.L. and AC. York 1989. Efl‘ects of ammonium fertilizers and BCH 81508 S on antagonism with sethoxydim phrs bentazon mixtures. Weed Technol. 32450-454. Jordan, D.L., PR Vidrine, J.L. Grifin, and DB. Reynolds. 1996. Influence of adjuvants on eficacy of clethodim. Weed Technol. 10:738-743. McInnes, D., K.N. Harker, RE. Blackshaw, and W.H VandenBorn. 1992. The influence of ultraviolet light on the phytotoxicity of sethoxydim tank mixtures with various adjuvants. p. 205-213. In C.L. Foy (ed.) Adjuvants for Agrichernicals. CRC Press, Boca Raton, FL. McMullan, PM. 1994. Efl‘ect of sodium bicarbonate on clethodim or quizalofop eficacy and the role of ultraviolet light. Weed Technol. 8:572-575. 18 19 McMullan, PM. 1996. Grass herbicide eficacy as influenced by adjuvant, spray solution pH, and ultraviolet light. Weed Technol. 10:72-77. Nalewaja, JD. and R Matysiak. 1993. Optimizing adjuvants to overcome glyphosate antagonistic salts. Weed Technol. 7:337-342. Nalewaja, J .D., R Matysiala and E. Szelezniak. 1994. Sethoxydirn response to spray carrier chemical properties and environment. Weed Technol. 8:591-597. Penner, D. 1989. The impact of adjuvants on herbicide antagonism. Weed Technol. 3 :227-23 1. Rhodes, G.N., Jr. and HD. Coble. 1984. Influence of application variables on antagonism between sethoxydim and bentazon. Weed Sci. 32:436-441. Roggenbuck, F .C., L. Rowe, D. Penner, L. Petrofl‘, and R Burow. 1990. Increasing postemergence herbicide eficacy and rainfastness with silicone adjuvants. Weed Technol. 42576-580. Sandberg, C.L., W.F. Meggitt, and D. Penner. 1978. Effect of diluent volume and calcium on glyphosate phytotoxicity. Weed Sci. 26:476-479. Stahlman, P.W. and W.M. Phillips. 1979. Efl‘ects of water quality and spray volume on glyphosate phytotoxicity. Weed Sci. 27:38-41. Sun, 1., CL. Foy, and BL. Witt. 1996. Effect of organosilicone surfactants on the rainfastness of primisulfirron in velvetleaf (Abnn'lon rheophrasn). Weed Technol. 10:263-267. Thelen, K.D., E.P. Jackson, and D. Penner. 1995. Characterizing the sethoxydim- bentazon interaction with proton nuclear magnetic resonance spectrometry. Weed Sci. 43:337-341. Wanarnarta, G., D. Penner, and 1.]. Kells. 1989 (a). Identification of efl'rcacious adjuvants for sethoxydim and bentazon. Weed Technol. 3:60-66. Wanarnarta, G., D. Penner, and J .J . Kells. 1989 (b). The basis of bentazon antagonism on sethoxydim absorption and activity. Weed Sci. 37:400-404. Wanamarta, G., U. Kells, and D. Penner. 1993. Overcoming antagonistic effects of Na- bentazon on sethoxydim absorption. Weed Technol. 7:322-325. 20 Zorner, P., J. Hazen, R Evans, D. Gourd, and T. Fitzgerald. 1989. The influence of DASH adjuvant in limiting photodegradation of sethoxydim on leaf surfaces. p. 83. In Weed Sci. Soc. Amer. Abstracts. WSSA, Charnpaign, IL. 21 Table 1: Trinexapac-ethyl efficacy on ‘Georgetown’ Kentucky bluegrass: the effect of application rates and Sylgard 309' plus 28% urea ammonium nitrate. Time after application (days) 7 14 21 28 Trinexapac- ethyl rate -adj. +adj. -adj. +adj. -adj. +adj. -adj. +adj. kg ha“ Clipping production (96 of control} 0 100 89 100 107 100 111 100 106 (1095 76 53 88 60 90 78 101 78 0.191 60 45 56 31 70 47 75 66 0.3 82 49 36 36 23 51 19 70 45 0.763 31 28 21 15 16 10 37 24 LSD”, ' ---18-—- ---18-—— ---l7-—-- ---18-—- Table 2: Trinexapac-ethyl efficacy on ‘Mondial’ perennial ryegrass: the effect of application rates and Sylgard 309’ plus 28% urea ammonium nitrate. Time after application (days) 7 14 21 Trinexapac- ethyl rate -adj. +adj. -adj. +adj. kg ha" Clipping production (96 of control} 0 100 89 100 88 100 82 0.095 62 38 51 27 85 57 0. 191 46 41 30 20 54 37 0.382 59 38 39 23 55 27 0.763 39 29 19 10 27 17 LSDo,05 ---24-—- —--23—— —--29—-- Table 3: Trinexapac-ethyl efficacy on ‘Putter’ creeping bentgrass: the effect of application rates and Sylgard 309' plus 28% urea ammonium nitrate. Time alter application (days) 7 14 21 Trinexapac- ethyl rate -adj. +adj. -adj. +adj. -adj. +adj. kg ha-1 Clippingtlodnctiar(%ofcontrol} 0 100 104 100 95 100 119 0.095 55 38 35 17 60 41 0.191 51 38 31 13 60 22 0-332 34 38 14 12 24 14 0-763 33 32 15 6 12 10 LSD”, --—-30—— -—-15-— -——-35—-— 24 Table 4: Trinexapac-ethyl efficacy on ‘Triathlon’ tall fescue: the effect of application rates and Sylgard 309' plus 28% urea ammonium nitrate. Time after application (days) 7 14 21 28 Trinexapac- ethyl rate -adj. +adj. -adj. +adj. -adj. +adj. -adj. +adj. kg ha" Clipping prudtrcdon (96 ofcontrol‘r 0 100 102 100 101 100 118 100 106 0.043 91 76 93 79 109 93 105 103 0.095 85 63 82 50 104 74 95 82 0.191 75 60 65 44 90 68 78 68 0.382 62 57 39 38 52 52 66 57 LSDoos ---25—-—- --—-22-—— ---34-—- ---25-—- 25 Table 5: Effects of hard water cations and ammonium sulfate (AMS) on trinexapac-ethyl activity in perennial ryegrass. Time after application (days) 7 14 21 Treatment - AMS + AMS - AMS + AMS - AMS + AMS Clipping plantation (96 ofcontrol) 90m”! 100 121 100 141 100 121 Twahvlin 63 66 38 50 12 37 deromzedwater 'l‘n'11133751111119«ethylin 5-0 71 82 55 43 38 38 gl ealcrum acetate Terence-tarmac 80 69 78 36 53 33 gl magnesrumacetate 13130.05 —--39—-— —47--- -—-29--- 26 Table 6: Effects of hard water cations and ammonium sulfate (AMS) on sethoxydim activity in perennial ryegrass. Time after application (days) 12 16 20 Treatment -AMS +AMS -AMS +AMS -AMS +AMS 4% injury ’0 Control 0 0 0 0 0 0 sethoxydim 41 48 68 79 74 83 deromzed water schoxydiminwtr' 34 49 61 77 68 83 calcium acetate sahoxydiminwsr' 34 42 59 73 64 77 magnesrum acetate LSD”, ---8-—- --ll-—— -——17--- t %injmyvalueswu'econvutedfiumnaigina10-10mungsmlewhae0-1minjmed, 10-dead 27 an m— w.— 8.53 .....v. 8 8 ma 8 a. R an .m on on .m S on 3.9.85. .38. mm an we 2. on an mu 3 on 8 2. an m. 888...... .....u. 3. an em. .a an mm mm 8 .n me mm on o 885%... .....u. an 9 me 2. v. e. a. 9. R S a. S 282 892...... 8. 8. 8. 8. 8. 8. 8. 8. 8. 8. 8. 8. 0.82 .880 985835833... 39...... 8......a 238 Sn 338 SM $33 8. .28.... ...-...... clue. as? 982 can}. can}. m2... 982 23.»... cans... e..)? 982 :28? 88:8... 23. .555. .553... .N v. e A933 528.3% .3... 2...... .8!»th ...—.53.— 8 ts..—.... iforoaanuoet. be 823.8322? 3052:55— # 033—. 28 h e— A: Sen—my— ; 2. X. on on mm 2 cv we 3 mm 3 cm 5......3 on 8 we on on S 3. we 3. a... an a. m . 5.5.2.3 2. an 8 mm Q. S .v .v on 3. ea .u o 8.8.9.3 on mm on 3 2. z. X. we 3 mm on co 282 5.5.2.3 o o c c o c c c c c c o 0:62 .2230 3.5. xv 8......5 m2<+§ Sm 32$..." 8m 238 8. .55.... 983 .593 m2... 982 .53 253 ms... 982 ...-.3 25...... ms... 982 been? .558; 53.6.4. 5.5.5.. 53.5.4. on o. N. A933 5.56.39. .3... 2:3. 6229.0. 35.9.2— 3 3:1...— 5.1.3223... 3.8.8.6223 82.3555— 8 933—. 29 1 r 6 Clipplngprodlsctlonfiolcontrol) 8 Figure 1: Effects of spray carrier volume and ammonium sulfate on ‘Mondial’ perennial ryegrass 14 days after treatment. _n 88888388 Clipplngproductlonfiofcontrot) .- O .L O 1 Figure 2: Effects of spray carrier volume on clipping production 21 days after treatment for ‘Blacksburg’ Kentucky bluegrass mowed at 5, 7.5, and 10 cm. APPENDIX A 31 Table A1: Specifications for greenhouse spray carrier volume treatments Spray Carrier Volurneaha") t SprayPressurthPa) Tee-Jet5 Nozzle Tm Used 187 p 68 8001B (100 mesh screen) 561 265 80015 (100 mesh screen) 935 136 8004VS (50 mesh screen) 1309 224 8004VS (50 mesh screen) 1683 272 888005 (50 mesh screen) tAflWapphcafimwuemdewifianmlehdfluofmcmmmemopymfice 32 Table A2: Specifications for field spray carrier volume treatments. Spray Carrier Volume SprayPressure Walking Speed Tee-Jet“ Nozzle Type aha") t (kPa) (erns‘) Used 187 177 152.40 8002 (50 mesh screen) 561 258 60.96 8002 (50 mesh screen) 1683 252 30.48 8006 (50 mesh M) tAnmmappfimfimswuemndewithsnonleheigmoflOunabovemecnopymfnce CHAPTER2 "C-TRINEXAPAC-ETHYL ABSORPTION AND TRANSLOCATION IN KENTUCKY BLUEGRASS “C-TRINEXAPAC-ETHYL ABSORPTION AND TRANSLOCATION IN KENTUCKY BLUEGRASS Matthew James Fagerness ABSTRACT Trinexapac-ethyl is a foliar-applied growth regulator for turfgrass that can reduce mowing fi'equency, clipping production, and enhance turfgrass color. l‘C-Trinexapac-ethyl was used to evaluate absorption and subsequent l‘C-trinesrapac-ethyl translocation in hydroponically grown ‘Blacksburg’ Kentucky bluegrass (Pan pratensr's L.). The magnitude and rate of 1‘C trinexapac-ethyl absorption by various organs was as follows: plant base > leaf blade > roots. Over the time period of 0 to 24 hours, mardmum absorption by the plant base was obtained in 8 hours and by the leaf blade in 24 hours. Absorption by the roots was negligible. Addition of an activator organosilicone adjuvant, Sylgard 309°, significantly enhanced l‘C-tlinertapac-ethyl absorption by the leaf blade one hour afler application but did not enhance absorption by the plant base. Of the “C- trinexapac-ethyl absorbed by the plant base, over 50% was translocated to the plant foliage after 24 hours. Of the l‘C-trinexapac-ethyl absorbed by the leaf blade, one-third was translocated after 24 hours; the direction of movement was predominantly basipetal. Less than 5% of absorbed l‘C-trinerltapac-ethyl fi'orn either site was translocated to roots or to rhizomes with daughter plants, explaining the lack of inhibition of lateral turfgrass growth. Combined efl‘ects of enhanced leaf blade absorption, basipetal translocation fiorn the leaf blade, and acropetal translocation from the plant base of I‘C-tlinexapac-ethyl helped explain the positive impact of Sylgard 309‘ on efficacy and rainfastness of trinexapac-ethyl. 33 34 Abbreviations: HAT - hours alter treatment INTRODUCTION The use of “C-labeled xenobiotics is a common tool for evaluating absorption and translocation patterns of these compounds. Trinexapac-ethyl is a foliar absorbed cyclohexanedione plant growth regulator. Specificity for uptake of trinexapac-ethyl among foliar sites is unknown for trinerapac-ethyl although leaf blades come into more direct contact with the spray droplets than do other tissues. Barrett and Bartuska (1982) found the Stern, as compared to the leaves, to be a preferred site of uptake for an experimental growth retardant, PP333. Speed of absorption of foliar applied chemicals can be important in both emcacy and chemical rainfastness. Results fiom Al-Khatib et al. (1992) suggest that most herbicide absorption occurs within 24 hours of application. Trinexapac-ethyl, labeled as being rainfast within one hour of application, should conform to this pattern. The absorption pattern of a given herbicide can be afl‘ected positively by adjuvants or negatively by antagonistic chemicals and/or ions. The absorption of sethoxydim by leaves was dramatically reduced in the presence of bentazon (Rhodes and Coble, 1984). Ammonium salts were able to counteract this antagonism (Wanarnarta et al., 1993) and were able to increase leaf absorption of picloram (Moxness and Lym, 1989). Crop oil concentrate adjuvants increase the absorption of fluazifop, quizalofop, and sethoxydim in cats (Manthey et al., 1992). Organosilicone surfactants enhance the eflicacy of many herbicides and can be excellent surface wetting agents (Jansen, 1973; Roggenbuck et al., 1990). The optimal adjuvants for enhancing trinexapac-ethyl eflicacy have not been determined. 35 36 Systemic herbicide translocation from the site of uptake to the meristerns can be as important a factor in determining eficacy as absorption for herbicides afl’ecting meristematic activity. Phloem transport can be slow (Derr et al., 1985; Peregoy et al., 1990; Camacho et al., 1991). Instances where plants transclocate herbicides basipetally through phloem more rapidly may relate to observed patterns of emcacy (Hart and Penner, 1993). Response to fluazifop was dependent on translocation (Derr et al., 1985). When slow basipetal translocation does not coincide with observed patterns of herbicide efficacy, alternate sites of uptake and translocation patterns may contribute to activity (Achhireddy et al., 1985). Dicamba, for example, is translocated acropetally in several Species, the mechanism for which may or may not be phloem transport (Al-Khatib et al., 1992). Translocation of the cyclohexanedione herbicide, sethoxydim, is both acropetal and basipetal (\Vrlls, 1984). Potential impact of trinexapac-ethyl on lateral turf growth was addressed by Calhoun (1996), who reported divot closure rates in creeping bentgrass treated with trinexapac-ethyl to be equal to or higher than those for untreated controls. The translocation pattern for trinexapac-ethyl is not known but may be largely dependent on the preferred site of absorption. The objectives of this research were to (i) determine the preferred site of absorption for “C-trinexapac—ethyl, (ii) determine the effect of an organosilicone surfactant on absorption of l‘C-trinexapac-ethyl, and (iii) determine the translocation patterns of "C-trinexapac-ethyl over a 24 hour period fiom three sites of uptake. The goal was to increase the understanding of foliar absorption and translocation patterns for trinexapac-ethyl such that eficacy may be maximized. MATERIALS AND METHODS Plant material for all studies was 2-year old ‘Blacksburg’ Kentucky bluegrass (Poa prarensis L.), originally obtained as sod fi'orn the Hancock Turfgrass Research Center in EastLansing, MI. Sodpieceswereplacedinto25 cmby50cmflatscontainingacoarse soil (23% gravel, 77% sand)inthegreenhouseandallowedtlueeweekstoestablish. The soiltypewaschosensothattheplantscouldlaterberemovedfi'omthesoilwithaminimal amount of root damage. The greenhouse was at 25 C +/- 2 C with supplemental lighting from high-pressure sodium lights providing 1200 umol photons rn'2 s'1 of light during 18 hours of daylight. All flats were irrigated regularly and received 10 kg nitrogen ha" per week. The rate of nitrogen application was higher than normal for maintenance of greenhouse plants because the soil used was coarse and leached readily. Plants were established hydroponically prior to the l‘C-trinexapac-ethyl absorption and translocation studies. The hydroponic solution used was a modified Hoagland’s solution derived for hydroponic establislunent of creeping bentgrass. Aliquots of this solution were made using six stock nutrient solutions, each containing one or more essential nutrients (Menu and McBee, 1970). Deionized water was the carrier for stock solutions and the Hoagland’s nutrient solution. One-half and one-quarter Strength versions of the same Hoagland’s solution were used in some instances. Treatment solutions were made using 1,2,6-“c-1abeled trinexapac-ethyl with 94.8% purity and a specific activity of 1,139,600 Bq mg’l (Figure 1). 37 38 mm Individual plants were established in full strength Hoagland’s solution by removingthemfromthe soil, rinsingthemfreeofdebris, andwrappingthebaseofthe plant with a slit cut foam plug. Plugs were then inserted through small holes in a colored plastic lid so the roots could suspend fieely in the solution below. The solution was aerated using an air stone attached to an electric air pump. Plants studied were selected for both health and size-based uniformity. Experiments were conducted with individual plants wrapped in a foam plug and suspended into 120 ml jars containing 100 ml of one-half strength Hoagland’s solution. A one-quarter strength solution was used for root absorption experiments to minimize solute interference with l"C-trir'rexapac-ethyl uptake. All jars were wrapped with aluminum foil. The specifications for 200 ul treatment stock solutions differed according to the type of experiment (Table 1). Absorption experiments for the roots, the leaf blade, and the plant base were conducted over a three week period in May and June of 1996, while those involving the adjuvant were conducted in November, 1996. Leaf blades and plant bases were treated with a 2 pl droplet of treatment stock solution containing a quantity of trinexapac-ethyl based on a 0.191 kg ha‘1 rate (Table 1); delivery was with a 10 111 Harnilton° syringe. 4 ul droplets were placed in the rootzone solution for the root absorption experiment. Treated plant parts were washed for 45 seconds with a 1:1 methanolzwater solvent at 0, l, 4, 8, or 24 hours after treatment (HAT). (The 8 HAT wash was excluded in the adjuvant study). The choice of solvent and duration of wash were based on solubility of trinexapac-ethyl, previous work with cyclohexanedione herbicides, and a time threshold, beyond which the solvent could extract absorbed material 39 (Devine et al., 1984; Wills, 1984). Each combination of treated plant site and washofl‘ time had four replications. Rinsateswere collectedin20mlliquid scintillationvials, eachcontaining3 mlof 1:1 methanolzwater. Two 2-ml aliquots fiom each rootzone solution which had directly received l‘C-trinexapac-ethyl for the root absorption experiment were similarly collected. The solutions were then diluted to a total volume of 15 ml with Safety Solve. liquid scintillation cocktail and analyzed in a liquid scintillation counter. Difl‘erences in radioactivity between the applied droplets and the collected rinsates were presumed to be the amounts of 1‘C taken up by each plant. “C in 2 ml aliquots taken fi'om each rootzone solution was determined to evaluate the amount of “C in root absorption experiments that was not absorbed. T ranslocag’on Studies Studies were initiated in June, 1996 to measure translocation of "C-trinexapac- ethyl from three sites of uptake. A similar study evaluating two sites of uptake with plants that had rhizomes with daughter plants was initiated in September, 1996. Plants for the June study were established and selected as for absorption studies. Plants for the September study were planted into a coarse sand and allowed to generate rhizomes with daughter plants. The parent and daughter plants were subsequently transferred to a one- half strength Hoagland’s solution. The colored lid for pots was prepared with a 0.5 cm by 7. 5 cm slit so both the parent and daughter shoots could be equally supported above the solution. 40 Treatments were applied fiom a stock solution containing the following: 50 ul “C-tlinexapac-ethyl, 118.85 111 formulation blank (taken flour a 1:100 stock solution), and 31.15 ul deionized water. Sites ofuptakewerethe leafblade, the plant base, and the roots for the June study, which had six replications per treatment. The September study had four replications per treatment and did not evaluate roots as a site of uptake. Two 2-ul droplets containing 3700 Bq of 1‘C were applied to the leaf blade or the plant base and in one 4-ul droplet to the rootzone solution A ten-fold increase in radioactivity, as compared to absorption studies, was used to ensure traceability of the l‘C as it was translocated. Twenty four hours after treatment, treated plant parts were rinsed with 1:1 methanol:water", rinsates and aliquots fiom directly treated rootzone solutions were analyzed for 1‘C. Plants were segmented based on the original site of 1‘C uptake (Table 2). Segments were kept at -10 C until they were oxidized in a biological oxidizer. Eficiency of the oxidizer was greater than 95% and was determined by oxidizing a known quantity of 1‘C placed on a piece of paper towel and then calculating the percent recovery. 1‘CO; from oxidized samples was trapped in a 2:1 Safety Solve°:CarbosorbII° cocktail. Each sample was then radioassayed in a liquid scintillation spectrometer. Combined radioactivity fi'om rinsates and all plant segments was used to determine the percent recovery of 1‘C for each site of uptake. Percent recoveries of “C were 96%, 75%, and 82% for “C-trinexapac-ethyl applied to the roots, base of the plant, and leaf blade, respectively, in June, 1996. Percent recoveries of "C were 76% and 79% for l‘C- trinexapac-ethyl applied to the base of the plant and leaf blade, respectively, in September, 1996. The foam plugs that supported the plant bases accounted for a portion of the 41 unrecovered “C. Distributions of translocated l‘C-trinexapac-ethyl were expressed as percentages of absorbed “C. All studies were completely randomized designs and were repeated. Data reflect combined means fiom two experiments. Statistical analyses were based on analysis of variance and/or simple linear regression with significance set at the 5% level. RESULTS AND DISCUSSION W There were marked difl‘erencesbetweentheleafblade, theplantbase, andthe roots, in terms of both rate and total absorption of 1‘C- trinexapwethyl (Figure 2). The leaf blade absorbed only 4% of applied l‘c-trinernrpnc-ethyl before the immediate washorr but had absorbed 31% by one HAT. Absorption increased consistently over the next three washofl‘ periods, culminating in maximum absorption of 70% by 24 HAT. The base of the plant absorbed 29% of applied l‘C- trinexapac-ethyl before the immediate washofl' and had absorbed 80% by one HAT. At this point, the absorption rate slowed down, reaching a maximum by 8 HAT and a level of 94% 24 HAT. The roots absorbed little l‘C-trinexapac- ethyl, taking up only 5% of applied material after 24 hours. Commercially, trinexapac- ethyl is foliar applied so the results fiom the root absorption experiment were not surprising. The plant base was the preferred Site of absorption for l‘C-trinertapac-ethyl. Thus, the amount of trinexapac-ethyl reaching the plant base may be a major factor in determining efiicacy. The organosilicone surfactant, Sylgard 309°, is an activator adjuvant that was determined to be capable of enhancing trinexapac-ethyl eficacy (data not presented). Absorption studies with “C-trinexapac-ethyl plus Sylgard 309' were conducted with the intent of determining whether enhanced absorption, surface movement to a preferred site of absorption such as the plant base, or both were involved in the observed enhancement of trinexapac-ethyl eflicacy. 42 43 Absorption of "C-trinatapac-ethyl by the plant base was unafl‘ected by the addition of Sylgard 309' at any of the washoff times (Figure 3b). Absorption of 1‘C- trinexapac-ethyl bythe leafbladewasunafl‘ectedbytheaddition ofSylgard 309°at 0, 4, and 24 HAT. However, significant enhancement of absorption with Sylgard 309' occurred 1 HAT, the increase being fiom 21% to 51% (Figure 3a). Absorption of l‘C-trinexapac- ethyl was so rapid that a measurable amount of absorption occurred in less than the 45 seconds it took to wash the “C-trinexapac-ethyl of of either the leaf blade or the plant base for the zero time treatment. Enhanced early absorption of 1‘C trinexapac-ethyl by leaf blades when Sylgard 309’ was included could provide greater trinexapac-ethyl rainfastness. Absorption by both the leafblade and plant base was greater than 50% aaer one hour when Sylgard 309' was added to the treatment solution. The reduction of surface tension of spray droplets with Sylgard 309° allow more of the spray solution to move on the surface of the plant to the base of the plant, which is the preferred site of absorption. Translocation Studies The direction and extent of ”C-trinexapac-ethyl translocation was dependent on the site of uptake. The base of the plant absorbed 96% of applied "C-trinexapac-ethyl while the leaf blade absorbed 70% of applied "C-trinexapac—ethyl in 24 hours. Roots absorbed 2% of applied l‘C-trinexapac-ethyl in 24 hours and translocated acropetally 50% of what was absorbed (Table 3). The leaf blade retained greater than 60% of the 1‘c it absorbed after 24 hours (Table 3). One percent of absorbed “C moved acropetally while 32% moved basipetally 44 and accumulated in a variety of tissues. The plant base accumulated 11% of absorbed "C. Roots accumulated 5% ofthe absorbed “c. Greater than 75% of"C absorbed by the base ofthe plant was translocated acropetally to the plant foliage over 24 hours. Roots accumulated less than 5% of absorbed 1‘C while the other 18% remained in the base of the plant (Table 3). The leaf blade and the plant base were the two sites of absorption evaluated in translocation studies involving rhizomes/daughter plants. Greater than 70% of the “c- trinexapac—ethyl applied to the leaf blade was absorbed while the plant base absorbed greater than 85% of applied “C-trinexapac-ethyl. Translocation of absorbed 1‘C from the leaf blade was as follows: 36% translocated to other foliar tissues, 4% translocated to the roots, and 3% translocated to the rhizome/daughter plant after 24 hours (Table 4). The other 57% remained in the leaf blade. Translocation of 1‘C absorbed by the base of the plant was as follows: 61% translocated acropetally, 3% translocated to the roots, and 3% translocated to the rhizome/daughter plant after 24 hours (Table 4). The other 33% remained in the base of the plant. Acropetal translocation of 1‘C from the plant base and basipetal translocation of 1“c from the leafblade observed in both studies appeared related to the activity of trinexapac-ethyl observed in turfgrasses. Preferred directional translocation fi'om both sites of uptake would result in a convergence of the active ingredient at the intercalary meristems, which are the primary sites of growth inhibition for trinexapac-ethyl. The leaf blade retained more 1‘C after 24 hours than did the plant base. The rinsate from the leaf 45 blade washofl‘ after 24 hours also contained more “C than that from the plant base. These results were again indicative of the plant base being the preferred site of absorption for trinexapac-ethyl and the potential for an adjuvant to increase leaf blade absorption. The lack of significant “c translocation to the roots suggested that any efl'ects trinexapac-ethyl mayhave onthe rootsareafilnctionofitsefi‘ectson shootgrowth. Thesimilarlackof significant translocation to rhizome tissues suggested that trinexapac-ethyl has little impact on lateral development of rhizomatous turfgrass species such as Kentucky bluegrass. LIST OF REFERENCES Achireddy, NR, RC. Kirkwood, and W.W. Fletcher. 1985. Foliar absorption and translocation of isoproturon, and its action on photosynthesis in wheat (Tritium aestr’vum) and slender foxtail (Alopecms myowroides). Weed Sci. 33:762-765. Al-Khatib, K., R Parker, and E.P. Fuerst. 1992. Foliar absorption and translocation of herbicides from aqueous solution and treated soil. Weed Sci. 40:281-287. Al-Khatib, K., R Parker, and E.P. Fuerst. 1992. Foliar absorption and translocation of dicamba from aqueous solution and dicamba-treated soil deposits. Weed Technol. 6:57-61. Barrett, J.E. and CA. Bartuska 1982. PP333 efl‘ects on stem elongation dependent on site of application. Hort. Sci. 17(5):737-73 8. Calhoun, RN. 1996. Efi‘ect of three plant growth regulators and two nitrogen regimes on grth and performance of creeping bentgrass. MS. thesis. Michigan State Univ., East Lansing. 57 p. Camacho, RF. and LJ. Moshier. 1991. Absorption, translocation, and activity of CGA- 136872, DPX-V93 60, and glyphosate in rhizome johnsongrass (Sorghum halepense). Weed Sci. 39:354-357. Derr, 11“., TJ. Monaco, and T.J. Sheets. 1985. Uptake and translocation of fluazifop by three annual grasses. Weed Sci. 33:612-617. Devine, M.D., H.D. Bestman, C. Hall, and W11 Vandenbom. 1984. Leafwash techniques for estimation of foliar absorption of herbicides. Weed Sci. 32:418- 425. Hart, SE. and D. Penner. 1993. Atrazine reduces primisulfilron transport to meristems of giant foxtail (Setan'afaben) and velvetleaf (Abutr’lon theophrastr). Weed Sci. 4 1 :28-3 3. Jansen, LL. 1973. Enhancement of herbicides by silicone surfactants. Weed Sci. 21:130- 135. 47 Manthey, EA, E.F. Szelezniak, ZM. Anyszka, and JD. Nalewaja. 1992. Foliar absorption and phytotoxicity of quizalofop with lipid compounds. Wad Sci. 40:558-562. Menn, W.G. and 6.6. McBee. 1970. A study of certain mstritional requirements for Tifgreen bermudagrass (WMIMXC. tram-miendsLJusinga hydroponic system. Agron. J. 62:192-195. Moxness, KB. and RG. Lyln 1989. Environment and spray additive efl‘ects on picloram absorption and translocation in leafy spurge (Eupharbia esula). Weed Sci. 37: 181-186. Peregoy, RS, L.M. Kitchen, P.W. Jordan, and J.L. Grifin. 1990. Moisture stress effects on the absorption, translocation, and metabolism of haloxyfop in johnsongrass (Sorghum halepense) and large crabgrass (Digital-ta swrgm'nalis). Weed Sci. 3 8:33 1-337. Rhodes, G.N., Jr. and H.D. Coble. 1984. Influence of bentazon on absorption and translocation of sethoxydim in goosegrass (Eleusine indica). Weed Sci. 32:595- 597. Roggenbuck, F.C., L. Rowe, D. Penner, L. Petrofl, and R Burow. 1990. Increasing postemergence herbicide eficacy and rainfastness with silicone adjuvants. Weed Technol. 4:576-580. Wanarnarta, G., J. J. Kells, and D. Penner. 1993. Overcoming antagonistic efi‘ects of Na- bentazon on sethoxydim absorption. Weed Technol. 72322-325. Wills, GD. 1984. Toxicity and translocation of sethoxydim in bermudagrass (Cymdon dacrylon) as afl‘ected by environment. Weed Sci. 32:20-24. Figure 1: Chemical structure for 1,2,6-"C-trinexapac-ethyl. 49 Table 1: Content of absorption study “C tr'eatlnent solutions. Absorption Deionized 320 Cold trinexapac- ‘Y: W Form Adqu 3 gm, ethyl t ethyl Hulk 1’ “I W 57.43 108.8 10 (370 Bq) 23.77 none Plant base ' W13!“ base 7.43 108.8 10 (370 Bq) 23.77 50 with Adjuvant fAmomtslistedtakenfrunaldOOmcezdeioniudmsolution. :Ammthstedtakenrrornal:503ylpnrd309‘:deioniaedmaoluuon Table 2: Plant segmentation in translocation studies Site of Uptake Segments Plant base Roots t Plantbase Treatedleart Leaverabovethetreatedlear Stemoftreatedleaf Sheathorueatedlear Rhizomerwithdmghterplantt Otherroliaget ‘lApplicabletotheSeptanber, l996studybutnotthe1me, 1996M. $Segnentsana1yzedintheltme, 1996andinthe8qstemba’,1996mldies. 51 100 so i so it -i - so 3 so 40 l +Roota 3° ‘ +Laarslada so +Plantaaaa 10 _ o 0 t l 4 s 12 to so 24 .‘. 0 mm Figure 2: Absorption patterns of “C-trinexapac-ethyl over a 24 hour period from three sites of uptake. The patterns are represented by the following regression data: Roots: Y-0.219x + .042 r’-.901 Lear Blade: Y’-lo43.1x“ + 15.9 r‘-.995 Plant Base: Y-o.9olnx + as r’-.951 52 3a so so 70 2! so i 50 g 40 I!) Z) +m ad“! 10 +M 0 t t : -10 ' 16 Z) 24 Tina Ms) 3b +no adlumu +adlmara 8 12 to 20 24 Tina 0101“) Figure 3a,b: The effect of Sylgard 309' on absorption of "C-trinexapac-ethyl by the leaf blade (Figure 3a) and by the plant base (Figure 3b). 53 Table 3: Translocation of “C-trinexapae-ethyl from three sites of uptake. Site of “C-trinexapac—ethyl uptake Site of Analysis LeafBlade Plant Base Roots % of abaubed "C Roots 5 b s 4 c 50 a Foliage t N/A 78 a 25 a Plant Base 11b 18 b 25 a Treated Leaf 67 a N/A N/A Leaf Sheath for 5 b N/A N/A Treated Leaf Stem of Treated 6 b N/A N/A Leaf Leaves Above 1 b N/A N/A Treated Leaf Other Tillers 5 b N/A N/A tReferstoallfoliartissueothertbantheplantbase. tNumbers foflowedbydifl‘auuleuusinechrespecdvewhmmmfisfiaflydifl'amtattbe5% level. Table 4: Translocation patterns of l‘C-tr'iuexapae-ethyl when rh'nomes were present. Site of uptake Site of Analysis Leafblade Plant base % of abaubed "C Foliage t 36 b t 61 a Treated Plant Part 57 a 33 b Roots 4 c 3 d Rhizome/Daughter Plant 3 c 2 d tReferstoallfoliartismeotherthanthetreatedplantpart. tNumbers foflowedbydifl'amtleuusineachrespeefivecolmmsnfisfiaflydifi‘amtuthe5%level. APPENDIX B 55 Table Bl: Contents of stock nutrient solutions. Chamwaeasan Quuifitylsr‘) ICNO; 101.1 ca(NOa)2 236.2 NHtHzPOa 1 15.1 MgSOax 7H20 246.5 SequestreneO 1 1.2 KCl t 3.7 H3303 t . 1.5 MnSOa x H20 t 0.3 ZnSOa x 7HzO t 0.6 CuSOa x 5H20 t 0.1 H2M004 t 0.1 tDenotesreagentswhoseammmtswereallcombinedinasingleliterofmicronuuientstocksolution. Table B2: Contents of full strength Hoagland’s solution. Nutrient Stock Solution Amount (ml 101") KNO: 60 030103): 40 NHJ'12P04 20 MgSOa x 71-120 10 Sequestrene" 10 Micronutrients t 10 1' Containedall sixreagents highlightedinTable l. CHAPTER3 EVALUATION OF V-10029 AND TRINEXAPAC-ETHYL AS TURFGRASS GROWTH REGULATORS ON FIVE COOL-SEASON SPECIES EVALUATION OF V-10029 AND TRINEXAPAC-ETHYL AS TURFGRASS GROWTH REGULATORS ON FIVE COOL-SEASON SPECIES Matthew James Fagerness ABSTRACT An experimental turfgrass growth regulator, V-10029, was compared with trinexapac-ethyl to evaluate growth inhibition patterns and suppression of seedhead formation; the latterwas evaluated forturfgrass speciesand forthe infestingweed species, annualbluegrass. Tall fescue, creepingredfesweKentuckybluegrass, pererulialryegrass, and creeping bentgrass were transferred fi'orn the field into the greenhouse. V-10029, at three rates (0.015 kg/ha, 0.029 kg/ha, and 0.059 kg/ha), was compared to an untreated control and trinexapac-ethyl at the label rate. Of the eight replications for each treatment, four were not mowed for the purpose of evaluating suppression of seedhead formation and four were used for weekly clipping collection to evaluate growth inhibition. Compared to the untreated control, V-10029 at all three applied rates caused significant suppression of seedhead formation in both tall fescue and perennial ryegrass pots. Trinexapac-ethyl was not as efi‘ective. V-10029 caused discoloration in all turfgrass species depending on the rate of application. Patterns of growth inhibition for tall fescue, Kentucky bluegrass, and perennial ryegrass, in response to V-10029 at all rates and to trinexapac-ethyl, were similar on a percent of control basis. Growth of creeping bentgrass was only inhibited by V-10029 at high rates. In contrast, creeping red fescue was significantly injured by V- 10029. The greatest growth inhibition, in response to all treatments, occurred 2 to 3 weeks after application. The efi‘ect of trinexapaeethyl faded after 4 weeks while V-10029 57 58 was eficacious slightly longer. Observed efi‘ects of V-10029 support its known activity as an ALS inhibiting herbicide and therefore as a Class D turfgrass growth regulator. 1552mm: V-10029 {sodium 2.6-bis[(4.6-dimethoxypyfinfidin-2-yl)oxy]bmzoate}. “imam-ethyl {Hcydopmvyl-a-hydmxy-methylweH.S-dioxocyclohmwcarboxyfic acid ethyl ester}, annual bluegrass (Paa m L. ), tall fescue (Fesnroa mordirroceo Schreb.), creeping red fescue (Fesnrca rubra L.) , Kentucky bluegrass (Paa prutensis L.), perminl ryegrass (Lafimmw L). creepins banal!” (WWW Hinds.) mandamus: growth inhibition, suppression of seedhead formation, discoloration, clipping collection INTRODUCTION Plant growth regulators (PGRs) in turfgrass systems are commonly used to reduce vegetative growth, production of seedheads, or both. Watschke (1985) discussed the impact of plant growth regulators that inhibit mitosis, as compared to plant growth retardants that disrupt cell division and elongation through impact on gibberellin biosynthesis. This breakdown of PGR mode of action was later used to classify PGRs as Type I or Type II (Watschke et al. 1992). Type I PGR’s, such as maleic hydrazide {1,2- dihydro-3,6-pylidazinedione} and mefluidide {N-[2,4-dimethyl-5-[[(trifiuoromethyl) sulfonyl]amino]phenyl]acetamide}, cause excellent seedhead formation suppression and growth inhibition in turfgrasses but are also injurious (Schott et al. 1980, McCarty et al. 1985, Diesburg and Christians 1989, Johnson 1989, Spak et al. 1993). Type II PGRs, such as paclobutrazol {(+/-)-(R",R')-B-[(4-chlorophenyl)methyl]-a—(1,l-dirnethylethyl)—1H- 1,2,4-triazole-l-ethanol}, flurprimidol {a-(l-methylethylya-[4-(trifluoro- methoxy)phenyl]-5-pyrimidine-ethanol}, and trinexapac-ethyl, are less injurious but are also less effective in seedhead formation suppression. Watschke and DiPaola (1995) reclassified Type II PGRs, based on the site of uptake and the site of inhibition in the gibberellin (GA) biosynthesis pathway. Trinexapac- ethyl is foliar absorbed, inhibits GA biosynthesis late in the pathway, and is classified as a Class A PGR Paclobutrazol and fiurprimidol are root absorbed, inhibit GA biosynthesis earlier, and are thus classified as Class B PGRs. Mitotis inhibiting PGRs that previously were classified as Type I are now Class C. A final class of PGRs, Class D, includes 59 60 chemicals with herbicidal activity applied at non-lethal rates. V-10029 is a member of this group, showing ALS inhibiting activity (Hopkins 1994, Lirn et al. 1997). Proper classification of both current and future plant growth regulators is essential for proper selection of a PGR to use for specific management needs. Evaluation ofnew plant growth regulators is a function ofseveral factors that, collectively, determine PGR eficacy. Growth inhibition, seedhead formation suppression, visual quality, anddensityintlu'fgrassesarecommonlyevaluatedinPGRstudies. Evaluation of experimental PGRs using a range of applied rates is also common. Schott et al. (1980) tested mefluidide at five rates and at difi‘erent timings of application to evaluate the efi‘ects of mefluidide on shoot growth, seedhead initiation, and turfgrass quality. Johnson (1992) evaluated the persistence of trinexapac-ethyl eficacy in bermudagrass. Many studies with new compounds also include a commercial PGR as a tool for comparison (Hofi'man and Ilnicki 1989, Johnson 1989, Sawyer and Jagschitz 1989). Plant grth regulators have been tested on both cool and warm season turfgrasses. Most of these individual studies concentrate on a limited number of species and/or PGRs. Conversely, few studies directly compare the effects of a given PGR across multiple species. Such studies could lead to more flexible uses for PGRs and be useful in instances where multiple species are afi‘ected by a single application. The objectives of this study were to (a) evaluate the ability of V-10029 to inhibit growth and/or suppress seedhead formation, as compared to trinexapac-ethyl, (b) recommend a chemical rate of V-10029 that, when applied, causes growth inhibition without negatively impacting turfgrass color and quality, and (c) compare five cool-season 61 turfgrass species, with respect to patterns of growth inlu'bition over time, as impacted by V-10029 at three rates and trinexapac-ethyl MATERIALS AND METHODS StudieswereinitiatedinApril, 1996toevaluatethegrowthinhibitingand seedhead formation suppressing abilities of the experimental V-10029 turfgrass grth regulator, as compared to trinexapac-ethyl. Plant material was ‘Penncross’ creeping bentgrass, ‘Viva/Columbia’ Kentucky bluegrass, ‘BrightstarManhattanlDimension’ perennial ryegrass, ‘Triathlon’ tall fescue, and ‘Hector’ creeping red fescue. All species wereirnportedassod plugs 10cmindiameterfi'omtheI-IancockTurfgrassResearch Center in East Lansing, MI to the greenhouse, where they were placed into 946 ml pots containing professional potting medial and allowed to acclimate over a two week period before treatments were applied. The greenhouse was at 25 C +/- 2 C with supplemental lighting fi'om high-pressure sodium lights providing 1200 uEJm’ls of light during 18 hours of daylight. All pots were irrigated as needed and received nitrogen at 5 kyha weekly throughout establishment and the duration of the study. Treatments were applied using a continuous link-belt sprayer at 170 kPa and 230 W spray pressure and carrier volume, respectively, on May 8 and May 9, 1996. V- 100292 as the 80W formulation was applied to each species at three rates: 0.015 kg/ha, 0.029 kyha, and 0.059 kyha. Trinexapac-ethyl’ was applied as the 1E formulation at a ‘Bacctoo isaproductoflldidlingeatCompany,Houston,'lX 2v.10029 isaproductofValentUSACorp.,WalnutCreek,CA 1‘I‘rintrxapllc-ethyl, sold as PrimoO, Ciba-Geigy Caporation, Greensboro, NC 62 63 label rate of 0.382 kg/ha for perennial ryegrass and 0.287 kg/ha for the other four species. A 0.25% v/v level ofx-77“ nonionic surfactant was included in all treatments but the untreated controls. Each treatment had eight replications: four were used to evaluate seedhead formation suppression and four were used to evaluate growth inhibition. Replications evaluated for growth inhibition were maintained at a 4 cm (2 cm for creeping bentgrass) cutting height before chemical applications. Evaluation of growth suppression was a function of clipping fi'esh weight, with clippings harvested weekly for seven weeks after treatment. Replications for seedhead formation suppression were unmowed throughout the course of the study and were evaluated through counts of numbers of seedheads per pot at 3 and 6 weeks afier treatment. Subjective assessment of turfquality was conducted throughout the course ofthe study. The study was a completely randomized design and was repeated. Data reflect combined means from two experiments. All data were analyzed using analysis of variance and/or simple linear regression with significance set at the 5% level. ‘xmo, Valent USA Corp., Walnut Creelt. CA RESULTS AND DISCUSSION S E l E i' 51 . Seedhead initiation was insignificant in Kentucky bluegrass, creeping red fescue, andcreepingbentgrass. Tallfescuepotshadasignificantmimberofseedheads, allof which were annual bluegrass. Only perennial ryegrass produced enough ofits own seedheads to be measurable. Long day light conditions allowed for seedhead formation in perennial ryegrass. Compared to the untreated control, V-10029 at all three rates caused significant suppression of annual bluegrass seedhead formation in tall fescue pots and of perennial ryegrass seedhead formation. Trinexapac-ethyl was not as efi'ective in seedhead formation suppression as V-10029 (Table 1). Growth Inhibig'on V-10029 and trinexapac-ethyl were efi'ective in inhibiting growth in all of the turfgrass species (Tables 2 to 6). The pattern of growth inhibition, as evaluated across PGR treatments, difi‘ered fiom species to species, both in terms of magnitude and persistence of inhibition. Growth of perennial ryegrass, Kentucky bluegrass, and tall fescue was similarly inhibited by all PGR treatments. V-10029 inhibited growth of perennial ryegrass, Kentucky bluegrass, and tall fescue in a rate dependent manner (Tables 2 to 4). The pattern of growth inhibition for trinexapac-ethyl in these species was most analogous to V-10029 at the 0.029 W rate. The growth inhibiting efi‘ects of trinexapac-ethyl and V-10029, at all three rates, were less 64 65 pronounced in creeping bentgrass than in the other species, perhaps due to the tendency for creeping bentgrass to grow laterally when mature (Table 5). Growth of creeping red fescue was inhibited by V-10029 but injury to the turf was largely responsible for this efi‘ect and caused huge variation among treatment replications (Table 6). Trinexapac-ethyl did not injure creeping red fescue but caused little, ifany, growth inhibition. ThegreatestgrowthhuubifioninaflueannentsguleraflyocwnedaroundZoH weeks after treatment (WAT). The efi‘ects of trinexapac-ethyl faded by 4 WAT; however, V-10029, at the 0.029 kg/ha and 0.059 kg/ha rates, was efiicacious for a longer period of time. V-10029 at the 0.015 kyha rate was the least persistent treatment across species with efi'ects lasting only 2 to 3 weeks. V-10029 at the 0.059 kg/ha rate still caused significant growth inhibition in some species when the study concluded 7 WAT but much of this was attributable to unacceptable injury it caused in the turf. Seven WAT, trinexapac-ethyl, V-10029 at 0.015 kg/ha, and V-10029 at 0.029 kg/ha treatments had no longer reduced clipping yield, compared to the untreated control. The predictability of growth inhibition of Kentucky bluegrass by V-10029, as a function of increasing rate, diminished as the study progressed (Figure 1). V-10029 caused significant discoloration in all the turfgrass species at all three rates of application. The specific extent of discoloration was not rated but was characterized by a yellowing of the turf that appeared within a week of application and persisted until 2 to 3 weeks after application. The extent of discoloration increased as the application rate of V-10029 increased. Trinexapac-ethyl treatments did not cause discoloration in any species and the turfgrasses actually assumed a darker green color as soon as 2 weeks after application. 66 Visible injury to the turfgrass species occurred with V-10029 at both the 0.029 kg/ha and 0.059 kg/ha rates. V-10029 at 0.029 kg/ha caused injury but the turfgrass was ableto recoverby4to 5WAT.TheturfgrassinjuredbyV-10029at0.059kglhadidnot firllyrecoverby7WATinatleast 50% ofthetreatrnents. Amongthefivespecies, creeping bentgrass showed the greatest recovery fiom injury caused by V-10029 at 0.059 kg/ha. Injuryobservedwasleaftipburnandreducedmrfdensityperpot. Theextentof injuryobserved seemedtobeproportionaltoamnralbhiegrasspmhferationwithinthepot. Observations made during the study suggested that armual bluegrass, as a weed species, appeared most abundantly in tall fescue and creeping red fescue pots. The creeping red fescue exhibited the least growth among the five species. Annual bluegrass becmnepronunentinsomepouofaeepingredfeswemdcataiMyhnpaaedthedm Annual bluegrass appeared less fi‘equently in pots treated with V-10029 than in control pots or those treated with trinexapac-ethyl. Arumal bluegrass was observed in some of the tall fescue pots but wasn’t abundant enough to have any impact on collected data V-10029 at 0.029 kg/ha and 0.059 kg/ha caused significant growth inhibition but, in many cases, also caused unacceptable discoloration and/or injury to the turf. Therefore, these rateswere detenninedtobetoohighforacceptableturfgrassgromh regulation in the greenhouse. V-10029 at 0.015 kg/ha was the least efi’ective growth regulator treatment, both in terms of extent of growth inhibition and persistence. V-10029, at this rate, however, caused little or no injury to the turf and discoloration was only slight. V- 10029, at an intermediate rate between 0.015 kg/ha and 0.029 kg/ha, would have the greatest potential to combine good growth inhibition with maintenance of acceptable turfgrass quality. A suitable range of application rates that is this narrow would reduce the 67 practicality of V-10029 as a turfgrass growth regulator used for intensively managed turfgrass systems. The extent of seedhead formation suppression caused by V-10029, coupled with its potential to injure turfgrass at higher rates, were useful in supporting the classification ofV-10029asaPGR ClassAandClassBPGRsarenotinjuriousandarenotefi‘ectivein suppressing seedhead formation (Watschke and DiPaola 1995). Conversely, Class C and Class D PGRs, such as mefluidide and glyphosate (N-(phosphonomethyl)glycine), can be injurious and are very effective in seedhead formation suppression. The known activity of V-10029 as an ALS inhibitor confirms its place as a Class D PGR An alternate method of evaluating PGR eficacy was employed by directly comparing the growth suppression patterns for all five species, as influenced by a single PGR treatment (Figures 2 to 5). Perennial ryegrass, Kentucky bluegrass, and tall fescue clipping production, on a percent of control basis, responded similarly to V-10029 at each of the three rates and in response to trinexapac-ethyl at the label rate. Creeping bentgrass was less sensitive to V-10029 than the perermial ryegrass, Kentucky bluegrass, or tall fescue. Creeping red fescue and creeping bentgrass responded similarly to trinexapac-ethyl at the label rate. The injury caused by V-10029, at all rates, to creeping red fescue, however, discounts the merit of including this species in such an evaluation. Ranking the species responses by treatment, would therefore be as follows: V-10029 (all three rates): Kentucky bluegrass == perennial ryegrass = tall fescue > creeping bentgrass Trinertapac-ethyl (label rate): Kentucky bluegrass = perennial ryegrass = tall fescue > creeping red fescue = creeping bentgrass 68 These results may be useful not only in determination of a proper PGR application rate but also in predicting response patterns of difl‘erent turfgrasses to a given PGR application, especiaflymmstanceswherenmlfiplespedesmaycoedumafixedmanaganausyuan (Johnson and Murphy 1996). LIST OF REFERENCES Diesburg, KL. and NE. Cluistians. 1989. Seasonal application of ethephon, flurprimidol, mefluidide, paclobutrazol, and amidochlor as they afl‘ect Kentucky bluegrass shoot morphogenesis. Crop Sci. 29:841-847. Hofinan, K. G. and RD. Ilnicki. 1989. Triasulfuron in combination with some growth regulators in turf. Proc. Northeast Weed Sci. Soc. 43:76. Hopkins, W.L. 1994. P.80. Global Herbicide Directory. Ag Chem Information Services, Indianapolis, IN. Johnson, B.J. 1989. Response of tall fescue (Fes1uca amndinacea) to plant growth regulators and mowing fiequency. Weed Technol. 3:54-59. Johnson, B.J. 1989. Response of tall fescue (Festuca mmdinacea) to plant growth regulator application dates. Weed Technol. 3:408-413. Johnson, B.J. 1992. Response of bermudagrass (Cynodon spp.) to CGA 163935. Wad Technol. 6:577-582. Johnson, B.J. and TR Murphy. 1996. Suppression of a perennial subspecies of annual bluegrass (Poo annua spp. reptans) in a creeping bentgrass (Agrostr‘s stolom’fera) green with plant growth regulators. Weed Technol. 10:705-709. Lim, J.S., Y.T. Bae, 111 Lee, and SJ. Koo. 1997. Mode of acetolactate synthase inhibition of the new herbicide LGC-40863. Abstr. Weed Sci. Soc. Amer. 37:169. McCarty, L.B., J.M. DiPaola, W.M Lewis, and W.B. Gilbert. 1985. Tall fescue response to plant growth retardants and fertilizer sources. Agron. J. 77:476-480. Sawyer, OD. and 1A Jagschitz. 1989. Evaluation of ACP-21 10 as a turfgrass growth regulator. Proc. Northeast Weed Sci. Soc. 43:98-99. Schott, P.E., H. Will, and H.H~ Nolle. 1980. Turfgrass growth reduction by means of a new plant growth regulator. p. 325-328. In 13. Beard(ed.) Proc. 3rd Int. Turfgrass Res. Conf., Munich, Germany. Amer. Soc. of Agron., Madison, WI. Spak D.R., J .M. DiPaola, W.L. Lewis, and CE. Anderson. 1993. Tall fescue sward dynamics: 11. Influence of four plant growth regulators. Crop Sci. 33:304-310. 69 70 Watschke, TL. 1985. Turfgrass weed control and growth regulation. p. 63-80. In F. Lemaire(ed.) Proc. 5th Int. Turfgrass Res. Conf., Avignon, France. 1-5 July. Inst. Natl. de la Recherche Agron., Paris. Watschke, T.L., MG. Prinster, and J.M. Brueninger. 1992. Plant growth regulators and turfgrass management. p. 557-588 In D.V. Waddington et al. (ed.) Turfgrass Science. Agron. Monogr. 32 ASA, CSSA, and SSSA, Madison, WI Watschke, TL. and J .M. DiPaola. 1995. Plant growth regulators. Golf Course Man. 64(3):59-62. 71 Table l: Seedhead formation suppression in tall fescue and perennial ryegrass pots, as a function of turfgrass growth regulator treatment. Tall fescue and Perennial Tall fescue and Perennial Treatment annual bluesms W883 m hm W888 3 WAT‘ 3 WAT 6 WAT 6 WAT lumber ofseorflleads per pot 003m“ 6.0 10.9 12.4 25.0 V-loo29 0.5 6.0 1.1 12.9 (0.015 kg/ha) v-10029 0.4 4,4 1.0 6.6 (0.029 kg/ha) moon 0 3.0 o 5.0 (0.059 kg/ha) Trinexapac-ethyl 1.9 8.5 6.9 15.8 (0.382 kg/ha) 15130.05 2.8 5.5 ‘WATswuksaflertreatrnent 72 Table 2: Growth inhibiting effects of V-10029 and trinexapae-ethyl on perennial WW Time after application (weeks) Treatment 1 2 3 4 5 5 7 clipping puductim (9s ofcontrol} "Comm“ 100 100 100 100 100 100 100 (0.3333,) 57 59 73 79 85 109 111 (03:03:) 47 46 62 73 74 96 99 ($30,338) 35 20 22 32 39 59 63 mm”. «11le ' 58 42 60 85 92 1 14 128 (0.382 kyha) LSDo.os 9 13 17 20 16 34 29 Table 3: Growth inhibiting effects of V-10029 and trinexapae-ethyl on Kentucky bluegrass. Time afier application (weeks) Jilipping poductiar (% of castrol} Um 100 100 100 100 100 100 100 (0.313%) 68 6O 98 97 88 103 93 (01333;) 50 32 57 67 59 81 73 (075303;) 39 23 31 46 58 65 70 WW, - 74 46 61 82 86 101 108 (0.287kg/ha) LSD0.05 18 16 23 27 23 31 33 74 Table 4: Growth inhibiting effects of V-10029 and trinexapae-ethyl on tall fescue. Time after application (weeks) Treatment 1 2 3 4 5 5 7 Clippingpoduction(%ofcontrol} mm 100 100 100 100 100 100 100 Control (0:30:33) 55 53 91 82 102 l 15 102 (0.3330333) 44 25 57 81 86 88 88 (0.0390333) 3 8 l3 3 1 41 5 3 6O 60 “mm, - 69 40 57 74 95 103 107 (0.3821ryha) LSDo.os l4 8 l8 17 20 149 29 75 Table 5: Growth inhibiting effects of V-10029 and trinexapac-ethyl on creeping bentgrass. Time after application (weeks) Treatment 1 2 3 4 5 6 7 Clipping poduction (% of clnrtrolrP mm 100 100 100 100 100 100 100 Control (03:02:13) 95 105 96 100 97 88 100 (03:03:13) 80 89 102 1 17 106 101 100 (03;;0282’18) 54 59 67 94 91 85 87 “$1,“. 64 69 100 103 1 12 120 120 (0.382 kglha) LSDooa l4 14 22 29 24 20 25 76 Table 6: Growth inhibiting effects of V-10029 and trinexapac-ethyl on creeping red fescue. Time after application (weeks) Treatment 1 2 3 4 5 6 7 Clipping poductim (% of control) ”DWI 100 100 100 100 100 100 100 (033033“) 137 88 99 71 71 74 75 (0.979033%) 126 74 69 43 37 38 49 (0059033111) 85 64 76 42 48 48 56 TW‘ 109 75 98 109 113 127 117 (0.382 lrg/ha) LSDo.05 71 36 67 44 46 56 42 rur- olV-lssas Mn) Figure l: Decreased predictability in the patterns of growth inhibition over time in Kentucky bluegrass, as a function of increasing rate of V-10029. LSD.” values for 2, 4, and 6 war were 16, 27, and 31, respectively while 11’ values for the depicted curves for growth inhibition at 2, 4, and 6 WAT were 0.73, 0.47, and 0.25, respectively. . O ,¢,-.--.' .r H o -— o H e' :.‘.‘.L'. .. ‘0 ---Parnnnflryngnaa - --Knntl.lcllybnlaunas 20" Tituscua ------ emu-rou- 0 1 4 t 2 3 4 M‘wwm Figure 2: Growth inhibiting effects of V-10029 (0.015 kg/ha) across four turfgrass species. The LSD“; values for weeks 1, 2, 3, and 4 were 17, 19, 28, and 24, respectively. 79 Thunderwnonaks) Figure 3: Growth inhibiting effects of V-10029 (0.029 kg/ha) across four turfgrass respectively. ‘ l l I A L A 1 J. U I V I I V I emprutluoaonaoroornron ‘ O 1 Y O J .a N a l- & Figure 4: Growth inhibiting effects of V-10029 (0.059 kg/ha) across four turfgrass species. The LSD”; values for weeks 1, 2, 3, and 4 were 13, 15, 21, and 25, respectively. 81 1” -. Figure 5: Growth inhibiting effects of trinexapac-ethyl (0.287 kglha for Kentucky bluegrass; 0.382 kg/ha for perennial ryegrass, tall fescue, and creeping bentgrass) across four turfgrass species. The LSD.” values for weeks 1, 2, 3, and 4 were 34, 19, 32, and 34, respectively. CONCLUSIONS The spectrumoftheresearchdetailedinthisthesiswasdiverse. Assuch, valuable information regarding the performance of turfgrass growth regulators was obtained, using techniques not confined to a specific discipline. Research investigating spray application parameters that may afi‘ecttheperformance oftrinexapac-ethyl illustratedthe multi- faceted benefits of adjuvants, as pertains to eflicacy, rainfastness, and activity in a hard water carrier of trinexapac-ethyl. The potential benefit of adjuvants was filrther explored using "C-labeled trinexapac-ethyl. The direct role of adjuvants in enhancing absorption and the ancillary role of promoting surface movement to alternate sites of absorption are believed to be very crucial in obtaining a good understanding of how trinexapac-ethyl exerts its activity and how eflicacy may be enhanced. Patterns of absorption and translocation for trinexapac-ethyl seem to correspond with activity at the intercalary meristems in turfgrass plants. V-10029 was compared with trinexapac-ethyl as a turfgrass growth regulator and was superior in terms of suppressing seedhead formation. However, levels of growth inhibition were comparable between the two PGRs and more injury was seen with V- 10029 than with trinexapac-ethyl, due to the activity ofV-10029 as an ALS inlubiting herbicide. Species comparisons showed tall fescue, perennial ryegrass, and Kentucky bluegrass to respond similarly to both trinexapac-ethyl and V-10029, ofl'ering some usefulness in growth inhibition predictability where multiple species coexist. 82 I“11301117121110.3031“11100I