.N sen ’1 ”~70 2 - 4 LIBRARY 2 0'54 Michigan State University This is to certify that the dissertation entitled CHARACTERIZATION OF THE ARABIDOPSIS THALIANA CBF1 TRANSCRIPTION FACTOR: FUNCTIONAL ROLE OF TWO EVOLUTIONARILY CONSERVED SIGNATURE SEQUENCES presented by Donatella Canella has been accepted towards fulfillment of the requirements for the Doctoral degree in Cell and Molecular Biologx Q: “4/ 4/74 4A / 'Major P‘Fofessor’s Signature ///7- 6A ‘7" ate MSU is an affinnative-acfion, equal-opportunity employer 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 6/07 p:/CIRCIDaIeDue.indd-p.1 CHARACTERIZATION OF THE ARABIDOPSIS THALIANA CBF1 TRANSCRIPTION FACTOR: FUNCTIONAL ROLE OF TWO EVOLUTIONARILY CONSERVED SIGNATURE SEQUENCES By Donatella Canella A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Program in Cell and Molecular Biology 2007 ABSTRACT CHARACTERIZATION OF THE ARABIDOPSIS THALIANA CBF1 TRANSCRIPTION FACTOR: FUNCTIONAL ROLE OF TWO EVOLUTIONARILY CONSERVED SIGNATURE SEQUENCES By Donatella Canella Over the course of evolution plants have adapted to the environment by developing ways to cope with different biotic and abiotic stresses. Among the abiotic stresses, low temperatures represent a major limiting factor for the growth and development of plants. The CRT /DRE-Binding Factors (CBFs) are transcriptional activators that are rapidly activated in response to low temperature and in turn induce the expression of a battery of cold-regulated (COR) genes to increase plant freezing tolerance. Arabidopsis plants overexpressing CBF1, CBFZ or CBF 3 are constitutively freezing tolerant, indicating that these regulators are master regulators of cold adaptation. CBF proteins belong to the APETALAZ/Ethylene-Response Binding Protein (AP2/EREBP) family, which includes 145 members in Arabidopsis. Proteins in this family share high similarity within their AP2/EREBP DNA-binding domain. A unique feature of the CBF proteins is that they contain two conserved sequences flanking the DNA-binding domain. These sequences, represented by the consensus motifs PKK/RPAGRxKFxETRHP and DSAWR, are also found in CBF-like proteins fi'om evolutionarily diverse plant species, suggesting that these “signature” sequences play a role in CBF activity. Overexpression of wild type CBF1 in Arabidopsis results in constitutive COR gene expression. Transgenic lines overexpressing CBF1 carrying alanine substitutions in the signature sequences showed reduced or no COR gene expression, indicating that these sequences play an important role in CBF1 activity. Analysis of protein levels revealed that alanine mutations throughout the DSAWR motif affect protein accumulation in planta, and could explain the lower COR gene accumulation in those plants. On the contrary, mutations in the PKKPAGR motif did not affect protein steady state levels; instead they impaired the ability of CBF1 to bind its cognate CRT/DRE element from the COR gene promoters. The most pronounced effect was observed when two conserved Arg and Phe residues were substituted with Lys and Ala, respectively; these substitutions were sufficient to abrogate DNA binding, indicating an essential role of those residues and potentially a base-specific recognition. Altogether these results indicated that DNA binding activity in the AP2/EREBP family extends beyond the canonical DNA-binding domain previously described to include the N- flanking PKKPAGR region. Based on these observations and secondary structure prediction studies, we developed a computational model describing CBF1 bound to the DNA. According to this model, CBF1 binds the DNA major groove through a three- stranded beta sheet, as described for other AP2/EREBP proteins. In addition, a helical stretch within the PKKPAGR motif makes essential interactions with the DNA minor groove in close proximity to a conserved thymine that is a specificity determinant in the CRT/DRE element bound by the CBF proteins. Additional investigations will elucidate whether residues within the PKKPAGR motif represent specificity switches for the recognition of the CRT/DRE promoter element by CBF proteins and whether a similar mechanism has been conserved in other protein of the APZ/EREBP family. ACKNOWLEDGMENTS I came to the United States to work in the Plant Research Laboratory as a visiting scholar. The plan was to stay one year. Several years later, I prepare to leave with a doctoral degree. Being here has been a great experience, at the professional and personal level, and I am very gratefiil to the many people that have accompanied me in this amazing journey. First of all, I want to thank my mentor, Mike Thomashow, for his great guidance and support. I am very grateful for the freedom, trust and encouragement that he has given me all throughout my PhD. I feel privileged to have met him and very fortunate to have worked in his lab. I have learned a lot from our meetings and discussions; these times will always stay with me wherever I’ll go. I would like to thank my committee members: Sheng Yang He, Lee Kroos, Ken Keegstra, and Steve Triezenberg. Thank you for your time and your constructive criticism. Your advice over the years has been critical in the completion of my thesis work and has been an important contribution to my growth as a scientist. I want to thank Leslie Kuhn, with whom I have collaborated for the last few years of my PhD. It has been a great pleasure to work together; I have greatly benefited from your scientific experience, and your encouraging attitude had made me a more confident scientist. Thank you to all the present and past members of the Thomashow lab: Sarah Gilmour, Keenan Amundsen, Marcela Carvallo, Diane Constan, Daniel Cook, Colleen Doherty, iv Malia Dong, Carri Duncan, Sarah Fowler, Chin-Mei Lee, Michael Mikkelsen, Susan Myers, Kanchan Pavangadkar, Ritu Shanna, Bonnie St.John, Heather Van Buskirk, Jonathan Vogel, Vandana Yadav, Dan Zarka and Xin Zhang. I can’t believe I won’t be coming to the lab next week. I will miss you. A very big thank you to all my friends. It’s been wonderful to have you in my life. Thanks to each and all of you for being with me during all this time and for making my life so much richer and happier. It’s amazing to think of how many little everyday things we have shared; all these times are such a big part of who I am today; there’s no doubt in my heart I could not have come so far without you. A special thank you to Alberto, who has been by my side with his love and support. Thank you for never doubting that I could make it. Your strength, patience and constant presence have been a wonderful gift that will stay with me forever. A warm thank you to Kathryn and Edward Boucher, my American family. You have been my home away from home. Thank you for the many suppers and for our chats after supper. Your positive and strong spirit has touched me deeply; I will never forget it. My family has always been there for me. Thank you mom, dad, Robi and Cira. You have been my rock with your endless love, constant presence and encouragement. Thank you for believing in me and being proud of me, no matter what. Thank you very much. TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... viii LIST OF FIGURES ........................................................................................................... ix CHAPTER ONE Literature Review Freeze-induced cell and role of cold acclimation .................................................. 1 Changes in gene expression underline the cold acclimation response ................... 3 The CBF cold-response pathway in Arabidopsis thaliana ..................................... 4 Upstream events in the regulation of the CBF pathway ........................................ 9 Conservation of the CBF pathway in other plant species .................................... 10 Mechanisms of transcriptional activation in the CBF family of proteins ............ 12 Conclusion .................................................................................. 18 Literature Cited ........................................................................... 19 CHAPTER TWO Functional roles of the PKKPAGR and DSAWR sequences in CBF1 transcriptional activity Introduction .......................................................................................................... 27 Results .................................................................................................................. 32 FUNTIONAL ROLE OF THE PKKPAGR MOTIF IN CBF1 ACTIVITY ........ 37 Mutations in the PKKPAGR motif affect COR gene activation .......................... 37 Reduced COR gene induction in the pkkpagr lines is not caused by lower protein levels ........................................................................................ 43 Role of the PKKPAGR motif in nuclear targeting of CBF1 ................................ 47 Identifying regions in CBF1 involved in nuclear transport ................................. 51 Role of the PKKPAGR motif in DNA binding .................................................... 57 Secondary structure predictions to investigate structural the preferences of the PKKPAGR motif .................................................................................................. 61 Specific amino acids within the PKKPAGR motif have important roles in binding to the CRT/DRE promoter element ........................................................ 67 Side chain conservation in AP2/EREBP-like proteins that use their flanking sequences to stabilize DNA binding ..................................................................... 73 A working model for DNA binding by CBF1 .......................... . ................ 80 FUNTIONAL ROLE OF THE DSAWR MOTIF IN CBF1 ACTIVITY ............ 87 Mutations 1n the DSAWR sequence affect COR gene activation ........................ 86 CBF1 protein detection In dsawr lines .................................................. 89 Role of the DSAWR motif 1n DNA binding activity by CBF1 ...................... 91 Point mutations within the DSAWR region from CBF 2 reveal the importance an Asp residue to CRT/DRE recognition by CBF proteins ................................. 93 DISCUSSION and FUTURE DIRECTIONS ...................................................... 98 vi MATERIAL AND METHODS ......................................................................... 106 Mutagenesis of the PKKPAGR and DSAWR motifs ........................................ 106 Preparation of constructions for plant transformation .............................. 108 Growth conditions and Northern blot analysis of transgenic Arabidopsis plants .................................................................................................................. 110 Analysis of variance (ANOVA) for COR/CBF1 transcript ratios ..................... lll Subcloning of CBF1 mutant ORFs into bacterial expression vectors for expression and purification of 6xHiszCBF1 and MBP:CBF1 proteins .............. 112 Protein isolation and immunoblot analyses ........................................................ 114 Electro-Mobility Shift Assays (EMSA) ............................................................. 115 Fluorescence imaging of Arabidopsis root tips overexpressing CBF1 :GFPzGUS transgenes ................................................................................. 116 Literature Cited ........................................................................... l 18 vii LIST OF TABLES Table 2.1. Secondary structure predictions of the PKKPAGR motif ......................... 63 Table 2.2. Sequery analysis of the crystallographic structures observed for tetrapeptide sequences in the PKKPAGR region of CBF1, and mutations to AAA in this region that were tested experimentally ........................................................................ 65 Table 2.3. Quantification of CRT /DRE-containing probes bound by CBF1 wild type and mutated proteins .................................................................................... 75 Images in this dissertation are presented in color. viii LIST OF FIGURES Figure 2.1. PKKPAGR and DSAWR signature sequences in the CBF family of plant transcription factors ................................................................................. 28 Figure 2.2. Site-directed mutagenesis was used to generate CBF1 ORFs containing specific mutations or deletions of the signature sequences ................................... 33 Figure 2.3. Arabidopsis pqupagr and dsawr lines ............................................. 35 Figure 2.4 A. Analysis of COR gene induction in APKK, M1 and M2 pkkpagr lines by Northern blot analysis ....................................................................................................... 38 Figure 2.4 B. Analysis of COR gene induction in M3, M4 and M5 pkkpagr lines by Northern blot analysis ....................................................................................................... 39 Figure 2.5A. Analysis of variance of COR78/CBF1 transcript ratios in transgenic plants overexpressing wild type or mutated CBF1 transgenes ....................................... 41 Figure 2.5B. Analysis of variance of COR6.6/CBF1 transcript ratios in transgenic plants overexpressing wild type or mutated CBF1 transgenes ....................................... 42 Figure 2.6. Comparison of CBF1 transcript and protein levels in Arabidopsis plants overexpressing 6xMyc:CBFl transgene ......................................................... 45 Figure 2.7. Northern blot analysis of Arabidopsis seedlings overexpressing different CBF1 :GFPzGUS constructs ........................................................................ 48 Figure 2.8. Localization studies of CBF1 :GFPzGUS chimeras .............................. 50 Figure 2.9. Localization studies of 5’ deletions of CBF1 :GFPzGUS chimeras ............ 54 Figure 2.10. Western blot analysis of total protein extracts from Arabidopsis seedlings overexpressing different CBF1 :GFPzGUS constructs or GFPzGUS alone .................. 56 Figure 2.11. Triple alanine mutations abrogate the binding activity of a full length 6xHis-CBF1 protein in vitro. ..................................................................... 58 Figure 2.12. Triple alanine mutations abrogate the binding activity of MBP-CBFZH 12 proteins in vitro ...................................................................................... 60 Figure 2.13 A. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR 78 gene promoter ..................... 69 ix Figure 2.13 B. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR15a gene promoter ..................... 70 Figure 2.13 C. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR6. 6 gene promoter ..................... 71 Figure 2.14. Three-dimensional structures of proteins that use their three-stranded B- sheets to recognize the major groove of their target DNAs .................................... 77 Figure 2.15. Comparison of B-sheet—flanking sequences experimentally shown to be important in DNA binding .......................................................................... 78 Figure 2.16A. NMR structure of the AP2/EREBP DNA-binding domain of ERF] bound to its cognate DNA .................................................................................. 81 Figure 2.16B. Working model describing DNA binding by CBF1 .......................... 82 Figure 2.17. Nucleotide sequences of the cis-acting elements in CBF and ERF proteins. ............................................................................................................................ 85 Figure 2.18. Accumulation of CBF and COR transcripts in Arabidopsis plants overexpressing wild type or dsawr CBF1 ........................................................ 88 Figure 2.19. Northern and western blot analysis of dsawr plants and plants overexpressing wild type CBF1 ................................................................... 90 Figure 2.20. DNA binding activity of an MBP-CBFl protein carrying alanine substitutions in the DSAWR region ............................................................... 92 Figure 2.21. Northern blot analysis of transgenic plants overexpressing wild type and mutated CBF2 transgenes ........................................................................... 94 Figure 2.22. DNA binding activity of MBP:CBF2 proteins carrying point mutations in the DSAWR motif ................................................................................... 96 CHAPTER ONE LITERATURE REVIEW Plant growth and development can be greatly affected by low temperatures. Plants differ in their ability to cope with low temperatures. Plants native to tropical regions, such as tomato and rice, are typically very sensitive to low temperatures and will suffer chilling injuries when temperatures drop to the range of 0-12°C. In contrast, plants originating from temperate regions, such as the model plant Arabidopsis thaliana, rye and canola, are generally more resistant, and are not only chilling tolerant but also freezing tolerant (Sakai and Larcher, 1987). The ability of plants to tolerate freezing can be significantly enhanced by pre-exposure to low non-freezing temperatures, an adaptive process called cold acclimation. For instance, upon cold acclimation wheat can increase freezing tolerance from -5°C to -20°C (Thomashow, 1998). The events occurring inside a plant upon cold acclimation reflect a complex network of changes that the plant mounts against freeze-induced damage (Levitt, 1980; Thomashow, 1999). These include changes at the physiological, biochemical and transcriptional level. In this chapter, I will summarize the advances in the field of cold acclimation, with specific focus on the CBF family of transcriptional regulators and their role in the CBF cold response pathway in the model plant Arabidopsis. Freeze-induced cell damage and role of cold acclimation. The negative effects of freezing temperatures on cell survival have been long known (Levitt, 1980). The plasma membrane is the primary target of freeze-induced cell damage. This injury is mainly caused by freeze-induced dehydration, which occurs in the cell upon exposure to sub-zero temperatures. The first event following freezing inside the cell is ice formation in the extracellular space; this is due to the presence of ice nucleating agents and lower solute concentration in the extracellular space relative to the intracellular space. This process can cause membrane damage due to adhesion of membranes. However, the most severe damage results from freeze-induced dehydration. When ice crystals form at sub-zero temperatures, the chemical potential in the outer space drops, and causes an outward movement of intracellular water to the outer space. The water loss can be as severe as 90% at approximately -10°C (Thomashow, 1999). The type of injury at the membrane site varies depending on the freezing temperatures: freeze-thaw cycles between -2°C to -4°C can cause expansion-induced cell lysis; a temperature range of —4°C to -10°C typically causes phase transition of bilayer lipids from larnellar to hexagonal II; in the most severe cases (below -10°C), fracture jump lesions will arise (Steponkus and Webb, 1992). Membrane damage induced by freezing temperatures is greatly reduced upon cold acclimation. The protective role of cold acclimation has been under scrutiny for many years, and includes a plethora of changes at the biochemical, physiological and transcriptional levels. Studies by several investigators have established that one of the roles of cold acclimation is to prevent cellular damage by stabilizing membranes. Changes in membrane lipid composition are among the first and best documented events that underline cold acclimation (Thomashow, 1999). Additional metabolic changes also accompany cold acclimation, including the accumulation of small cryoprotective molecules such as soluble sugars and proline (Rudolph and Crowe, 1985; Strauss and Hauser, 1986; Carpenter and Crowe, 1988). Changes in gene expression underline the cold acclimation response. In more recent years, changes in gene expression have been correlated to adaptation to low temperatures. The first direct evidence that cold acclimation was accompanied by changes in gene expression was presented by Guy and colleagues (1985), who reported that differential pools of translatable mRNAs accumulate in non- acclimated and acclimated spinach leaves (Guy et al. , 1985). This discovery was followed by the identification of numerous cold-responsive genes in other plant species, including Brassica napus (J ohnson-F lanagan and Singh, 1987), alfalfa (Mohapatra et al. , 1987) and Arabidopsis (Gilmour et al., 1988; Kurkela et al. , 1988). Those genes were given names such as low temperature-induced genes (LT 1), cold-regulated genes (COR), KIN (cold-induced), or RD (responsive to dehydration) (reviewed in Thomashow, 1999). Analysis of their expression patterns revealed a positive correlation between their transcript accumulation in response to low temperatures and the level of freezing tolerance achieved by plants (Guy and Haskell, 1987; Mohapatra et al. , 1987). Despite their highly divergent sequences, the proteins encoded by these genes share a set of properties that suggest that they might share a common mechanism of action in protecting the cell against freezing. Indeed, some of those properties have been previously described for cryoprotectant molecules, and include high hydrophilicity, presence of amphipathic helices, and solubility in aqueous buffers at high temperatures (V olger and Heber, 1975). The observation that changes in gene expression accompany cold acclimation suggested that such changes might represent the initial signal regulating the events required to achieve freezing tolerance. However, overexpression of COR15a in Arabidopsis had little effect on overall plant freezing tolerance, despite its significant role in protecting the chloroplast envelope (Steponkus et al. , 1998). This observation indicated that individual COR genes were not sufficient to trigger the whole set of changes induced by cold acclimation. It was soon hypothesized that an upstream regulator of gene expression was responsible for the activation of a whole set of COR genes. Consistent with this idea, characterization of a battery of COR genes from Arabidopsis suggested that they might be coordinatively regulated in response to low temperatures. Consistent with this idea, COR6. 6, COR15a, COR4 7, and COR 78 expression followed similar kinetics; they were upregulated as early as four hours after a treatment at 4°C, reached a peak at about 24 hours and remained upregulated as long as the plants were kept under inducing conditions (Hajela et al. , 1990) (Horvath et al. , 1993). Furthermore, analysis of COR gene promoters unveiled the presence of a common cis-regulatory element, named C-repeat/Drought Responsive Element CRT/DRE (Baker et al., 1994; Yamaguchi-Shinozaki and Shinozaki, 1994). Similar induction patterns and the presence of a conserved element in their promoters suggested that the COR genes might be regulated by a common factor. The CBF cold-response pathway in Arabidopsis thaliana. Discovery of the CBF pathway. A major breakthrough in the field of cold acclimation was the discovery of the transcriptional regulator CRT/DRE Binding F actorl (CBF1) (Stockinger et al., 1997) and its key role in regulating cold acclimation in Arabidopsis. CBF1 was identified in a yeast- one-hybrid screen in which the CR T /DRE promoter element was used as bait. CBF1 could specifically bind this cis-element in vitro and induce the expression of a COR15a::lacZ reporter gene in yeast (Stockinger et al., 1997). Arabidopsis plants overexpressing CBF1 constitutively expressed the whole battery of COR genes and displayed constitutive fi'eezing tolerance similar to freezing tolerance of cold-acclimated non-transgenic plants, indicating that this transcription factor is a master regulator of the cold acclimation response in Arabidopsis (Jaglo-Ottosen et al. , 1998). The discovery of CBF1 and its role in cold acclimation was significant because it provided the first evidence that genetic manipulation was indeed possible to generate more freezing tolerant plants, whereas years of breeding attempts had resulted in only modest improvement of freezing tolerance. Furthermore, it revealed that the whole set of changes necessary to induce freezing tolerance could be induced by overexpressing a single gene (Thomashow, 1999). Shortly after, it was discovered that Arabidopsis has three CBF genes, CBF1, CBF2, and CBF3 — also known as DREBI B, DREBI C, and DREBIA, respectively - arranged in tandem array on chromosome IV (Gilmour et al. , 1998; Liu et al. , 1998; Gilmour et al. , 2000). They follow a similar induction pattern in response to low temperatures, whereby they are induced within 10-15 minutes of cold treatment, peak at 2-3 hours, and return to a basal expression level by 24 hours of cold treatment. Overexpression of CBF1, CBF2, or CBF 3 in Arabidopsis rendered plants constitutively freezing tolerant (Jaglo-Ottosen et al., 1998; Liu et al. , 1998; Gilmour et al., 2000; Gilmour et al. , 2004) and induced constitutive expression of the same pool of genes, indicating that the three transcriptional activators have matching transcriptional activities. Three additional CBF homologs were identified more recently. CBF 4/DREBI D (Haake et al. , 2002; Sakuma etal., 2002), was described as a regulator of drought response in Arabidopsis, in that its expression was up-regulated by drought stress, but not by low temperature. Its overexpression in Arabidopsis resulted in constitutive accumulation of CRT /DRE-containing genes, and conferred to plants both drought and freezing tolerance, similarly to what observed for overexpression of the other CBFs (Haake et al. , 2002). This result is not surprising, given the physiological overlap between dehydration and freezing stress. DDF I/DREBI F and DDF 2/DREBI E (Sakuma et al. , 2002; Magome et al. , 2004) also show high sequence homology to the CBF family, but little is known on their role in response to abiotic stresses. DDF I/DREBI F expression is induced upon salinity stress (Sakuma et al. , 2002; Magome et al. , 2004), similarly to what happens to CBF1-3, suggesting that these genes may display overlapping activity. The role of DDF I/DREBI F and DDF 2/DREBI E in low temperature response, however, has not been elucidated. Predominant role of CBF in configuring low temperature responses in Arabidopsis. Technological advances and the availability of whole genome sequences have made it possible to conduct high-throughput experiments to identify the molecular changes at the transcriptional and post-transcriptional level occurring in response to low temperature. Microarray technology has proven a powerful tool for surveying global gene expression in Arabidopsis in response to cold treatment. Transcriptome profiling studies using cDNA microarrays and Affymetrix GeneChips have been very usefirl to identify novel genes, determine their kinetics of induction, and help delineate the complex network of cascades that are configured. These analyses have revealed that hundreds of genes are regulated during cold acclimation in Arabidopsis (Seki et al. , 2001; Fowler and Thomashow, 2002; Kreps et al., 2002; Seki et al., 2002; Gilmour et al., 2004; Maruyama et al., 2004; Vogel et al. , 2005). Among the up-regulated genes are several transcription factors, some of which are induced in parallel with CBF. An important conclusion from these studies has been that additional cold response pathways exist besides the CBF pathway (Seki et al., 2001; Fowler and Thomashow, 2002; Kreps et al. , 2002; Seki et al. , 2002). This is consistent with previous reports of Arabidopsis mutants that are constitutively more freezing tolerant than wild type plants despite the fact that COR genes are not induced. One example of a CBF-independent response is represented by the eskr'moI Arabidopsis mutant (Xin and Browse, 1998). eskimoI plants display constitutive freezing tolerance and yet do not show constitutive expression of the CBF regulon (Xin and Browse, 1998; Xin et al., 2007). Indeed, transcriptome profiling analysis showed an overlap of about 12% in gene transcripts between the eskimoI mutants and CBF2- overexpressing plants (Xin et al., 2007). Interestingly, eskimoI plants are not drought tolerant, which makes them a valuable tool for understanding the players and mechanisms that are specific for the cold response in Arabidopsis. Similarly to the eskimoI mutation, ada2b mutant plants in Arabidopsis are constitutively more freezing tolerant than wild type plants, and yet they do not show constitutive COR gene expression (V lachonasios et al. , 2003); therefore ADA2b is implicated in a CBF- independent pathway. Despite the presence of additional cold response pathways, the CBF family plays a predominant role in configuring the changes observed when plants respond to cold. Comparison between genes that are cold- and CBF-responsive has helped define the CBF regulon as the pool of approximately 100 genes that are regulated by both low temperature and CBF (Fowler and Thomashow, 2002; Vogel et al. , 2005). Eighty five genes in the CBF regulon are up-regulated and account for almost 30% of the transcripts that accumulate in response to low temperatures. In addition, members of the CBF regulon are among the most highly induced upon cold treatment (V ogel et al. , 2005). For instance, 84% of the cold-responsive genes that are up-regulated more than 15 fold and 50% of the genes that are up-regulated between 5-10 fold belong to the CBF regulon. Metabolite profiling in Arabidopsis has enabled the analysis of changes occurring at low temperature to changes in small organic compounds associated with low temperature response. The role of CBF in mediating metabolic changes that occur upon cold exposure was monitored by metabolomic profiling of wild type Arabidopsis plants and plants overexpressing CBF 3 . Major rearrangements were detected during cold acclimation, with 75% of the metabolites analyzed showing changes. Of those changes, the majority (79%) could be attributed to CBF 3 overexpression (Cook et al. , 2004). Therefore, the CBF pathway plays a major role in configuring the changes that are needed for plants to adapt to low temperatures, both at the transcriptional and post- transcriptional level. Upstream events in the regulation of the CBF pathway. Much of the research effort in the field of cold acclimation has focused on elucidating the upstream events regulating the induction of CBF genes. Reporter gene analyses and mutant screens have proven very valuable to identify novel genes that play important roles during cold acclimation. Promoter analysis of CBF2 unveiled a minimal 125-bp promoter region that is cold responsive (Zarka et al. , 2003). Within this promoter fragment, two regions named ICErl and ICEr2 (induction of CBF expression region 1 and 2) cooperated in directing cold regulation of CBF expression (Zarka er al., 2003). ICErl contains a cis-acting element which is recognized by ICEl (Inducer of CBF Expression 1). ICE] was isolated through a mutant screen of plants that were defective in CBF 3 cold-induction; however, CBF1 or CBF 2 expression was not affected, suggesting that one of the other 139 bHLH proteins present in Arabidopsis might be involved in the specific recognition of this element at low temperatures (Chinnusamy et al., 2003). Interestingly, overexpression of ICE] is not sufficient to confer constitutive freezing tolerance, but a cold stimulus is required. Presumably, post-translational mechanisms or the presence of a co-activator must occur to activate ICEl (Chinnusamy et al. , 2003). More recently, it was discovered that a transcriptional regulator member of the MYB family, MYB15, can bind to the promoter of CBF1, CBF 2, and CBF 3, and physically interacts with ICEl in vitro (Agarwal et al., 2006). MYB15 overexpression negatively regulates CBF expression in the cold and is most evident in the early stages of induction (approximately 3 hours). However, MYB15 overexpression does not significantly influence COR gene expression (Agarwal et al., 2006). The significance of these results remains unclear at the present. Recently, Vogel et al. (2005) have characterized ZAT12, a zinc-fmger transcription factor that appears to be coordinately regulated with CBF2 in response to low temperature. Interestingly, ZAT12 overexpression in Arabidopsis has a negative effect on cold-regulated induction of the CBF genes, whereas in the zat12 knock-out lines CBF expression is induced. Presmnably this repressive activity is mediated by an EAR- like motif present in ZAT12 (Ohta et al. , 2001). At the same time, ZAT12 induces the expression of a Z4T12 regulon, which partly overlaps with the CBF regulon, suggesting that CBF and ZAT12 might cooperate to induce expression of cold responsive genes (V ogel et al., 2005). It is evident from the recent advances that the regulation CBF expression during cold stress is complex and involves many factors. Identification of additional components and a better understanding of their mechanisms of action will help gain a more detailed picture of how transcriptional networks are regulated in response to low temperatures. Conservation of the CBF pathway in other plant species. A major interest in the field of cold acclimation has been to determine whether the CBF pathway has been conserved in other plant species. This is of particular interest for agronorrrical application to crop plants. CBF-like genes have been identified in a wide variety of plant species, including both chilling sensitive and freezing tolerant plants. A central question is to determine whether these represent functional homologs of the Arabidopsis CBFs, and whether any differences in the CBF pathway of these plants can explain different tolerance to low temperatures. 10 Among freezing tolerant plants, CBF-like genes have been identified in Brassica napus (Jaglo et al., 2001), poplar (Benedict et al. , 2006), barley (Skinner et al. , 2005), maize (Qin et al., 2004) and others. Some of these CBFs have been shown to be functional homologs of the Arabidopsis CBFs based on their sequence similarity, transcriptional activity, and the ability to enhance freezing and drought tolerance when overexpressed. Furthermore, a number of CR T/DRE-containing genes present in those plants and their transcripts can be constitutively accumulated when CBF is overexpressed (Choi et al., 1999; Dal Bosco et al., 2003). Chilling sensitive plants also contain CBF-like genes. Some of these genes represent functional homologs of the Arabidopsis CBFs, since they are cold-responsive and can induce constitutive expression of a CBF regulon when overexpressed in Arabidopsis, and results in increased tolerance to freezing and dorught. Some examples include CBFs from tomato (Hsieh et al., 2002; Zhang et al. , 2004) and rice (Dubouzet et al., 2003). The tomato CBF locus includes three CBF genes (Zhang et al., 2004). LeCBFI is cold-responsive and can induce constitutive expression of a CBF regulon and irrrpart freezing tolerance when overexpressed in Arabidopsis. Tomato plants overexpressing LeCBFI or AtCBFI can also activate a CBF regulon. However, microarray analysis of a quarter of the tomato genome indicates that, despite the fact that similar classes of proteins are induced in the two plants, transgenic tomato plants can only slightly increase their chilling tolerance (Zhang et al. , 2004). Current efforts are in place to determine whether the inability of chilling sensitive plants to show a functional CBF response can be attributed to smaller and consequently less diverse CBF regulons or additional defects that limit CBF function. 11 Mechanisms of transcriptional activation in the CBF family of proteins. The CBF family of proteins. CBF proteins are members of the AP2/EREBP family of transcription factors (Riechmann and Meyerowitz, 1998). Members of this family are widespread among plants, but a few examples have been recently found in other organisms such as bacteria and bacterial viruses (Magnarri et al., 2004; Wuitschick et al. , 2004; Balaji et al. , 2005). AP2/EREBP proteins play a variety of roles in plant growth and development as well as in biotic and abiotic stress responses (Riechmann and Meyerowitz, 1998). A common feature of these proteins is the presence of a conserved AP2/EREBP DNA-binding domain. The Arabidopsis AP2/EREBP family includes 145 predicted proteins sharing high homology within this domain. Based on the sequence similarity within the DNA- binding domain, Sakuma et al. (2002) have divided the Arabidopsis AP2/EREBP family into 5 different groups: the DREBs (56 members); the ERFs (65 members); the AP2$ (14 members); the RAVs (6 members); and a fifth group (4 members). The DREB group includes the CBF family. CBF proteins are characterized by an N-terminal domain including 32 amino acids of unknown function, the AP2/EREBP DNA-binding domain, and a C-terminal domain resembling acidic activation domains from other transcriptional activators such as the herpesvirus VP16, the mammalian p53 protein and RelA (Triezenberg, 1995). A distinctive feature of the CBF family in Arabidopsis is the presence of two conserved signature sequences flanking the DNA-binding domain, defined by the consensus PKKP/RAGRxKFxETRHP and DSAWR (J aglo et al., 2001). Strikingly, these sequences have been highly conserved throughout evolution, and they can be found in very diverse plant species (J aglo et al., 2001). This Observation suggests 12 that they might play an important role in mediating CBF transcriptional activity. The functional characterization of these motifs will be presented in this dissertation. DNA-binding properties of CBF proteins and other APZ/EREBP family members. The DNA-binding activity of the AP2/EREBP proteins was frrst described by Ohme-Takagi and Shinshi (1995). They reported that a 59-amino acid region that was highly conserved among four ethylene-binding proteins was required for specific binding to their target DNA, called the GCC box (Hart et al., 1993; Ohme-Takagi and Shinshi, 1995; Sato et al., 1996). Shortly after, additional transcription factors with the same conserved domain were identified, and the conserved region named AP2/EREBP DNA- binding domain (Riechmann and Meyerowitz, 1998). One of the most significant contributions towards characterizing the binding activity of the AP2/EREBP proteins has been the NMR analysis of this domain from AtERFl bound to its target cis-regulatory element. This domain is defined by a three stranded B-sheet followed by a C-terrninal helix; the B-sheet is packed against the major groove of the DNA and makes specific contacts through seven amino acids, mainly arginines and tryptophans (Allen et al. , 1998). The degree of similarity among the Arabidopsis AP2/EREBP proteins ranges from approximately 60 to 90% and six of the seven amino acids involved in direct DNA contact in ERF 1 are conserved in the AP2/EREBP proteins (Hao et al. , 2002). The high degree of conservation implies that these domains fold similarly and might make similar DNA contacts. Very little information is available on the DNA binding preferences within the AP2/EREBP family. One of the reasons is that the target cis-acting elements have been 13 identified only for a few AP2/EREBP proteins, including ABI4, ORCA, CBFs, DREBZS, and ERFs (Stockinger et al. , 1997; Allen et al., 1998; Liu et a1. , 1998; van der Fits and Memelink, 2001; Acevedo-Hernandez et al. , 2005). Most of the current knowledge on the binding properties of the AP2/EREBP family has been derived from the analysis of the ERF and CBF proteins. The cis-acting elements targeted by the CBF and ERF proteins have long been known and are represented by the CR T/DRE and GCC box promoter elements, respectively (Baker et al., 1994; Yamaguchi-Shinozaki and Shinozaki, 1994; Ohme-Takagi and Shinshi, 1995). The specificity determinants in these two cis-acting elements are defined by the two core sequences A/GCCGACNT and AGCCGCC for the CBF and ERF family, respectively (Hao et al., 1998; Hao et al., 2002; Maruyama et al. , 2004). Much remains to be understood in terms of specificity switches within proteins of the two subfamilies. Most of the work has focused on elucidating the role of amino acids within the AP2/EREBP domain that differ between ERFs and CBFs. One of the most important findings is that an alanine residue that is highly conserved in the ERF subfamily can impart specific sequence recognition to proteins in this subgroup. In fact, when alanine is substituted with the corresponding valine residue from the CBF proteins, the mutated ERF alters its binding specificity and can recognize the CR T/DRE promoter element (Hao et al. , 2002). However, there is no evidence indicating the corresponding mutation in the AP2/EREBP domain of CBF can modify its binding preference. Sakuma et al. (2002) have proposed that two conserved amino acids within the AP2/EREBP domain, which differ between ERFs and CBFs, function as specificity determinants. The authors showed that swapping of the residues between the two groups of proteins could abrogate binding to their respective cis-acting elements; however, there is no direct l4 evidence that residue swapping can drive CBF proteins to the ERF promoter. Based on the current studies, it is unclear whether additional residues are essential for DNA binding within the AP2/EREBP family. Hao et al. (1998) observed that short stretches of amino acids flanking the DNA-binding domain of several ERF proteins are important for binding to the GCC box. Since the flanking regions of ERF proteins share little or no similarity, it was suggested that the role of these residues is to stabilize the protein-DNA complex (Hao et al., 1998). Characterization of CBF1 trans-activating properties. Despite the fact that CBF1 was identified a decade ago, not much is known on the mechanism whereby CBF proteins activate COR gene expression. The C-terrninal domain of CBF (CBF1116-213) functions as the activation domain in both yeast and Arabidopsis (Wang et al., 2005). Within this domain are several hydrophobic clusters; each of them provides some functional redundancy, presumably to maintain efficient activation of the CBF regulon under cold stress conditions (Wang et al. , 2005). More recently, investigators have explored the role of chromatin-modifying factors in facilitating CBF trans-activating properties. The CBF1 activation domain is acidic in nature, similar to that of other transcription factors such as VP16 and Gcn4 (Hope and Struhl, 1986; Seipel et al., 1994). These proteins stimulate transcription in part by recruitment of chromatin-modifying complexes such as Ada or SAGA (Kuo et al. , 1998; Ikeda et al. , 1999). The transcriptional adaptors ADA2 and ADA3 and the histone acetyltransferase GCNS are part of these large multi-protein complexes (Grant et al. , 1997). They were shown to be essential for CBF1 transcriptional activity in yeast 15 (Stockinger et al. , 2001). More recently, it was discovered that CBF1 can also interact in vitro with AtGCN5 and AtADA2b, the Arabidopsis homologues of GCNS and ADA2 from yeast (Mao et al. , 2006). Contrary to what expected, these interaction occur through the DNA-binding domain of CBF1, and not through the activation domain. Furthermore, they are not specific to CBF1 but appear to be conserved within the AP2/EREBP family, as a similar interaction occurs with another member of the family, TINY (Wilson et al. , 1996; Mac et al., 2006). To better understand the role of AtGCN5 and AtADA2b in vivo, T-DNA insertion mutations for the two genes were analyzed. Disruption of AtGCN5 and AtADA2b affects cold-regulated induction of the CBF regulon in Arabidopsis, by reducing the level of COR gene'expression and delaying the kinetics of induction (V lachonasios et al. , 2003). Taken together, the emerging evidence is that these two adaptor proteins might be recruited by CBF1 to mediate COR gene induction. In the effort to identify components of the cold acclimation response in Arabidopsis, a set of mutants sensitive to freezing was isolated (McKown et al. , 1996). In one of those mutants, sfi6, cold-induced activation of CRT/DRE-containing genes was affected in one of the steps that follow CBF1 induction (Knight et al., 1999). CBF1 transcript still accumulated in response to low temperatures, yet COR gene expression was compromised (Knight et a1. , 1999; Boyce et al., 2003). The nature of the sfi'6 mutation is still unknown, and thus the contribution of this factor to the CBF pathway has not been elucidated. Given that the CBF transcript is unaffected in sfi6 plants, the role of the SFR6 protein is post-transcriptional. Among other explanations, it is reasonable that SFR6 might act as a co-factor of CBF. 16 Additional post-transcriptional events regulating CBF activity. One way to modulate the activity of transcription factors is by actively controlling their cellular localization (Tran and Wente, 2006). Whether nuclear shuttling of CBF1 is dependent on a nuclear localization signal (N LS), and whether this is a low temperature-regulated process is not clear. It is expected that some CBF1 protein can accumulate in the nucleus without low temperature treatment, since Arabidopsis lines overexpressing CBF1, CBF 2 or CBF 3 can accumulate high levels of COR gene transcripts compared to the non transgenic plants in the absence of a cold stimulus (J aglo- Ottosen et al. , 1998; Gilmour et al. , 2000). On the other hand, when CBF-overexpressing plants are cold treated, COR gene expression increases (Gilmour et al., 2000). Among other reasons, this may be due to more efficient nuclear transport. Recent studies focused on the subcellular localization of several members of the AP2/EREBP family have revealed that nuclear targeting within this family can be either constitutive or stimulus- dependent. For instance, localization studies on AINTEGUMENTA (Klucher et al. , 1996) have indicated that nuclear targeting of this AP2/EREBP protein is constitutive and depends on the presence of specific basic residues (Krizek and Sulli, 2006). On the contrary, the cytokinin response factors CRFs accumulate in the nucleus in response to cytokinin treatment (Rashotte et al. , 2006). Moreover, a high-throughput localization study of Arabidopsis proteins has identified an AP2/EREBP protein (id. At2g22880) that can be found both in the nucleus and in the cytoplasm and four that are constitutively nuclear localized (id. At1g13260, At2g39250, At2g33710) under standard grth conditions (Koroleva et al. , 2005). Altogether these studies indicate that nuclear localization of proteins in the AP2/EREBP family is regulated differently, and a case by 17 case study is required to understand the mechanisms dictating nuclear import of each protein. Based on its similarity to known nuclear localization motifs found in other plant and non-plant transcription factors, the PKKPAGR motif has been proposed to be the nuclear targeting signal of CBF1 (Stockinger et al. , 1997). Whether this motif is required and/or sufficient for nuclear import of CBF proteins remains to be determined. Conclusion Since the discovery of CBF 1 in Arabidopsis, several lines of evidence have suggested that additional cold response pathways are present. While it is important to define these alternative pathways, furthering our knowledge of the CBF action mechanisms could better our understanding of how these novel pathways are activated and promote freezing tolerance in plants. The relevance of the CBF pathway has been emphasized over the recent years by large-scale analyses, such as transcriptome and metabolome profiling, and by numerous reports showing that this pathway is conserved across diverse plants species. Understanding how CBF proteins are activated and regulate the changes that lead to freezing tolerance will help define some of the basic mechanisms that have been conserved not only the CBF pathway but also in novel cold-responsive pathways, and it will ultimately help establish how plants sense and respond to low temperatures. To contribute to this effort, in dissertation I will discuss the functional characterization of the Arabidopsis CBF1 protein, by describing the functional properties of two conserved domains in the CBF family of transcriptional activators. 18 LITERATURE CITED Acevedo-Hernandez, G.J., Leon, P., and Herrera-Estrella, LR. (2005). Sugar and ABA responsiveness of a minimal RBCS light-responsive unit is mediated by direct binding of ABI4. The Plant Journal 43, 506-519. Agarwal, M., Hao, Y., Kapoor, A., Dong, C.-H., Fujii, H., Zheng, X., and Zhu, J.-K. (2006). A R2R3 type MYB transcription factor is involved in the cold regulation of CBF genes and in acquired freezing tolerance. J. Biol. Chem. 281, 37636- 37645. Allen, M., Yamasaki, K., Ohme-Takagi, M., Tateno, M., and Suzuki, M. (1998). A novel mode of DNA recognition by a beta-sheet revealed by the solution structure of the GCC-box binding domain in complex with DNA. EMBO J 17, 5484-5496. Baker, 8., Wilhelm, K., and Thomashow, M. (1994). The 5'-region of Arabidopsis thaliana corl 5a has cis-acting elements that confer cold-, drought- and ABA- regulated gene expression. Plant Mol Biol 24, 701-713. Balaji, S., Babu, M., Iyer, L., and Aravind, L. (2005). Discovery of the principal specific transcription factors of Apicomplexa and their implication for the evolution of the AP2-integrase DNA binding domains. Nucl. Acids Res. 33, 3994-4006. Benedict, C., Skinner, J.S., Meng, R., Chang, Y., Bhalerao, R., Huner, N.P.A., Finn, C.E., Chen, T.H.H., and Hurry, V. (2006). The CBF1-dependent low temperature signalling pathway, regulon and increase in freeze tolerance are conserved in Populus spp. Plant, Cell & Environment 29, 1259-1272. Boyce, J., Knight, H., Deyholos, M., Openshaw, M., Galbraith, D., Warren, G., and Knight, M. (2003). The sfr6 mutant ofArabidopsis is defective in transcriptional activation via CBF/DREBI and DREB2 and shows sensitivity to osmotic stress. The Plant Journal 34, 395-406. Carpenter, J.F., and Crowe, J.H. (1988). The mechanism of cryoprotection of proteins by solutes. Cryobiology 25, 244. Chinnusamy, V., Ohta, M., Kanrar, S., Lee, B.-h., Hong, X., Agarwal, M., and Zhu, J.-K. (2003). ICE]: a regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev. 17, 1043-1054. Choi, D., Zhu, B., and Close, T. (1999). The barley (Hordeum vulgare L.) dehydrin multigene family: sequences, allele types, chromosome assignments, and expression characteristics of 11 Dhn genes of cv Dicktoo. Theor. Appl. Genet. 98, 1234—1247. 19 Cook, D., Fowler, 8., Fiehn, 0., and Thomashow, M.F. (2004). A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. PNAS 101, 15243-15248. Dal Bosco, C., Busconi, M., Govoni, C., Baldi, P., Stanca, A.M., Crosatti, C., Bassi, R., and Cattivelli, L. (2003). car gene expression in barley mutants affected in chloroplast development and photosynthetic electron transport. Plant Physiol 131, 793—802. Dubouzet, J.G., Sakuma, Y., Ito, Y., Kasuga, M., Dubouzet, E.G., Miura, S., Seki, M., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2003). OsDREB genes in rice, Oryza sativa L., encode transcription activators that function in drought-, high-salt- and cold-responsive gene expression. The Plant Journal 33, 751-763. Fowler, 8., and Thomashow, M.F. (2002). Arabidopsis transcriptome profiling indicates that multiple regulatory pathways are activated during cold acclimation in addition to the CBF cold response pathway. Plant Cell 14, 1675-1690. Gilmour, S.J., Hajela, RK, and Thomashow, M.F. (1988). Cold Acclimation in Arabidopsis thaliana. Plant Physiol. 87, 745-750. Gilmour, S.J., Fowler, S.G., and Thomashow, M.F. (2004). Arabidopsis transcriptional activators CBF1, CBF 2, and CBF3 have matching functional activities. Plant Mol Biol 54, 768-781. Gilmour, S.J., Sebolt, A.M., Salazar, M.P., Everard, J.D., and Thomashow, M.F. (2000). Overexpression of the Arabidopsis CBF3 transcriptional activator mimics multiple biochemical changes associated with cold acclimation. Plant Physiol. 124, 1854-1865. Gilmour, S.J., Zarka, D.G., Stockinger, E.J., Salazar, M.P., Houghton, J.M., and Thomashow, M.F. (1998). Low temperature regulation of the Arabidopsis CBF family of AP2 transcriptional activators as an early step in cold-induced COR gene expression. The Plant Journal 16, 433-442. Grant, P.A., Duggan, L., Cote, J., Roberts, S.M., Brownell, J.E., Candau, R., Ohba, R., Owen-Hughes, T., Allis, C.D., Winston, F., Berger, S.L., and Workman, J .L. (1997). Yeast Gcn5 functions in two multisubunit complexes to acetylate nucleosomal histones: characterization of an Ada complex and the SAGA (Spt/Ada) complex. Genes Dev. 1 1, 1640-1650. Guy, C.L., and Haskell, D. (1987). Induction of freezing tolerance in spinach is associated with the synthesis of cold acclimation induced proteins. Plant Physiol. 84, 872-878. Guy, C.L., Niemi, K.J., and Brambl, R. (1985). Altered gene expression during cold acclimation of spinach. PNAS 82, 3673-3677. 20 Haake, V., Cook, D., Riechmann, J.L., Pineda, 0., Thomashow, M.F., and Zhang, J.Z. (2002). Transcription factor CBF 4 is a regulator of drought adaptation in Arabidopsis. Plant Physiol. 130, 639-648. Hajela, R.K., Horvath, D.P., Gilmour, S.J., and Thomashow, M.F. (1990). Molecular cloning and expression of cor (Cold-Regulated) genes in Arabidopsis thaliana. Plant Physiol. 93, 1246-1252. Hao, D., Ohme-Takagi, M., and Sarai, A. (1998). Unique mode of GCC box recognition by the DNA-binding domain of ethylene-responsive element-binding factor (ERF domain) in plant. J Biol Chem 273, 26857-26861. Hao, D., Yamasaki, K., Sarai, A., and Ohme—Takagi, M. (2002). Determinants in the sequence specific binding of two plant transcription factors, CBF1 and NtERF2, to the DRE and GCC motifs. Biochemistry 41, 4202-4208. Hart, C., Nagy, F., and Meins, F. (1993). A 61 bp enhancer element of the tobacco beta-1,3-glucanase B gene interacts with one or more regulated nuclear proteins. Plant Mol Biol 21, 121-131. Hope, I.A., and Struhl, K. (1986). Functional dissection of a eukaryotic transcriptional activator protein, GCN4 of Yeast. Cell 46, 885. Horvath, D.P., McLarney, B.K., and Thomashow, M.F. (1993). Regulation of Arabidopsis thaliana L. (Heyn) cor78 in response to low temperature. Plant Physiol. 103, 1047-1053. Hsieh, T.-H., Lee, J.-T., Yang, P.-T., Chiu, L.-H., Charng, Y.-y., Wang, Y.-C., and Chan, M.-T. (2002). Heterology expression of the Arabidopsis C- Repeat/Dehydration Response Element Binding Factor 1 gene confers elevated tolerance to chilling and oxidative stresses in transgenic tomato. Plant Physiol. 129, 1086- 1 094. Ikeda, K., Steger, D.J., Eberharter, A., and Workman, J.L. (1999). Activationd domain-specific and general transcription stimulation by native histone acetyltransferase complexes. Mol. Cell. Biol. 19, 855-863. Jaglo-Ottosen, K.R., Gilmour, S.J., Zarka, D.G., Schabenberger, 0., and Thomashow, M.F. (1998). Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 280, 104-106. Jaglo, K.R., Kleff, S., Amundsen, K.L., Zhang, X., Haake, V., Zhang, J.Z., Deits, T., and Thomashow, M.F. (2001). Components of the Arabidopsis C- Repeat/Dehydration-Responsive Element Binding Factor cold-response pathway are conserved in Brassica napus and other plant species. Plant Physiol. 127, 910- 91 7. 21 Johnson-Flanagan, A.M., and Singh, J. (1987). Alteration of gene expression during the induction of freezing tolerance in Brassica napus suspension cultures. Plant Physiol. 85, 699-705. Klucher, K.M., Chow, H., Reiser, L., and Fischer, RL. (1996). The AINTEGUMENTA gene of Arabidopsis required for ovule and female gametophyte development is related to the floral homeotic gene APETALAZ. Plant Cell 8, 137-153. Knight, H., Veale, E.L., Warren, G.J., and Knight, MR. (1999). The sfi'6 mutation in Arabidopsis suppresses low-temperature induction of genes dependent on the CRT/DRE sequence motif. Plant Cell 11, 875-886. Koroleva, 0., Tomlinson, M., Leader, D., Shaw, P., and Doonan, J. (2005). High- throughput protein localization in Arabidopsis using Agrobacterium-mediated transient expression of GFP-ORF fusions. Plant J 41, 162-174. Kreps, J.A., Wu, Y., Chang, H.-S., Zhu, T., Wang, X., and Harper, J.F. (2002). Transcriptome changes for Arabidopsis in response to salt, osmotic, and cold stress. Plant Physiol. 130, 2129-2141. Krizek, B., and Sulli, C. (2006). Mapping sequences required for nuclear localization and the transcriptional activation function of the Arabidopsis protein AINTEGUMENTA. Planta 224, 612-621. Kuo, M.-H., Zhou, J., Jambeck, P., Churchill, M.E.A., and Allis, CD. (1998). Histone acetyltransferase activity of yeast Gcn5p is required for the activation of target genes in vivo. Genes Dev. 12, 627-639. Kurkela, S., Franck, M., Heino, P., Lang, V., and Palva, E. (1988). Cold induced gene expresssion in Arabidopsis thaliana L. Plant Cell Reports 7, 495-498. Levitt, J. (1980). Responses of plants to environmental stresses. (New York: Academic Press). Liu, Q., Kasuga, M., Sakuma, Y., Abe, H., Miura, S., Yamaguchi-Shinozaki, K., and Shinozaki, K. (1998). Two transcription factors, DREBI and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10, 1391-1406. Magnani, E., Sjolander, K., and Hake, S. (2004). From endonucleases to transcription factors: evolution of the AP2 DNA binding domain in plants. Plant Cell 16, 2265- 2277. Magome, H., Yamaguchi, S., Hanada, A., Kamiya, Y., and Oda, K. (2004). dwarf and delayed-flowering 1, a novel Arabidopsis mutant deficient in gibberellin 22 biosynthesis because of overexpression of a putative AP2 transcription factor. The Plant Journal 37, 720-729. Mao, Y., Pavangadkar, K., Thomashow, M., and Triezenberg, S. (2006). Physical and functional interactions of Arabidopsis ADA2 transcriptional coactivator proteins with the acetyltransferase GCNS and with the cold-induced transcription factor CBF1. Biochirnica et Biophysica Acta 1759, 69-79. Maruyama, K., Sakuma, Y., Kasuga, M., Ito, Y., Seki, M., Goda, H., Shimada, Y., Yoshida, S., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2004). Identification of cold-inducible downstream genes of the Arabidopsis DREBlA/CBF3 transcriptional factor using two microarray systems. The Plant Journal 38, 982-993. McKown, R., Kuroki, G., and Warren, G. (1996). Cold response ofArabidopsis mutants impaired in freezing tolerance. Journal of Experimental Botany 47, 1919- 1925. Mohapatra, S.S., Poole, R.J., and Dhindsa, RS. (1987). Changes in protein patterns and translatable messenger RNA populations during cold acclimation of alfalfa. Plant Physiol. 84, 1 172-1 176. Ohme-Takagi, M., and Shinshi, H. (1995). Ethylene-inducible DNA binding proteins that interact with an ethylene-responsive element. Plant Cell 7, 173-182. Ohta, M., Matsui, K., Hiratsu, K., Shinshi, H., and Ohme—Takagi, M. (2001). Repression domains of Class II ERF transcriptional repressors share an essential motif for active repression. Plant Cell 13, 1959-1968. Qin, F., Sakuma, Y., Li, J., Liu, Q., Li, Y.-Q., Shinozaki, K., and Yamaguchi- Shinozaki, K. (2004). Cloning and functional analysis of a novel DREBI/CBF transcription factor involved in cold-responsive gene expression in Zea mays L. Plant Cell Physiol. 45, 1042-1052. Rashotte, A.M., Mason, M.G., Hutchison, C.E., Ferreira, F.J., Schaller, G.E., and Kieber, J.J. (2006). A subset of Arabidopsis AP2 transcription factors mediates cytokinin responses in concert with a two-component pathway. PNAS 103, 11081-11085. Riechmann, J.L., and Meyerowitz, EM. (1998). The AP2/EREBP family of plant transcription factors. Biol Chem 379, 633-646. Rudolph, A.S., and Crowe, J.H. (1985). Membrane stabilization during freezing: The role of two natural cryoprotectants, trehalose and proline. Cryobiology 22, 367. Sakai, A., and Larcher, W. (1987). Frost survival of plants: responses and adaptation to freezing stress. (Berlin: Springer-Verlag). 23 Sakuma, Y., Liu, Q., Dubouzet, J.G., Abe, H., Shinozaki, K., and Yamaguchi- Shinozaki, K. (2002). DNA-binding specificity of the ERF/AP2 domain of Arabidopsis DREBs, transcription factors involved in dehydration- and cold- inducible gene expression. Biochem Biophys Res Commun 290, 998-1009. Sato, F., Kitajima, S., and Koyama, T. (1996). Ethylene-induced gene expression of osmotin-like protein, a neutral isoform of tobacco PR-5, is mediated by the AGCCGCC cit-sequence. Plant Cell Physiol. 37, 249-255. Seipel, K., Georgiev, 0., and Schaffner, W. (1994). A minimal transcription activation domain consisting of a specific array of aspartic acid and leucine residues. Biol Chem Hoppe Seyler 375, 463-470. Seki, M., Narusaka, M., Abe, H., Kasuga, M., Yamaguchi-Shinozaki, K., Carninci, P., Hayashizaki, Y., and Shinozaki, K. (2001). Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a Full-Length cDNA microarray. Plant Cell 13, 61-72. Seki, M., Narusaka, M., Ishida, J., Nanjo, T., Fujita, M., Oono, Y., Kamiya, A., Nakajima, M., Enju, A., Sakurai, T., Satou, M., Akiyama, K., Taji, T., Yamaguchi-Shinozaki, K., Carninci, P., Kawai, J., Hayashizaki, Y., and Shinozaki, K. (2002). Monitoring the expression profiles of 7000 Arabidopsis genes under drought, cold and high-salinity stresses using a full-length cDNA microarray. The Plant Journal 31, 279-292. Skinner, J., van Zitzewitz, J., Szucs, P., Marquez-Cedillo, L., Filichkin, T., Amundsen, K., Stockinger, E., Thomashow, M., Chen, T., and PM., H. (2005). Structural, functional, and phylogenetic characterization of a large CBF gene family in barley. Plant Mol Biol 59, 533-551. Steponkus, P., and Webb, M. (1992). Freeze-induced dehydration and membrane destabilization in plants. (Berlin: Springer-Verlag). Steponkus, P.L., Uemura, M., Joseph, R.A., Gilmour, S.J., and Thomashow, M.F. (1998). Mode of action of the COR15a gene on the freezing tolerance of Arabidopsis thaliana. PNAS 95, 14570-14575. Stockinger, E.J., Gilmour, S.J., and Thomashow, M.F. (1997). Arabidopsis thaliana CBF1 encodes an AP2 domain-containing transcriptional activator that binds to the C-repeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in response to low temperature and water deficit. PNAS 94, 103 5- 1040. Stockinger, E.J., Mao, Y., Regier, M.K., Triezenberg, S.J., and Thomashow, M.F. (2001). Transcriptional adaptor and histone acetyltransferase proteins in Arabidopsis and their interactions with CBF1 , a transcriptional activator involved in cold-regulated gene expression. Nucl. Acids Res. 29, 1524-1533. 24 Strauss, G., and Hauser, H. (1986). Stabilization of lipid bilayer vesicles by sucrose during freezing. PNAS 83, 2422-2426. Thomashow, M.F. (1998). Role of cold-responsive genes in plant freezing tolerance. Plant Physiol. 118, 1-8. Thomashow, M.F. (1999). Plant cold acclimation: fi'eezing tolerance genes and regulatory mechanisms. Annual Review of Plant Physiology and Plant Molecular Biology 50, 571-599. Tran, E.J., and Wente, S.R. (2006). Dynamic Nuclear Pore Complexes: Life on the Edge. Cell 125, 1041. Triezenberg, SJ. (1995). Structure and function of transcriptional activation domains. Current Opinion in Genetics & Development 5, 190. van der Fits, L., and Memelink, J. (2001). The jasmonate-inducible AP2/ERF-domain transcription factor ORCA3 activates gene expression via interaction with a jasmonate-responsive promoter element. The Plant Journal 25, 43-53. Vlachonasios, K.E., Thomashow, M.F., and Triezenberg, SJ. (2003). Disruption mutations of ADA2b and GCN5 transcriptional adaptor genes dramatically affect Arabidopsis growth, development, and gene expression. Plant Cell 15, 626-638. Vogel, J.T., Zarka, D.G., Van Buskirk, H.A., Fowler, S.G., and Thomashow, M.F. (2005). Roles of the CBF2 and ZAT12 transcription factors in configuring the low temperature transcriptome of Arabidopsis. The Plant Journal 41, 195-211. Volger, H., and Heber, U. (1975). Cryoprotective leaf proteins. Biochim Biophys Acta 412, 335-349. Wang, Z., Triezenberg, S., Thomashow, M., and EJ., S. (2005). Multiple hydrophobic motifs in Arabidopsis CBF1 COOH-terminus provide functional redundancy in trans-activation. Plant Mol Biol 58, 543-559. Wilson, K., Long, D., Swinburne, J., and Coupland, G. (1996). A dissociation insertion causes a semidominant mutation that increases expression of TINY, an Arabidopsis gene related to APETALA2. Plant Cell 8, 659-671. Wuitschick, J.D., Lindstrom, P.R., Meyer, A.E., and Karrer, KM. (2004). Homing endonucleases encoded by germ line-limited genes in T etrahymena thermophila have APETALA2 DNA binding domains. Eukaryotic Cell 3, 685-694. Xin, Z., and Browse, J. (1998). eskimoI mutants ofArabidopsis are constitutively freezing-tolerant. PNAS 95, 7799-7 804. 25 CHAPTER TWO FUNCTIONAL ROLE OF THE CONSERVED PKKPAGR AND DSAWR MOTIFS IN CBF1 ACTIVITY INTRODUCTION CBF proteins are members of the AP2/EREBP family of transcription factors in Arabidopsis (Riechmann and Meyerowitz, 1998). This multi-gene family is highly conserved among plants and is characterized by the presence of the AP2/EREBP DNA- binding domain (Jofuku et al. , 1994; Ohme-Takagi and Shinshi, 1995). NMR studies on AtERFl (Allen et al. , 1998) revealed that this DNA-binding domain is composed of a three-strand B-sheet structure followed by an or-helix. The B-sheet, at the N-terrninus, is the fold which mediates DNA recognition. High identity within the DNA-binding domain and conservation of the critical amino acids required for protein-DNA interaction suggests that proteins in the AP2/EREBP family likely adopt a similar 3D structure. Within the AP2/EREBP family in Arabidopsis, CBF1/DREB1B, CBF2/DREB1C, CBF3/DREB1A, CBF4/DREB1D, DREBlE/DDFI and DREBlF/DDF2 (Stockinger et al., 1997; Gilmour et al. , 1998; Liu et al., 1998; Haake et al. , 2002; Sakuma et al. , 2002; Magome et al. , 2004) define a small sub-family that is characterized by the presence of two short polypeptide sequences that flank the AP2/EREBP domain: PKKP/RAGRxKFxETRHP (abbreviated as PKKPAGR) and DSAWR, at the N-and C- terrninus, respectively (Figure 2.1), designated the “signature” sequences (Jaglo et al., 2001). 27 1 32 47 105110 213 Figure 2.1. PKKPAGR and DSAWR signature sequences in the CBF family of plant transcription factors. Top panel. Schematic diagram of CBF1 protein. Yellow boxes represent PKKPAGR and DSAWR motifs. The grey box at the C-terminus represents the activation domain. The AP2/EREBP DNA-binding domain is indicated Bottom panel Sequence alignment of AtCBF1-3 and CBF-like proteins. The signature sequences are included in the yellow boxes. The consensus sequence is indicated below. At, Arabidopsis thaliana; B. napus, Brassica napus; N. tabacum, Nicotiana tabacum; S. cereals, Secale cereale; T. aestivum, Tritium aestivran; G. max, Glycine max; H. vulgare, Horderan vulgae. 28 Orthologues of the CBF genes are present in a wide variety of plant species. A sequence alignment of CBF representatives from evolutionarily distant plant species showed that most of the identity among the Arabidopsis CBF family and CBF-like proteins from other species is due to a high degree of identity between their AP2/EREBP domains and the signature sequences (J aglo et al., 2001). Conservation of these motifs across evolutionarily diverse plants species suggests that they have an important functional role in CBF activity. When AtCBFl was first isolated, the PKKPAGR motif was proposed to represent the signal responsible for nuclear localization of the protein (Stockinger et al. , 1997). This observation was based on the similarity of this sequence to known nuclear localization signals from other plant proteins, such as the early induced Aux/1AA genes (Abel et al. , 1994), the photomorphogenic repressor protein COPl (von Arnim and Deng, 1994) and the heat shock factors, HSF (Lyck et al., 1997). In addition, localization studies of the member of the AP2/EREBP family AINTEGUMENTA (Klucher et a1. , 1996) have indicated that nuclear targeting can depend on the presence of specific basic residues (Krizek and Sulli, 2006). Whether nuclear shuttling of CBF1 is dependent on the PKKPAGR motif, and whether this is a low temperature-regulated process is not clear. Recent studies focused on the subcellular localization of several members of the AP2/EREBP family have revealed that nuclear targeting within this family can be either constitutive or stirnulus- dependent (Klucher et al. , 1996; Rashotte et al. , 2006); therefore it is not possible to draw a general conclusion based on studies on other AP2/EREBP family members. It is expected that some CBF protein can accumulate in the nucleus in the absence of a cold 29 stimulus, as Arabidopsis lines overexpressing CBF1, CBF 2 or CBF 3 display constitutive accumulation of COR transcripts compared to the non-transgenic plants (J aglo-Ottosen et al., 1998; Gilmour et al., 2000). On the other hand, when CBF-overexpressing plants are cold-treated, COR gene expression increases (Gilmour et al. , 2000), suggesting that nuclear transport may be more efficient at low temperatures. For several families of DNA-binding proteins, regions flanking the DNA-binding domain can be an essential extension important not only for binding affinity but for binding specificity as Well (Crane-Robinson et al., 2006). Based on their proximity to the AP2/EREBP domain, a reasonable hypothesis is that the PKKPAGR and DSAWR motifs participate in binding to the CR T /DRE promoter element. In addition, the N-terrninal PKKPAGR region of this motif is rich in basic residues, bearing positive charges that may favor a DNA-protein interaction. Finally, deletion studies of ERFl have shown that the presence of 10 amino acids immediately upstream and 8 amino acids downstream of the AP2/EREBP domain can greatly stabilize DNA binding (Hao et al., 1998). Based on these observations, we hypothesized that the two signature motifs might play a role in CBF binding to its cognate CR T/DRE promoter element. The main goal of the experiments described in this chapter was to investigate the roles of the PKKPAGR and DSAWR motifs in CBF function. The question was addressed by taking a mutational approach. Arabidopsis transgenic lines were generated by overexpressing CBF1 transgenes harboring specific mutations in the PKKPAGR and DSAWR motifs. As a hallmark of CBF1 activity, COR gene expression was tested. The results demonstrate that both conserved motifs are required for CBF1 function, since Arabidopsis plants overexpressing a CBF1 transgene carrying mutations in the signature 30 sequences show greatly diminished COR gene expression compared to plants overexpressing the wild type CBF1 transgene. The effect of the mutations cannot be explained by reduced protein levels, as indicated by western blot analysis of plants overexpressing the 6xMyc: CBF1 transgenes. Contrary to what was originally hypothesized, the PKKPAGR motif alone is not required for nuclear localization of a CBF1 :GFPzGUS chimera. Electrophoretic-mobility gel shift experiments showed that the PKKPAGR and DSAWR motifs play a role in mediating the recognition of the CRT/DRE cis-acting element present in the COR gene promoters. Extensive mutational analysis within the PKKPAGR motif combined with secondary structure prediction methods suggested that certain residues in the predicted helical region represented by the RKKFRET sequence are involved in direct protein-DNA interaction. A computational model of CBF1 bound to its target CR T/DRE promoter element has led to the identification of specific residues that are essential to DNA binding and might mediate specific DNA recognition. 31 RESULTS A MUTATIONAL APPROACH TO INVESTIGATE THE FUNCTIONAL IMPORTANCE OF THE SIGNATURE SEQUENCES IN CBF1 FUNCTION. To investigate the importance of the signature sequences in CBF function, Arabidopsis transgenic lines were generated by overexpressing CBF1 transgenes harboring specific mutations in the PKKPAGR and DSAWR motifs. AtCBFI was selected as a representative model for the Arabidopsis CBF family, as it has been shown that overexpression ofAtCBFI, AtCBF 2, or AtCBF 3 results in activation of the same CBF regulon of genes, indicating that the three proteins have overlapping function (Gilmour et al. , 2004). Alanine scanning was used to introduce stretches of three alanines in place of the original amino acids within the PKKPAGR motif, which resulted in 5 mutant ORFs named M1-M5 (Figure 2.2). In addition, the entire PKKPAGR motif was deleted, and the resulting ORF named APKK. The wild type and mutagenized versions of CBF1 driven by the constitutive 35$ Cauliflower Mosaic Virus (3 SS CaMV) promoter were transformed into Arabidopsis and are referred hereafter as pkkpagr (APKK, M1-M5) and dsawr lines. For each construct, fifteen to twenty independent homozygous lines were selected, and five to seven representative lines were chosen for further characterization. 32 M51 142 M3 M4 105 no 213 II I II I WT DSAWR (is art: AAAAA Figure 2.2. Site-directed mutagenesis was used to generate CBF1 ORFs containing specific mutations or deletions of the signature sequences. Top. Schematic of full-length CBF1 and CBF1 lacking the PKKPAGR motif (APKK). Bottom. Description of specific mutations within the signature sequences. WT: wild type sequence for the two signature sequences. M1-M5, dsawr: Arrrino acids in red indicate where the wild type CBF1 sequence has been substituted with alanines. 33 Overexpression of AtCBF I , AtCBF 2, or AtCBF 3 leads to constitutive expression of the CBF regulon, a group of about 85 genes containing the cis-acting CRT/DRE promoter element (V ogel et al., 2005). As a result, CBF-overexpressing plants are constitutively freezing-tolerant. However, overexpression of wild type CBF1 results in transgenic plants that are smaller in stature and delayed in flowering compared to the wild type plants or transgenic plants harboring the empty vector. In addition, higher amounts of CBF transcripts correlate with smaller plants that are delayed in flowering (Haake et al., 2002; Gilmour SJ et al., 2004; Gilmour et al., 2004). Transgenic lines overexpressing CBF1-pkkpagr at levels similar to or higher than plants overexpressing wild type CBF1 were selected by Northern blot analysis (Figure 2.3 C) and their growth phenotype was compared (representative examples are shown in Figure 2.3 A). 34 \ Figure 2.3. Arabidopsis pkkpagr and dsawr lines. Seedlings harboring the 35S::CBFI construct (026 and G39), CBF1 mutants (M1 —5, APKK, and dsawr) or the empty vector (B6) were grown on plates for 14 days before being transferred to soil. Pictures were taken 30 days after transplanting into soil A pkkpagr lines. B. Left, dsawr transgenic line. C. Northern blot analysis of different 358::CBF1 Arabidopsis plants. Total RNA extracted from two-week old seedlings was tested for the expression of CBF1. RNA levels of the 188 ribosomal RNA were used as a normalization control. 35 Overall, overexpression of CBF1 carrying alanine mutations in the PKKPAGR motif resulted in plants showing different degrees of growth retardation, as compared to the control line B6, carrying the empty vector. However, the growth retardation observed in the pklgoagr lines was always less severe than that observed in the WT CBF1 over- expressing plants, when similar transcript levels were compared (compare plant growth and CBF1 expression levels in Figures 2.3A and C). These observations suggested that the PKKPAGR motif is required for CBF1 fimction. As was observed in the pkkpagr lines, the negative effects of CBF1-dsawr overexpression on plant growth were much milder than those in the wild type overexpressor (Figure 2.3B). It must be noted that CBF1 transcript levels in the dsawr plants were similar to or higher than in the plants overexpressing wild type CBF1 (representative example shown in Figure 2.3 C), suggesting that the DSAWR motif also plays a role in CBF1 function. In summary, overexpression of CBF1 transgenes mutated in the two signature sequences altered of one of the hallmark phenotypes of CBF-overexpressing plants, growth retardation, providing a first suggestion that these two conserved motifs play a role in CBF1 activity. 36 FUNCTIONAL ROLE OF THE PKKPAGR MOTIF IN CBF1 ACTIVITY. Mutations in the PKKPAGR motif affect COR gene activation. To assess the effect of alanine mutations on CBF1 activity, wild type and pkkpagr lines were analyzed for COR gene expression by northern blot analysis. Overexpression of AtCBF 1 leads to constitutive expression of the CBF regulon, including COR6. 6, COMM, COR47, and COR 78 (Jaglo-Ottosen et al., 1998). If the PKKPAGR motif is required for CBF1 activity, we would expect that the ability of CBF1 to activate these target genes would be impaired or absent in transgenic plants expressing CBF proteins with mutations in this motif. This possibility was tested by determining expression levels of target COR genes in transgenic plants overexpressing wild type and mutant versions of the PKKPAGR sequence. Total RNA was analyzed for the expression of CBF] and COR genes in independent pkkpagr lines, as well as in transgenic lines overexpressing the wild type CBF1 or harboring the empty vector. The experiment was carried out by growing Arabidopsis seedlings at control temperature (22°C). At this temperature the endogenous CBF and COR gene transcripts are barely detectable by northern blotting (Figure 2.4A and B, see B6 line, which is transformed with the empty vector). In contrast, transgenic lines overexpressing wild type CBF1 are able to activate constitutive expression of the COR genes (Jaglo-Ottosen et al., 1998; Figure 2.4A and B, see all WT lines). 37 WT APKK M1 M2 v-I‘O u-I BEss-Nnean-s_s—nsza COR6.6 -.. ."fltuo-"N-“fl COR” 040-" """‘"‘“-' -oo.q~-.l COR15a fl .- rss ‘i’W-MQ Figure 2.4 A. Analysis of COR gene induction in APKK, M1 and M2 pklrpagr lines by Northern blot analysis. Total RNA was extracted from 2-week Old seedlings and tested for the expression of CBF1, COR6. 6, COR 78, and COR15a. 18S ribosomal RNA levels are shown as loading control. B6, transgenic plants carrying an empty expression vector; WT, wild type CBF1 overexpressing plants; APKK, M1 and M2, transgenic lines overexpressing CBF1 lacking the entire PKKPAGR region or carrying triple alanine mutations, respectively. Numbers above lanes represent independent transgenic lines. 38 1 5 B6 G5 G6 G2 G26 G39 CBF1 COR6.6 - - n.- 188 uni... n COR78 nu... COR15a ”I! Q.“ 133 macaque. M3 M4 M5 N M O—i r-IMVIOb—r—iNINQ—I—IMBG—I—r .. we”- a. -1" Man. U "" .o .- "~i-i. .na---..§ " 7‘17 ---....- u‘-..‘- - H ’ I a it the .n afiwfiidgflmmflhm Figure 2.4 B. Analysis of COR gene induction in M3, M4 and M5 pkkpagr lines by Northern blot analysis. Total RNA was extracted from 2-week old seedlings and tested for the expression of CBF1, COR6. 6, COR 78, and COR15a. 18S ribosomal RNA levels are shown as a loading control. B6, transgenic plants canying an empty expression vector; WT, wild type CBF1 overexpressing plants; M3-M5, transgenic lines overexpressing CBF1 carrying triple alanine mutations. Numbers above lanes represent independent transgenic lines. 39 The pkkpagr lines showed higher COR gene induction than the control line B6, suggesting that the CBF1 mutated proteins had retained some ability to activate transcription (Figures 2.4 A and B, compare pkkpagr lines Ml-MS, and APKK with B6). However, this function was affected by alanine substitutions in the PKKPAGR motif. For instance, C 0R1 5a induction was greatly reduced in all the pkkpagr lines as compared to the transgenic plants overexpressing wild type CBF1 transgene (Figures 2.4 A and B, Ml-M5 lines), and almost undetectable in the lines overexpressing CBF1 transgene lacking the whole PKKPAGR motif (Figures 2.4 A, APKK lines). Induction of COR6.6 and COR78 appeared to be also affected, in that higher levels of mutated CBF1 transcript were needed to reach comparable COR transcript levels (Figures 2.4 A and B, compare COR6. 6 and COR 78 transcript levels and corresponding CBF1 transcript levels in WT and pkkpagr lines M1-M5). To determine whether mutations in the PKKPAGR motif had a significant effect on COR6.6 and COR 78 induction, we compared the COR/CBF1 transcript ratios in the transgenic lines overexpressing wild type and mutated CBF1 by ANOVA (Figures 2.5A and B). 40 150- 100- COR78/CBF1 50- 0.0 ‘ " ‘ WT APKK M1 M2 40004 3200— 24.00- 160)“ COR78/CBF1 800— o'm __7 7,, . .- 7;..11 'i’i. . L"»' 1‘ ,w. 'z, ("MW "1‘51"". WI" M3 M4 M5 Wl'versusAPlPro mutation was designed to test the effect of a helix-breaking residue on the stability of the protein-DNA complex. Based on these criteria, a reciprocal ArgHLys substitution and Phe—*Tyr were chosen as conservative mutations; Arg—>Ser, Lys—*Ala, and Phe—>Ala represented non-conservative substitutions. Preserving the positive charge, in the case of Arg and Lys, and the aromatic ring in the case of Phe—>Tyr, may result in a negligible effect on DNA binding; on the contrary, the non-conservative mutations would directly test the importance of the positive charge, in the case of Arg and Lys, or the contribution of the aromatic ring, in the case of Phe. Recombinant proteins carrying point mutations were tested by gel mobility shift assays in the presence of CRT/DRE promoter elements from COR15a, COR6. 6, and COR78 (Figure 2.13A, B, C). 68 Conservative Non-Conservative A121 LysZ Lys3 Phe4 Arg5 WT Lys Ser Arg Ala Arg Ala Ala Pro Tyr Ser WT _AAAAAAA —4l AAAé Figure 2.13 A. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR 78 gene promoter. Top panel. Substitution table indicating the point mutations designed within the RKKFRET region of CBF1 . Conservative and non-conservative mutations are indicated. H and C below the PKKPAGR motif indicate the predicted helical and coiled structures, respectively. Bottom panel. Gel mobility shift assays indicating the binding activity of different MBP: CBF127-112 proteins. DNA binding reactions were carried out in a 15 ul reaction containing increasing amounts (300 and 600 ng) of each recombinant protein in the presence of 0. 5 ng of radiolabeled CRT/DRE-containing probe fi'om COR 78 promoter. 69 Arg] LysZ Lys3 Phe4 Arg5 Phe4 MBP WT Lys Ala Arg Ala Ala Pro Ser Tyr AZEAAA _AAAA21 Figure 2.13 B. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR15a gene promoter. Gel mobility shift assays indicating the binding activity of different MBP:CBF127-112 proteins. DNA binding reactions were carried out in a 15 ul reaction containing increasing amounts (300 and 600 ng) of each recombinant protein in the presence of 0.5 ng of radiolabeled CRT/DRE—containing probe from COR15a promoter. 70 Argl Lysz Lys3 P_he4 A___rg_5 -2222222 _222 w Argl Lys2 Lys3 WT Lys Ser Arg Ala Arg Ala ._u. ’1 .fi'. n--uw§ tw". ”:‘3. " H4 MBP:CBF1 Figure 2.13 C. Effect of point mutations in the predicted RKKFRET helical region of CBF1 on binding activity of the protein to the COR6. 6 gene promoter. Top panel. Gel mobility shift assays indicating the binding activity of different MBP: CBF127- 112 proteins. DNA binding reactions were carried out in a 15 ul reaction containing increasing amounts (300 and 600 ng) of each recombinant protein in the presence of 0. 5 ng of radiolabeled CRT/DRE-containing probe from COR 78 promoter. Bottom panel. Western blot analysis (or-MBP antibody) showing protein loading. Twenty-five and 50 ng of each recombinant protein were analyzed. 71 The binding results observed at the three different promoters showed that most of the residues tested play an important role in DNA binding by MBP:CBF127412. The most striking observation was that the two mutations in Argl (Arg—*Lys and Arg—>Ser), and two of the three mutations on Phe4 (Phe—>Ala and Phe—>Pro) caused almost complete loss of DNA binding to the three promoters tested. The loss of binding by substitution of Argl with Lys, of similar side-chain length and charge, was significant and suggested that the interaction of this residue with DNA is possibly occurring through the DNA base. Loss of binding for the Phe4—>Ala substitution suggested that the aromatic ring in this residue is essential for the interaction with the DNA. Pro is a less conservative mutation that is likely to interrupt helicity, and therefore loss of DNA binding was expected. The conservative Phe—>Tyr mutation did not cause a significant change in binding activity; in fact, this mutated protein displayed an increase in DNA binding affinity compared to the wild type protein. This result suggests that the aromatic ring contributes favorable interactions with DNA bases; in addition, the terminal hydroxyl group, with its H- bonding donating accepting capacity, could add H-bonds between the side chain and either the DNA bases of with part of the backbone. Overall, substitutions at LysZ and LysB showed a moderate effect on DNA binding, and less effect when the mutant was electrostatically conservative (Lys—>Arg) than when the side chain was truncated (Lys—>A1a). These data suggested that the contribution of those residues to DNA binding is less important, and they could likely be involved in salt bridging to the phosphate backbone of the DNA. Finally, Arg5—>Ser substitution showed no significant effect on DNA binding, indicating that this residue is not important for this function. 72 Interestingly, we could observe some differences in the binding affinity of the mutant proteins to different promoters. For instance, Lys—>Arg mutation in LysZ affected binding to the CRT/DRE from the COR 78 promoter more strongly than the COR15a promoter (compare Figures 2.13A and B), and mutant CBF1 proteins carrying the Lys—+Ala mutation displayed different binding affinities to the three promoters. It must be noted, however, that binding to the COR 78 and the COR6. 6 promoters was tested only one time. Additional technical replicates will elucidate whether these results are reproducible. Taken altogether, these data indicate that specific amino acids within the PKKPAGR motif play a significant role in CBF1 binding to its target CR T/DRE promoter element. Based on the moderate loss of DNA binding for substitutions at LysZ and Ly53, it is likely that these residues are participating in electrostatic interactions at the protein- DNA interface through their positively charged side chains. The more dramatic effects observed for substitutions at Argl and Phe4 suggest that these amino acids might provide side chain interactions that are base-specific. Side chain conservation in AP2/EREBP-like proteins that use their flanking sequences to stabilize DNA binding. An important role of amino acid sequences flanking DNA-binding domains in DNA recognition has been previously reported for a few families of proteins, such as the homeodomain and high-mobility group families (Lnenicek-Allen et al., 1996; Wolberger, 1996; Dragan et al., 2004). The importance of specific amino acids in the PKKPAGR motif in DNA binding, made us wonder whether this phenomenon is conserved in DNA- 73 binding proteins that, like the AP2/EREBP proteins, use their B-sheets to interact with their target DNA. We addressed this question by searching the Protein Data Bank for structures similar to the three [i-stranded domain described for AtERFl (Allen et al. , 1998). Any consensus in this region would further support the idea that these residues sit close to the DNA, and participate directly to the binding activity of the protein, possibly by a common mechanism. The atomic coordinates for the NMR structure of the AP2/EREBP domain of AtERFl [PDB id. lgcc (Allen et al., 1998)] were used as a query to search the comprehensive Protein Data Bank (http://rcsb.org), which includes 3D structures obtained from crystallography and NMR studies. The method of choice was Dali (Holm and Sander, 1993), a well-established website for protein structure comparison. The server returned a list of 26 hits that shared significant structural similarity (Z-score >2.0) with l gcc, including ERF] itself (Table 2.3). Six of the other 25 structures did not contain any structural information on the sequences flanking the B-sheet at its N- terminus, and were discarded (data not shown). Of the remaining hits, the analysis was focused on nucleic acid-binding proteins, represented by ten structures. 74 Table 2.3. Summary of Dali analysis showing significant structural matches of nucleic acid binding proteins with the AP2/EREBP DNA binding domain of AtERFl. PDB Z- % Nucleic acid-binding proteins Nucleic Acid Flanking id. score, Idec containing a fl-sheet bound sequence, Manure r 2gcc 11.8 100 AtERFI fragment 7 N/A 2bn8 4.2 16 Cell division activator (E. colt) - 9L lkjk 3.9 10 Viral protein — intggrase \/ 7L lzlb 3.8 9 DNA integrase mutant ‘J 2L 2bb8 3.4 10 Tn9l6 integrase 7 9L 2d35 2.8 13 Cell division activator (E. colz') \/ 7L lrth 2.8 15 HIV—1 reverse transcriptase \/ 16L lmkm 2.8 8 T. maritime 0065, IclR family - 2L12H lbhi 2.5 8 DNA-bindinflotein cre-bpl - N/A 1di2 2.3 9 X laevis dsRNA binding protein A J 1L12H3L lytb 2.2 6 S. cerevisiae TATA-box binding protein J H/L 2gcc (Allen et al., 1998); 2bn8 (Chen et al., 2005); 1kjk(Wojciak et al., 2002); lzlb (Biswas et al., 2005); 2bb8 (Connolly et al., 1998); 2d35 (unpublished; http://www.rcsb.org/pdb/explore/explore.do?structureId=2D35); 1rth(Ren et al. , 1995); lmkm (Zhang et al., 2002); lbhi (Nagadoi et al., 1999); ldi2 (Ryter and Schultz, 1998); lytb (Kim et al., 1993). a Protein Data Bank identifier. b Degree of significant structural similarity (Z-score >2.0), according to Dali Ca matching. c percentage of sequence identity over positions of structural homology. d check marks under “Nucleic Acid bound in structure” indicate that the 3D structure of the protein is in complex with either DNA or RNA. e secondary structure of sequences N-terminal to their B-sheets, where available. H=helix; L=loop; N/A=structure of this region of the protein is not present or not defined in the crystal or NMR structure. 75 The similarity between three of them, represented by PDB identifiers lkjk and lzlb (representing the same protein), and 2bb8 and the AP2/EREBP domain of ERF] had previously been reported (Connolly et al., 1998; Wojciak et al. , 2002; Biswas et al., 2005; Chen et al. , 2005). An extensive comparative analysis of the DNA-bound structure of ERF 1 and Tn916 integrase (Wojciak et al., 1999) has previously indicated significant similarities in the mode of DNA recognition by these B-sheet containing proteins (Connolly et al. , 2000). This recognition involves a similar orientation of the B-sheet relative to major groove of the target DNA, and conserved protein-DNA contacts within the AP2/EREBP-like structure (Figure 2.14). We wanted to extend our analysis to the N-flanking regions, in search of protein- DNA complexes in which both the B-sheet and the N-flanking motif were involved in DNA-binding. Among the AP2/EREBP-like structures described in Table 2.3, lambda integrase and Tn916 integrase displayed this binding mode. NMR studies showed that the secondary structure of the N-flanking regions of these proteins differs from the helical structure predicted for the PKKPAGR region RKKFRET. This N-terminal region is represented by a loop in Tn916 integrase and by an unstructured segment in lambda integrase. However, despite the structural differences, some compelling similarities were found when we analyzed the amino acid composition of these motifs (Figure 2.15A, B). 76 Connolly et al., 2000 Figure 2.14.Three-dimensional structures of proteins that use three-stranded B-sheets to recognize the major groove of DNA. Left. The three-stranded B-sheet complexes of the DNA binding domain fi'om Tn916 integrase (left) and AtERFl (right). The Figure emphasizes the conserved orientation of the sheet structures relative to the DNA. The conserved B-sheets and the DNA backbone are dark and light grey, respectively. Right. DNA-binding domain of Tn916 in complex with DNA. labeled residues were mutated and the DNA binding activity of the mutated protein was tested by electrophoretic mobility shift assay. ArgS, in the N-flanking region of Tn916, contributes to DNA binding aflinity. DNA and protein backbones are shown as tubes, and the three- stranded B-sheet is represented by ribbons. N and C, N and C termini of the DNA- binding domain including the N-flanking region 77 A * Tn916-DBD B +++ +- + CBF1 (lgcc) PKKPAGRKKFRETRHPI - - - -YRGVRQR K-Int (lkjk) MGR_RRSHERRDLPP---NYRNNGY Tn916 (1tn9) MBKBR--DNRGRILKTGESQRKDG ' : : * Figure 2.15. Comparison of B-sheet—flanking sequences experimentally shown to be important in DNA binding. A. Schematic of the NMR derived structure of the Integrases Tn916 and Lambda (k-DBD and Tn9l6 DBD, respectively) and ERF] (GBD) as described by Wojciak et al. (1999). The NMR structures of ERFl and Tn916 were solved in complex with the DNA. The antiparallel B-sheet is blue, the C-terminal helix is red. Asterisks N-flanking motifs. B. Alignment of N-flanking regions from Lambda and Tn916 integrases and ERF]. Alignment is based on optimizing charged residue correspondence. Underlined residues are known to be important for binding. (*), exact conservation of residues; (z), strong conservation; (.), some degree of conservation in the residue chemistry. (+) and (—), positively and negatively charged residues, respectively. Area highlighted in grey, residues at the 5’ end of the beta sheets. lgcc (Allen et al., 1998); lkjk (Wojciak et al., 2002); 1b69 (Connolly et al., 2000). 78 Not only are the N-terminal residues proximal to or directly contacting the minor groove of the DNA in the NMR structure (Figure 2.15A), mutational analysis of residues in the N-flanking regions of these two proteins indicated that one or more positively charged arginines play important roles in DNA binding. For instance, mutation of an arginine residue in the N—flanking sequence of Tn9l6 integrase causes a significant loss in DNA-binding affinity (Connolly et al. , 2000), whereas deletion of two arginines in the [ES-sheet flanking region of lambda integrase could abrogate DNA-binding (Wojciak et al. , 2002). When we analyzed the sequence composition of those motifs we observed that the central core of those N-flanking regions includes a cluster of positively charged residues, and one negatively charged residue. In summary, we could identify a subgroup of B-sheet DNA-binding proteins in which one or more residues present in the N-flanking region provide a significant contribution to DNA-binding affinity. The presence of a positively charged cluster, their proximity to the DNA and their contribution to DNA binding, is consistent with our findings and supports the idea that the PKKPAGR motif provides a similar contribution at the CR T /DRE promoter site. 79 A working model for DNA binding by CBF1. Extensive mutagenesis and structural analysis of the PKKPAGR motif and comparison to AP2/EREBP-like proteins that use their N-flanking regions to contact the DNA, prompted us to describe CBF1 interaction with the DNA by computational modeling. The experimental results obtained by point mutations within the RKKFRET predicted helix drove the model design. The constraints used in modeling the orientation of the helix relative to CBF and the DNA were the following: (1) The C-terminus of the RKKFRET helix had to lie within two residues of the N-terminus of the B-stranded region of CBF; (2) Argl and Phe4 residues, critical to DNA binding, were placed in close proximity to the DNA, given that mutations to these positions have drastic effects on DNA binding. (3) The helix should pack well against the DNA and CBF1. The other features described below were derived from the placement of the RKKFRET helix to meet these three constraints. The NMR structure of DNA bound to the CBF1 homolog ERFl (PDB accession 1 gcc), was used as template to model the interaction between the DNA minor groove and the RKKFRET N-terminal flanking helix (Figure 2.16A). Figure 2163 illustrates the working model of CBF1 bound to the DNA. 80 Figure 2.16A. NMR structure of the AP2/EREBP DNA-binding domain of ERF] bound to its cognate DNA (Allen et al., 1998). Lefi and right figures represent two different views of the DNA binding domain in complex with its cognate promoter element. The three-stranded beta sheet is represented by ribbons numbered 1-3 that contact the DNA by reaching into its major groove. N and C termini of the AP2/EREBP domain are indicated by N and C, respectively. 5’ and 3’ end of the double-stranded DNA are indicated. 81 (Leslie Kuhn) Figure 2.16B. Working model describing DNA-binding by CBF1. The NTVIR structure of DNA bound to the CBF homolog ERF] (PDB id. lgcc), was used as the basis to model the interaction between the minor groove of DNA and the RKKFRET N-terminal flanking helix (yellow ribbon and side chains) relative to the main chain of the Iii-stranded DNA-binding domain of CBF1 (orange ribbon). The DNA is shown in tubes, with carbons in white, nitrogens in blue, oxygens in red, and phosphorus in magenta. In the RKKFRET helix, Arg 1 and Phe 4 residues bind within the DNA minor groove, where Argl forms hydrogen bonds and Phe4 in involved in stacking interactions. The main- chain structure of the B-strand domain of CBF was modeled with SwissModel. Two additional residues, Arg and His, bridge between the C-terminus of the KKFRET and the. N-terminus of the beta-stranded DNA-binding domain, as indicated by the orange line. This connection likely involves a slight reorientation of the first four residues of the beta-stranded region (orange ribbon), as observed in the structures of two DNA integrases (PDB id. lb69 and lzlb), which also use a B-stranded domain to bind DNA, with an N-terminal flanking region that enters the minor groove. 82 The AP2/EREBP domain and the RKKFRET region (orange and yellow, respectively) are placed against the DNA major and minor groove, respectively. The AP2/EREBP domain (shown in orange vertically oriented), is placed into the major groove, and its interaction with DNA forms a smooth pocket where the RKKFRET helix is packed. The proximity of this region to the DNA is also derived by the physical constraint of a short two-residue connection to the three-stranded B-sheet. Mutational analysis indicated that Argl and Phe4 are critically important for DNA binding. Argl——>Lys abrogated DNA binding, despite conserving side-chain length and charge. For the Phe mutations, Ala is helix-propensive and hydrophobic but less bulky than Phe, whereas Pro is a less conservative mutation and is likely to interrupt helicity. The loss of DNA binding upon these single-site mutations suggests that these residues interact intimately with the DNA, and both are positioned to bind within the minor groove. Phe—*Tyr mutation didn’t affect DNA binding. This is consistent with the phenyl moiety of Tyr maintaining favorable aromatic interactions with adjacent DNA bases, while its terminal hydroxyl group can both donate and accept hydrogen bonds either with bases in the groove of the DNA backbone. The phenyl moiety of Phe4 (or Tyr4) can make aromatic ring stacking interactions with adjacent bases in the groove, which have been reported when Phe interactions with DNA occur (Luscombe and Thornton, 2002; Moravek et al. , 2002) and this is likely to distort the local DNA conformation (Moravek et al. , 2002). Substitutions in the other residues of the RKKFRET motif resulted in a more moderate effect on DNA binding. Accordingly, Lys2 is shown salt bridging to the phosphate backbone and Lys3 is hydrogen bonding to a backbone ribose residue, consistent with mutations to these residues having moderate (non-base-specific) effects 83 on DNA binding, and less effect when the mutant is electrostatically conservative (Lys—>Arg) than when the side chain is truncated (Lys-+A1a). Arg5 and Glu6 are positioned such that they can form a salt bridge that stabilizes the helix, on the opposite side of the DNA. This is also consistent with Arg5—>Ser mutation having little effect on DNA binding, and with the Ser mutation also being able to form a stabilizing intra-helix hydrogen bond with Glu6. Based on the experimental constraints imposed on the model, the helical RKKFRET region can be placed close to the 3’end of the CRT/DRE motif - (G/A)CCGACNT. According to the model, Argl can make hydrogen-bond and pi-cation interactions with bases in the minor groove, corresponding to the region in which bases B11-B13 (from chain B) pair with C14-C16 (from chain C) in the 1 gcc structure (Figure 2.17). This Bl 1-B13/Cl4-C16 sequence follows the conserved consensus GCCGCC, representing the core sequence bound by ERF] via its B-sheet. Interestingly, the corresponding region from the CRT/DRE promoter element contains a conserved thymine that has been shown to be required for specific binding of CBFs to CR T/DRE-containing promoter elements (Maruyama et al. , 2004; Sakuma et al. , 2006). The proximity of the RKKFRET region to this DNA base suggests that either Argl or Phe4 could direct specific recognition of the DNA by CBF proteins. 84 CBF cognate DNA (CORISA) 5'- T G G C C G A C C T G - 3' Residue # in Chain B of lgcc 3 4 5 6 7 8 9 10 ll 12 13 ERF DNA in PDB lgcc 5'- T A G C C G C C A G C 3' ERF DNA in PDB 1 gcc 3'- A T C G G C G G T C G 5' Residue # in Chain C of lgcc 24 23 22 21 20 19 18 l7 16 15 I4 CBF cognate DNA (CORISA) 3'- A C C G G C T G G A C 5' Figure 2.17. Nucleotide sequences of the cis-acting elements in CBF and ERF proteins. The sequences shown for ERF DNA refer to the coding and complementary strand of DNA used in the NMR studies by Allen et al. (1998), in which the structure of the DNA- binding domain was solved in complex with the DNA. The conserved bases are shown in bold. 85 In summary, by integrating secondary structure predictions and experimental analysis, it was possible to describe CBF1 interaction with the DNA by computational modeling. We propose that the interactions of N-terminal flanking residues with the minor groove involve bases beyond those bound by the B-sheet, some of which might confer specificity for COR gene promoters relative to ERE. 86 FUNCTIONAL ROLE OF THE DSAWR MOTIF IN CBF1 ACTIVITY. Mutations in the DSAWR sequence affect COR gene activation. To determine whether the DSAWR motif plays a role in CBF1 function, Arabidopsis plants overexpressing dsawr mutant or wild type CBF1 transgenes were generated. Morphological analysis of those plants provided a first indication that residues within the conserved DSAWR motif might play a role in CBF1 function (Figure 2.3). The role of the DSAWR motif in CBF1 activity was assessed by northern blot analysis as described for the PKKPAGR motif. Representative results are shown in Figure 2.18. 87 CBF1 COR78 COR1 5a COR47 COR6.6 18$ 3 dsawr 86 621 G26 Figure 2.18. Accumulation of CBF and COR transcripts in Arabidopsis plants overexpressing wild type or dsawr CBF1. Total RNA was isolated from warm—grown seedlings (22°C), and 5 ug were analyzed by Northern blotting. Northern blot analyses were performed using probes for CBF1, COR6. 6, CORISA, COR47, and COR78 transcripts. B6, transgenic lines harboring the empty vector; WT, transgenic lines overexpressing wild type CBF1; dsawr, transgenic lines overexpressing CBF1 transgene carrying alanine substitutions in the DSAWR region. Numbers above lanes indicate independent transgenic lines. 88 Similarly to what was observed for the pkkpagr lines (Figures 2.4 and 2.5), Arabidopsis plants overexpressing CBF1-dsawr transgene still retained some ability to activate transcription at COR gene promoters, compared to the control plants (B6) harboring the empty vector only (Figure 2.18, compare B6 to independent dsawr lines). However, COR transcript accumulation was reduced when dsawr lines were compared to Arabidopsis lines overexpressing wild type CBF1 transgene Northern blot analysis showed that higher transcript levels of the mutated CBF1 were needed to achieve COR transcript accumulation similar to the wild type CBF1 overexpressing lines (Figure 2.18, compare WT and dsawr lines). Altogether these data indicated that alanine substitutions throughout the DSAWR motif of CBF1 affected its transcriptional activity. CBF1 protein detection in dsawr lines. To determine whether the dsawr mutation affected CBF1 protein accumulation, it was necessary to measure CBF1 protein levels in planta. The experiment was carried out by generating transgenic plants overexpressing wild type and mutated CBF1 transgenes fused to a Myc tag. Protein levels were measured by western blot analysis, as described for the pkkpagr lines. CBF1 gene and protein expression levels are shown in Figure 2.19. 89 CBF1 WT CBF1-dsawr o l O mooo—cmxo—rtx v-nrxooamxohoo ONxor—v—NNNNmm WOv-tv-tv-‘v-‘F-‘NNNN CBF1 'm‘ -'- “were . -4» .. COR6.6 ~...u..r_..u g... . 188 0.000....” W. CBF1 WT CBF1-dsawr c? '6‘ w v—i In l‘ V) \0 w 0 so .— N N .— N N N a-Myc “h- .— Ponceau an .: ”*Wfimtmmtzflwp~ Red ‘5" Figure 2.19. Northern and western blot analysis of dsawr plants and plants overexpressing wild type CBF1. Top. Northern blot analysis of 353::6xMyczCBF1 transgenic lines. Total RNA was analyzed for CBF1 and COR6. 6. Transcript amounts were normalized using the 183 loading control. Bottom. Western blot analysis showing the presence of 6xMyczCBFl protein in the transgenic plants (detected with a monoclonal anti-Myc antibody). WT and dsawr, transgenic plants overexpressing a 6xMyc tag fused to a wild type (WT) or mutated (dsawr) CBF1 transgene; Col-0, non-transgenic plants. 90 Several independent lines were analyzed for the expression of CBF1 and C 0R6. 6. Representative lines overexpressing similar levels of My-CBFI were chosen for protein detection from plant extracts. Western blot analysis clearly indicated that dsawr plants. were affected in CBF1 accumulation when compared to Arabidopsis plants overexpressing wild type CBF1 transgene at similar level (Figure 2.19, compare). Taken together, these results indicate that the inability of the dsawr lines to induce COR gene expression can be explained, at least in part, by reduced protein accumulation. Role of DSAWR in DNA binding activity of CBF1. Based on its proximity to the AP2/EREBP DNA-binding domain, and by analogy to the PKKPAGR sequence, we postulated that the DSAWR sequence might play a role in CR T/DRE binding by CBF1. Gel mobility shift assays were conducted to compare the binding activity of the wild type and mutated protein. Experiments were conducted as previously described for the PKKPAGR mutants. Binding reactions were set up by incubating wild type of mutated CBF1 proteins with Y32P-radiolabeled probe representing CRT/DRE elements from COR6. 6, COR15a, and COR 78 promoters. In all cases, alanine substitutions greatly reduced the binding activity of CBF1 , indicating that the DSAWR motif contributes to recognition (Figure 2.20). 91 WT dsawr WT dsawr WT dsawr mp4] A AAMBP 414] ”0““ 1 vol-.52.. I. “ ~u w Free probe CORISA COR78 COR6.6 Figure 2.20. DNA binding activity of an MBP-CBFlzulz protein carrying alanine substitutions in the DSAWR region. Upper panel. Gel shift mobility assay in which binding ofMBP:CBF127-1 12 proteins to COR6. 6, COR15a, and COR 78 is shown. 0.5 and 1.0 ug of protein were used in a 15.0 111 reaction in the presence of 0.5 ng of radiolabeled CRT/DRE-containing probes from COR6. 6, COR15a, and COR78 promoters. Lower panel. Western blot analysis showing protein loading. 25 and 50 ng of each recombinant protein were analyzed. Point mutations within the DSAWR region of CBF2 reveal the importance of an Asp residue in CRT/DRE recognition. To better understand the contribution of the DSAWR region to DNA binding by the CBF proteins, analyzing the effect of point mutations can be informative. In a mutant suppressor screen designed to identify regulators of the CBF pathway that act through the CBF regulon, several mutations in the CBF 2 transgene were identified (Gilmour, unpublished). Two mutations fell in the DSAWR region of CBF 2, Asp—>Asn and Arg—>Trp. Overexpression of those mutated transgenes in Arabidopsis produced plants that showed reduced accumulation of COR transcript compared to the plants overexpressing the wild type gene (Figure 2.21). 93 WT D/N WT R/W CBF1 MM '- --."" -rm..- 0!. COR15a an-” ' * a... .. .. 18s err-u quid i.e.-ea; mags-99 “I" Sarah Gilmour, unpublished Figure 2.21. Northern blot analysis of transgenic plants overexpressing wild type and mutated CBF2 transgenes. Total RNA was isolated from two-week old warm-grown (22°C) seedlings. Northern blot analysis was performed using probes for CBF1, COR15a, and 188 ribosomal RNA. WT, RNA samples from independent lines overexpressing wild type CBF1; DIN and WW, transgenic plants overexpressing a mutated CBF2 transgene harboring an Asp to Asn or Arg to Trp mutation, respectively. 94 Transcriptome profiling by Gilmour and colleagues (2004) revealed that CBF1, CBF 2, and CBF 3 display overlapping functions when overexpressed in Arabidopsis. In addition, the DNA—binding domains of CBF1 and CBF2 only differ in two residues, neither of which is involved in direct contact with DNA based on the AP2/EREBP structure described for AtERFl, so presumably the two DNA-binding domains can be considered equivalent. To determine the effect of Asp—>Asn and Arg—+Trp mutations on CBF 2 binding affinity, MBP:CBF227-112 fusions were subcloned in E. coli and the purified protein extracts tested in a gel mobility shift assay. As observed for CBF1, CBF2 could specifically bind a CRT /DRE-containing probe (Figure 2.22). The Asp—>Asn mutation, conservative in structure but with no charge, resulted in a significant loss of DNA binding, but retained its specific sequence recognition, as shown by the competition assay. Surprisingly, the Arg—>Trp substitution, despite the significant change in side chain structure, did not alter CBF2 binding affinity. 95 WT Asn/Arg Arg/Trp Protein —MBPA A A ————WTM——WTM——WTM C°"'P°“‘°" DNA -- In .- O WT Asp/Asn Arg/Trp MBP A /1 ._/_i MBP» - Q Q ‘4 MBP:CBF1 Figure 2.22. DNA binding activity of MBP:CBF 227-1 12 proteins canying point mutations in the DSAWR motif. Upper panel. Gel shift assay showing different DNA binding activity. Increasing amounts (150 and 450 ng) of wild type and mutated CBF2 proteins were tested in a 15 ul binding reaction containing 0.5 ng of radiolabeled CRT/DRE-containing probe from COR1 5a promoter. WT and M indicate a wild type and mutated unlabeled DNA (100 ng each), respectively used for the competition experiments indicated. Lower panel. Western blotting showing protein loading (15 and 45 ng of each protein were analyzed). a-MBP was used for protein detection. 96 Taken together, these observations suggest that the reduced ability to induce COR gene expression observed for the dsawr plants could be explained, in part, by inefficient binding of CBF1 to CR T/DRE-containing promoters. Based on the effect on protein accunulation observed when multiple alanines are introduced into the DSAWR motif, reduced ability to induce COR gene expression also appears to be due to reduced accumulation of the mutant protein. 97 DISCUSSION and FUTURE DIRECTIONS The main goal of this study was to determine whether two conserved signature sequences in the CBF family of proteins are required for CBF l transcriptional activity, and, if so, to elucidate their role. A mutational approach was taken to generate transgenic plants overexpressing wild type and mutated versions of the two motifs. Induction of several members of the CBF regulon - COR6. 6, COR15A, and COR 78 - was tested by northern blot analysis to evaluate the relevance of the signature sequences on CBF1 activity. In all cases tested, reduced COR/CBF1 transcript ratios in the pkkpagr lines compared to the transgenic plants overexpressing wild type CBF1 indicated that the PKKPAGR motif is required for CBF1 activity. Interestingly, we noticed a difference on the effect of these mutations at different COR gene promoters. For instance, COR15a induction was dramatically affected by all mutations tested compared to COR6. 6 and COR 78 induction. One possible explanation is that PKKPAGR confers higher binding affinity to different COR promoters, COR1 5a in this case. Once this motif has been mutated, the DNA binding function is lost or reduced and the effect will be more pronounced at the most highly induced genes. Alternatively, induction of COR15a-like genes might represent a sub-group of the CBF regulon that requires a CBF1-interacting factor that acts through the PKKPAGR motif. When the PKKPAGR motif is mutated, the interaction is lost and CBF 1 ability to recruit the transcriptional machinery will be greatly impaired. Yeast-two-hybrid screen by using the PKKPAGR motif as bait could help identify such factor. Overexpression of CBF1 in T-DNA knock- 98 out lines for that gene might show that a subset of COR15a-like genes is affected, similarly to what has been observed in the pkkpagr lines. Several hypotheses were postulated to assign a functional role to the PKKPAGR motif. Western blot analysis of transgenic plants overexpressing 6xMyc: CBF1 constructs ruled out the possibility that the PKKPAGR motif lowers CBF1 protein steady-state levels. On the contrary, substitution of the KKF residues with alanines resulted in higher CBF1 protein accumulation (Figure 2.5, compare CBF1 transcript and protein levels in WT and M3). Transcription factors tend to be very unstable proteins and it has been extensively reported that their activity and their turn-over can be tightly correlated (Tansey, 2001). Their degradation is usually mediated by the proteolytic activity of the proteasome, which recognizes a poly-ubiquitin chain attached to a lysine residue of the target protein. Increased CBF1 protein levels and reduced transcriptional activity when KKF residues are converted to alanines are consistent with a role of the ubiquitin pathway in modulating CBF1 activity. The obvious question is whether the difference in protein levels in wild type and mutated CBF1 proteins can be attributed to this pathway. To address this question it will be necessary to determine whether CBF1 is ubiquitylated in vivo, and whether the protein can accumulate to higher levels in the presence of specific proteasome inhibitors. If this is the case, the following step will be the identification of the ubiquitin ligase that is responsible for the addition of the ubiquitin chain that signals CBF1 degradation to the proteasome. Ultimately, the goal is to extend this analysis to wild type Arabidopsis plants and study the effect of low temperature on the regulation of CBF1 protein stability. 99 The PKKPAGR region, rich in basic residues, represents a potential NLS, and resembles NLSs previously identified in other plant transcription factors, including the member of the AP2/EREBP family AINTEGUMENTA (Krizek and Sulli, 2006). We tested whether the PKKPAGR motif played a role in the transport of CBF1 across the nuclear envelope by nuclear localization studies of Arabidopsis plants overexpressing translational fusions of CBF1 to GFP:GUS. Laser confocal microscopy of Arabidopsis plants overexpressing CBF1 APKK:GFPGUS revealed that this chimera is present in the nucleus of root tip cells, indicating that the PKKPAGR motif is not required for nuclear targeting of CBF1; on the contrary, localization studies of 5’ deletions of CBF1 fused to GFP:GUS revealed that in the absence of its N-terminal domain (residues 1-104) CBF1 :GFP:GUS loses its nuclear localization, thus mapping the nuclear targeting signal to the N-terminal of the protein which includes the DNA-binding domain. These data correlate with the fact that CBF1 1-1 15-VP1 6A1) can activate COR gene induction in stable Arabidopsis lines, and therefore must contain a nuclear targeting motif (Wang et al. , 2005). The involvement of the DNA-binding domain in nuclear targeting is not a new finding; it has been calculated that for 79% of DNA-binding proteins in which both NLS and DNA-binding domain have been characterized, the two functions map to the same protein domain and sometime the very same residues (LaCasse and Lefebvre, 1995). Some examples include the plant-specific SBP-domain proteins (Birkenbihl et al. , 2005), the yeast transcription factor GAL4 (Silver et al. , 1984; Carey et al. , 1989; Birkenbihl et al., 2005), and the transcription factors from the high-mobility group family (Sudbeck and Scherer, 1997; Prieve et al., 1998). Due to lack of a consensus signal for nuclear import, we could not identify any potential NLS in this region. However, the 100 AP2/EREBP domain contains several positively charged residues, mainly concentrated at its N-terminus, next to the PKKPAGR motif. It is possible that both the PKKPAGR region and this portion of the DNA-binding domain must be deleted to eliminate transport across the nuclear membrane. Based on its proximity to the DNA-binding domain, we hypothesized that the PKKPAGR motif might be helping CBF1 to bind to the CRT /DRE element, and found that triple alanine mutations within this region abrogated DNA-binding. Secondary structure predictions suggested that several residues in the PKKPAGR motif are likely to be coiled, while a seven-residue motif — RKKFRET - showed strong helical propensity. Point mutations within the RKKFRET motif revealed that specific amino acids are crucial for CBF1 DNA binding activity. Point mutations at either Argl or Phe4 (Arg—’Lys and Arg—>Ser and Phe—>Ala)’ caused the most dramatic loss in DNA-binding, whereas point mutations at LysZ and Lys3 (Lys—rArg and Lys—>Ala) had a minor effect and Arg5—*Ser substitution showed no significant effect. The CRT/DRE element is specifically recognized both by CBF and DREBZA proteins (Liu et al., 1998). A conserved thymine in the core sequence of this cis-acting element represents a specificity determinant for CBF1-dependent DNA binding activity (Sakuma et al. , 2006). Based on the NMR structure available for ERF 1 and the high homology among AP2/EREBP proteins, it was proposed that residues within this domain are involved in the specific binding to the CCGAC core of the CRT/DRE element (Allen et al., 1998 - verify); however the conserved thymine is located downstream of these DNA bases and therefore is unlikely to be contacted by residues within the AP2/EREBP domain. We integrated these observations with experimental data, obtained from mutational analysis and 101 secondary structure predictions, in a computational model that illustrates how CBF1 interacts with the DNA major groove through its B-sheet and with the DNA minor groove via the RKKFRET motif. In this model, the helical region of the PKKPAGR motif reaches into the DNA minor groove with both Argl and Phe4 oriented in close proximity to the conserved thymine. Arginines are frequently associated to base specific recognition, when present at the protein-DNA interface; this interaction occurs mainly via hydrogen bonding with a guanine, in fewer cases with thymine (Luscombe et al. , 2001; Luscombe and Thornton, 2002). Phenylalanines are not frequently found; however, when present they can contribute specific DNA recognition by providing stacking interactions with adjacent bases in the minor groove, preferentially at a thymine-adenine pair (Luscombe and Thornton, 2002; Moravek et al. , 2002). LysZ and Lys3, are placed in the vicinity of the DNA phosphates to provide favorable electrostatic interactions, whereas Arg5 does not participate to the protein-DNA complex. Overall, the experimental data can be nicely fit into this working model and explain how the PKKPAGR motif provides side chain interactions that are important or essential for CBF1 DNA binding activity. One important question that remains to be answered is whether Argl and/or Phe4 indeed represent specificity determinants and whether they are involved in direct recognition of the conserved thymine found in the CRT/DRE promoter element. NMR studies of CBF1 in complex with its cognate DNA will help shed some light on the nature of those interactions. A role for N-flanking regions in DNA-binding affinity has been recognized in homeodomain proteins (Wolberger, 1996). In this family, binding specificity is driven by a-helices mainly via hydrophobic interactions in the major groove of the target DNA, 102 whereas the N-terrninal tail makes contact with the DNA minor groove (Wolberger, 1996). DNA-binding by B-sheets has been extensively characterized at the structural and biochemical level in several DNA-binding proteins (Tateno et al. , 1998; Connolly et al. , 1998). However, most of the research has focused on defining the conserved structural features important for DNA binding within the B-sheet only. In contrast, very little is known about the role of their N-flanking sequences. Analysis of N-flanking sequences from other B-sheet-containing DNA-binding proteins revealed that these regions tend to be enriched in basic residues (Figure 2.14). Moreover, one or more residues in these regions are crucial for binding to the cognate DNA. These observations together with the results obtained with our mutational analysis suggest that the formation of protein-DNA complexes in the B-sheet family extends beyond the AP2/EREBP-like structure to include residues from the N-flanking region that provide important side chain interactions at the protein-DNA interface. Along these lines, identification of the binding determinants in the PKKPAGR motif could provide some insights into the mode of DNA binding in the AP2/EREBP family. To date, most of the characterization has been limited to the binding activity of the ERF and CBF proteins. It has been proposed that the specificity determinants in these two subfamilies lie within the AP2/EREBP DNA-binding domain (Hao et al. , 1998; Sakuma et al., 2002). However, high sequence similarity and the conservation of the residues that make direct contact with the DNA in these proteins suggests that the specificity switch may instead be found outside of this domain. Studies on several ERF proteins indicated that 10 amino acid residues upstream of the AP2/EREBP domain were essential for DNA binding (Hao et al. , 1998). Since proteins in the ERF group do not 103 share significant similarity in the region immediately upstream of the AP2/EREBP domain, the authors concluded that these residues do not contribute to specific binding, but instead their function is to stabilize the protein-DNA complex (Hao et al. , 1998). Interestingly, ERF proteins contain a cluster of positively charged residues flanking the AP2/EREBP domain at its N-terminus. It would be of interest to test the role of those basic amino acid residues in DNA binding by the ERF proteins. A large-scale comparison of the N-flanking sequences from members of this family combined with mutational analysis might reveal the presence of similar amino acids in other proteins within the AP2/EREBP family. In addition, sequence comparisons within different subgroups of the AP2/EREBP family could indicate whether there are any conserved residues that are specific for different subfamilies. If present, those residues would represent potential specificity determinants. Similarly to what had been observed for the pkkpagr lines, the growth retardation in the dsawr plants was not as severe as in for wild type CBF1-overexpressing plants. Consistently, COR gene induction was affected by this mutation. When overexpressed in Arabidopsis, dsawr CBF1 proteins do not accumulate as well as wild type CBF1 proteins. Inability to accumulate this protein in planta provides an explanation for reduced COR gene expression in the dsawr lines compared to the wild type CBF1-overexpressing lines. It is not clear at the moment whether this result is solely due to protein instability. Alanine substitutions within the DSAWR motif impaired CBF1 DNA binding activity. Analysis of two point mutation in the DSAWR motif - Asp—rAsn and Arg—>Trp - suggest that at least one residue (Asp) might play an important role in DNA binding. In fact the Asp—+Asn mutation, conservative in size though not in charge, may suggest that 104 no major change in protein folding occurred, and therefore a side chain interaction with the partner DNA might have been lost. On the contrary, mutation of Arg—*Trp did not show any effect on DNA-binding affinity. Interestingly, overexpression of a CBF 2 transgene carrying that mutation in Arabidopsis has a dampening effect on COR gene expression as compared with overexpression of wild type CBF2 (Figure 2.23). Whether protein stability is affected in these plants is currently unknown. An intriguing possibility is that this residue is critical for protein-protein interaction that is required for transcriptional activation at the COR gene promoters. 105 MATERIAL AND METHODS Mutagenesis of the PKKPAGR and DSAWR motifs. Alanine scanning mutagenesis. Site-directed mutagenesis (Li and Wilkinson, 1997) was performed to convert the original amino acids in the signature sequences of CBF1 to stretches of three alanines. Primers were designed to introduce the desired mutations to the PKKPAGR motif of CBF1 protein using the QuikChange mutagenesis kit (Stratagene), and the protocol was carried out according to the manufacturer instructions. A M restriction site was included in these primers for the screening of plasmids containing the mutated ORFs. The template DNA used was the full-length CBF1 ORF inserted into pBS/SK'.The primers used for the PKKPAGR mutants are the following: Mutant] (PKK), fwd (MT640): 5’ gcc acg agt tgt gcg gcc gca ccg gcg ggc cgt3’; rev. GVIT641) 5’acg gcc cgc cgg tgc ggc cgc aca act cgt gch’; Mutant2 — (PAGR), fwd. (MT642): 5’cga gtt gtc cga aga aag egg ccg ccg cta aga agt ttc gtg aga c3’; rev. (MT643): 5’ gtc tca cga aac ttc tta gcg gcg gcc gct ttc ttc gga caa ctt g3’; Mutant3 -— (KKF), fwd. (MT644): 5’ccg gcg ggc cgt gcg gcc gct cgt gag act cgt3’; rev. (MT645): 5’acg agt ctc acg agc ggc cgc acg gcc cgc cgg3’; Mutant4 (RET), fwd. (MT646): S’gcg ggc cgt aag aag ttt gcg gcc gct cgt cac cca att tac ag3’; rev (MT647): 5’ctg taa att ggg tga cga gcg gcc gca aac ttc tta cgg ccg cg3’; MutantS (RHP), fwd. (MT686): S’aag aag ttt cgt gag act gcg gcc gca att tac aga gga gtt cgt3’; rev (MT687): 5’acg aac tcc tct gta aat tgc ggc cgc agt ctc acg aaa ctt ctt3’. Primers used for DSAWR mutant: fwd. (MT817) 5’ gtc tca act tcg ctg ccg cgg ccg cgg cgc tac gaa tcc cgg ag3’; rev. (MT818): 5’ctc cgg gat tcg tag cgc cgc ggc cgc ggc agc gaa gtt gag ac3’. 106 The APKK mutant, that lacks the entire PKKPAGR region, was made using a modified protocol based on the QuikChange method (Wang and Malcolm, 1999). To overcome the tendency of the perfectly complementary mutagenic primers to anneal to each other rather than to the target sequence, a two-stage PCR was performed, running two separate single- primer reactions before the final PCR-amplification. Primers: fwd. (MT665) 5’taa atg tct ggt caa gca gtt tct ttg agc cg3’; rev. (MT666): 5’tgt tga gca ccg gtt gca gcc tgt tat tag ag3 ’. All constructs were verified by DNA sequencing. Point mutations in the RKKFRE T region of CBF1 . The point mutations in the RKKFRET motif were designed by using environment- specific substitution tables (Overington et al., 1992), which allowed to choose substitutions compatible with the helical structure. To introduce point mutations, the QuikChange mutagenesis protocol was followed as described in the Strategene manual. The following primers were used: Argl—>Lys, fwd: S’ccg aag aaa ccg gcg ggc aag aag aag ttt cgt gag act cg3’; rev: 5’cga gtc tca cga aac ttc ttc ttg ccc gcc ggt tct tcg3’; Argl—>Ser, fwd: 5’cga aga acc ggc ggg ctc gaa gaa gtt tcg tga gac tcg3’; rev: S’cga gtc tca cga aac ttc ttc gag ccc gcc ggt ttc ttc g3’; Lys2—rArg, fwd: 5’ccg gcg ggc aga aga aag ttt cgt gag act cg3’; rev: 5’cga gtc tca cga aac ttt ctt Actg ccc gcc gg3’; LysZ—>Ala, fwd: 5’ccg gcg ggc aga gcg aag ttt cgt gag act cgt cac3’; rev: 5’gtg acg agt ctc acg aaa ctt cgc tct gcc cgc cgg3’; Lys3—Arg, fwdz5’ccg gcg ggc aga aag aga ttt cgt gag act cgt cac c3’; revz5’ggt gac gag tct cac gaa atc tct ttc tgc ccg cc gg3’; Lys3—>A1a, fwd: 5’ccg gcg ggc aga aag gcg ttt cgt gag act cgt cac c3’; rev: 5’ ggt gac gag tct cac gaa acg cct ttc tgc ccg ccg g3’; Phe4—)Ala, fwd: S’ggg ccg taa gaa ggc tcg aga gac tcg tca ccc3’; rev: 5’ ggg tga cga gtc tct cga gcc ttc tta cgg ccc3’; Phe4—>Pro, fwd: 5’ gcg ggc cgt aag aag cct cga gag act cgt cac cc3’; rev: 5’ ggg tga cga gtc 107 tct cga ggc ttc tta egg ccc gc3’; Phe4—>Tyr, fwd: 5’ccg gcg ggc aga aag aag tac cgt gag act cgt c3’; rev: 5’ gac gag tct cac ggt act tct ttc tgc ccg ccg g3’; Arg5—+8er, fwd: 5’ccg gcg ggc aga aag aag ttt agt gag act cgt cac c3’; rev: 5’ ggt gac gag tct cac taa act tct ttc tgc ccg ccg g3’. For the screening of the clones harboring the desired mutation, the primers were designed to disrupt a pre-existing Sau96I restriction site in all mutant ORFs, except for Phe4—+Ala and Phe4—rPro mutations, where a new XhoI restriction site was inserted. Preparation of constructions for plant transformation. Plant expression vector harboring wild type and mutated CBF1. BglII restriction ends were added by PCR in order to subclone the ORF of wild type or mutated CBF1 into pGA643 (An et al. , 1988) which harbors the constitutive 358 CaMV promoters. Primers BglII fwd: 5’gaa gat cta tga act cat ttt cag ctt ttt ctg3’; BglII rev: 5’ gaa gat ctc tcg ttt cta caa caa taa aat aaa3’. A11 constructs were verified by DNA sequencing. Plant expression vector harboring 6xMyc: CBF1 ORFs. F usional translation of wild type and mutated CBF1 to 6xMyc, were generated by subcloning a SmaI/Sacl CBF1 ORF into the binary vector pKVB24 (V lachonasios et al. , 2003) containing a 6xMyc tag, under the control of CaMV 35S promoter. The translational fusion results in 6xMyc at the N-terminus of CBF1. Primers used to add SmaI/SacI ends by PCR amplification were: MT708, fwd: 5’agc ccg ggg atg aac tca tt3’ MT709, rev.: 5’cag agc tct tac taa ctc ca3’. Plant expression vector harboring CBF1:GFP:GUS. 108 Wild type CBF1 and CBFIAPKK ORFs were pEZT-CL(GUS). The vector was engineered by subcloning a BamHI GUS insert into the pEZT-CL plant expression vector (Schena et al. , 1991), resulting in an in frame fusion to GFP under the control of the CaMV 35 S promoter. The eGFP gene in pEZT is based on mGFP4 (Haseloff et al. , 1997) and contains additional mutations (S65T, Y66H) to increase intrinsic GFP fluorescence (Cormack et al. , 1996). BamHI restriction ends were added to GUS by PCR amplification using the following primers: fwd. (MT794): 5’cag gat ccg cat cga taa gct tga att cac c3’; rev. (MT795): 5’aga gga tcc cca att ccc gag gct gta3’. Full-length and 5’ deletions of CBF1 ORF were subcloned into the XhoI restriction site of pEZT-CL(GUS). In frame fusions of CBF1 to GFP:GUS were generated by inserting an XhoI fragment into the binary vector. A PCR approach was used to add the proper restriction ends, as follows: firll-length CBF 1 XhoI 5’ end (MT704): 5’ gcc tcg aga tga act cat ttt cag3’; XhoI 3’ end (MT825): 5’ccc tcg agg cgt aac tcc aaa gcg3’For the 5’ ORF deletions the following forward primers were used: construct 802 (MT802): S’acc tcg aga tgc cga aga aac cgg cgg gcc g3’; construct 801 (MT801): S’acc tcg aga tga ttt aca gag gag ttc g3’; construct 800 (MT800): 5’acc tcg aga tgg act cgg ctt ggc ggc tac g3’; construct 799 (MT799): 5’acc tcg aga tgc tac gaa tcc cgg agt caa c3’. The XhoI insert for NIa was PCR-amplified from the yeast plasmid pAVA367, containing the in frame fusion Nia-GFP (Schena et al. , 1991), with the following primers: fwd. (MT815): S’acc tcg aga tgg gga aga aga atc aga agc aca agc taa aga tga g3’; rev. (MT826): 5’ccc tcg agt gac ctg tca atg gat cca 03’. Plant transformation. All the transgenic lines used in this study were generated by the floral dip method described by Clough and Bent (1998). pEZT-CL(GUS) or pKVBZ4 constructs were 109 transformed into the A. tumefaciens strain LBA4404, whereas pGA643 constructs were transformed into strain GV3101. Generation and selection of transgenic lines was performed as follows. Arabidopsis plants (T0) were grown in soil to bolting and then transformed by A. tumefaciens-mediated transformation using the floral dip method (Clough and Bent, 1998). Seeds from T0 plants were screened for antibiotic or herbicide resistance (600uM Basta; 50ug/ml Kanamycin). Resistant T1 seedlings were grown to maturity to produce T2 seeds. T2 seeds were plated onto selective media containing antibiotic or herbicide and screened for a 3:1 survival ratio, indicating that there was an insertion in a single locus in the T1 genome. Homozygous lines were selected by growing plants on selective media and screened for 100% survival. Growth conditions and Northern blot analysis of transgenic Arabidopsis plants. Seeds were sterilized and stratified for 3 days at 4°C, and then plated on Gamborg’s plates supplemented with BS nutrients and 23.21 gr/liter of sucrose (Caisson Laboratories, Inc., Rexburg, ID, USA) and 0.8% phytagar (Life Technologies Inc., Gaithersburg, MD, USA). Seedlings were grown under continuous cool-white fluorescent light (100umol-m'2-s") at 22°C for two weeks. Tissue samples were harvested in liquid Nitrogen and total RNA extracted using either the RN easy Plant Miniprep kit (QIAGEN) or the Trizol reagent (Life Technologies, Gaithersburg, MD). Five to ten micrograms of total RNA were fractionated in 1% formaldehyde gels and transferred onto nitrocellulose membranes as described (Sambrook et al. , 1989). Membranes were hybridized in Church buffer (1% BSA, 1 mM EDTA, 0.5 M NaPO4 pH 7.2, 7% SDS) (Church and Walter, 1984) at 65°C, overnight. Blots were probed with a 32P-labelled fragments prepared using 110 the Random Primers DNA Labeling System (Invitrogen), according to the manufacturer instructions. Probes for all the transcripts were obtained by using the full-length ORF as a probe. After hybridization, membranes were washed three times in .IXSSC, 0.1%SDS at 55°C, for 15 minutes. mRN A levels in different samples were normalized by comparing them to the levels of 188 rRNA determined from the same blots. Following the washes, the membranes were exposed to a phosphorimager screen; the screen was later scanned and then quantified using a QuantityOne software from BioRad (Hercules, CA, USA). Analysis of variance (ANOVA) of COR/CBF1 transcript ratios. AN OVA tests were carried out using to determine the effect of mutations in the PKKPAGR motif on COR78/CBF1 mRNA ratios. mRNA values for CBF1 and different COR genes were obtained by exposing radioactive membranes to a phosphorimager screen. The screen was later scanned and the photointensity of different bands was measured using a QuantityOne software from BioRad (Hercules, CA, USA). Variables were log-transformed to meet the normality assumptions of statistical analyses. Each analysis was conducted by averaging ratios from two technical replicates. Each group analyzed was represented by five or six independent transgenic lines. Statistical difference in COR/CBF1 ratios of transgenic plants overexpressing a wild type CBF1 transgene or CBF1 transgenes mutated in the signature sequences was analyzed by analysis of variance (AN OVA) using SAS Proc Mixed procedures (SAS Institute, Cary, NC), version 9.1. To test for significant difference between COR/CBF1 ratios between transgenic plants overexpressing wild type CBF1 or the mutated versions described 111 earlier we estimated least-square means for each genotype and compared them to the least-square means of the control plants overexpressing wild type CBF1. These estimates were used to calculate the t-values and the statistical significance at P<0.0001 for all the transgenic lines analyzed. Subcloning of CBF1 mutant ORFs into bacterial expression vectors for expression _ and purification of 6xHiszCBFl and MBP:CBF1 proteins. Expression and purification of 6xHis:CBF 1 proteins. All CBF1 ORFs described earlier were introduced into the pET28a+ expression vector (N ovagen/EMD, San Diego, CA) under the control of the bacteriophage IPTG-inducible T7 promoter. The constructs were generated by NcoI/XhoI cleavage of the pSJG6 construct (ref), in which CBF1 is fused to 6His-T7 tag. The NcoI/XhoI fragments were ligated onto pET28a+. Sequences were confirmed for all constructs and plasmids were transformed into BL21-CodonPlus(DE3)-RIL (Stratagene, La Jolla, CA) E. coli strain for optimal expression of the recombinant proteins (Kleber-Janke and Becker, 2000). Protein expression and purification was started by inoculation of individual colonies containing plasmid DNA in 3mL of Luria-Bertani broth containing 50ug/ml Kanamycin and 40ug/ml Chloramphenicol (Sigma-Aldrich, St.Louis, MO) and grown at 37°C, overnight (250 rpm). 50ul of that culture were inoculated into 500ml of the same broth to an 0ng of approximately 0.9 before addition of lmM IPTG (Research Organics, Cleveland, OH) to induce protein expression for 4 hours, at 37°C. Cells were harvested by centrifugation, and resuspended in 101111 of cell lysis buffer (SOmM NaHzPOr-HzO, 300mM NaCl, IOmM Imidazole, pH8.0). Cells were lysed by sonication (BRANSON Sonifier150, Pittsburgh, PA) in the presence of 112 protease inhibitors (Complete EDTA-free protease inhibitor cocktail tablets, Roche, Mannheim, Germany). The soluble protein fiaction was separated by centrifugation, at 4°C. The supernatant fraction containing 6xHis-T7-CBF1 proteins was loaded over a nickel column equilibrated With wash buffer (SOmM NaHzPO4'HzO, 300mM NaCl, 20mM Imidazole, pH8.0). After three washes in wash buffer, 6xHis-T7-CBF1 proteins were eluted by washing with elution buffer (SOmM NaHzPO4-HZO, 300mM NaCl, 250mM Imidazole, pH8.0). Protein concentration was determined using the Bradford dye binding assay (Bio- Rad, Hercules, CA), with BSA used as standard. Expression and purification ofMBP:CBF127.1 12 proteins A 258bp anI/Xbal CBF1 fragment (aa27-112) was cloned into pMAL-c2x vector (New England BioLabs, Beverly, MA), downstream of the malE gene, which encodes Maltose Binding Protein (MBP) under the control of the strong Pm promoter. this resulted in the translational fusion of CBF1 to the C—terminus of MBP. The primers used for wild type CBF1 and all the PKKPAGR mutants were: MT908(fwd): S’ATA TTT TCT AGA TTC GTA GCC3’, and MT909(rev.): 5’CCA TGG AAG GAT TTC GGC CAC GAG TTG T3’. For the dsawr CBF1 mutant, MT910 was used as the reverse primer: 5’ATA TTT TCT AGA TTC GTA GCG C3’. Individual transformed cells were grown at 37°C in 3ml Rich Medium (10g tryptone, 5 g yeast extract 5g NaCl, 2g glucose, per liter) supplemented with lOOug/ml Ampicillin to an CD. of 0.5. Protein expression was induced by addition of lmM IPTG to the bacterial suspension, and let grow for 3 additional hours. Cells were harvested by centrifugation (6,000 rpm, 10min., 4°C), and lysed by sonication, as described above. The soluble protein fraction was separated by centrifugation at 13,000rpm, 30 min., at 4°C. The supernatant containing the fusion 113 proteins was loaded over an amylase column for protein purification by affinity. The amylase column was equilibrated in column buffer (20mM Tris-HCl, 200mM NaCl, lmM EDTA). MBP-CBF1 proteins were eluted in column buffer containing lOmM maltose. Protein concentration was determined using the Bradford dye binding assay (Bio-Rad, Hercules, CA). Protein isolation and immunoblot analyses. Immunodetection of CBF1 protein fiom E. coli extracts. Western Blotting and protein detection were carried out by first separating the proteins on a 10% bis-acrylamide tricine gel. Proteins were transferred in Towbin buffer (25mM2 Tris, 192mM Glycine, 15% MeOH, 0.02% SDS) to a PVDF membrane (Irnmobilon-P, Millipore corporations, Bedford, MA) using by electrophoretic transfer (Genie®, Idea Scientific, Minneapolis, MN), for 45min. at lAmp, 30 volts. The membrane was blocked in PBS-Tween (per liter: 80g NaCl, 2gr KCl, 14,4gr NazHPO4, 2.4gr KHzPO4, pH7.2, 0.1% (v/v) Tween 20) and 5% non-fat dried milk for 1h at RT with gentle rocking. Incubation with the primary antibody was conducted for 1h at room temperature in PBS-T and 5% non-fat dried milk for 1h at RT. Antibody dilutions were: 1/7,500 for a-CBFI (F1, IgG purified CBF1 antiserum raised against the full- length CBF1 protein fused to a 6xHis tag); 1/ 10,000 for a-MBP (New England Biolabs, Beverly, MA). The presence of the protein on the membrane was assayed by chemiluminescence detection (Amersham Biosciences, Piscataway, NJ). Immunodetection of CBF1 protein fi'om plant extracts. 114 Total protein extracts were prepared from two-week old Arabidopsis seedlings by grinding frozen tissue in protein extraction buffer (20 mM Tris-HCl, pH 8.0, SOmM NaCl, 5 mM EDTA, 0.05 % SDS) supplemented with protease inhibitor cocktail from Roche. Protein concentration was determined using the Bradford reagent (Bio—Rad, Hercules, CA, USA) with BSA as the standard. Equal amounts of proteins were subjected to SDS-polyacrylamide gel electrophoresis (PAGE) (4-20% gradient gel). After electrophoresis, the proteins were transferred to polyvinylidene difluoride (PVDF) membranes by electrotransfer. Membranes were blocked in 5% nonfat milk powder in Tris-buffered saline (TBS) - 0.1% Tween-20 for l h at room temperature, and then incubated overnight at 4°C with specific antibodies. Antibody dilutions were: 1/5,000 for the a-Myc monoclonal antibody Roche-Mannheim, USA), and 1/10,000 for the a-GFP monoclonal antibody (Pierce Biotechnology, Rockford, IL). The next day, membranes were washed three times and incubated in horseradish peroxidase-coupled anti-mouse IgG antibody [1/ 10,000 dilution of the rabbit anti-mouse IgG (Pierce Biotechnology, Inc., Rockford, IL)] for 1 h at room temperature. Immunoreactive bands were visualized by an enhanced chemiluminesence assay following the manufacturer's inst2ructions. Electrophoretic-Mobility Shift Assays (EMSA). 24-bp fragments from COR15a, COR 78 and COR6.6 promoters containing a CR T/DRE element were prepared by synthesizing both strands. The oligomers were suspended in 1X STE buffer (10mM Tris, pH8.0; 10mM NaCl; lmM EDTA), and the complementary strands were annealed. The resulting double strand oligonucleotides were labeled with [y-32P]ATP (NEN Life Science Products) by T4 polynucleotide kinase (New 115 England Biolabs, Beverly, MA) and purified through a Sephadex G-50 column. Mutant and wild type sequences were prepared following the same procedure. The DNA oligomers used for the binding assays are the following: COR15a: 5’ATT TCA TGG w CTG CTT TTT3’; COR78: S’AAT ATA CTA w ATG AGT TCT 3’; COR6. 6: S’AAA AAG CTA CCGAC ATA AGC CAA3’. A mutant version of COR15a in which the entire C-repeat, CCGAC, had been altered was used for all the competition assays: 5’ ATTTCATGGtatgtCTGCT'ITTT3’. The binding of 6xHis:CBF] or MBP:CBF127-112 to the CRT/DRE-containing DNA probes was tested in a total volume of 12.0 111 as follows: 0.5 ng of a 32P-labeled DNA probe encoding the CRT/DRE binding site was mixed with a gradient concentration of each recombinant protein (100 - 300 ng of 6xHis:CBF 1 and 0.3 — 15 ug for MBP:CBF127-112) and incubated in binding buffer (20 mM Tris-HCl, pH 8.0; 100 ug/ml BSA; 30mM KCl; 5mM MgClz; 4% Glycerol; lmM DTT) in the presence/absence of 100 ng unlabeled competitor DNA. Afier incubation at room temperature for 20 min., samples were loaded into nondenaturing polyacrylamide gels (4% wt/vol), and fractionated by electrophoresis at 150V, for 3 hours at 4°C. The gels were dried at 80°C, 30 min., and exposed to a phosphorimager screen. The screen was later scanned and the photointensity of complexed DNA was measured using a QuantityOne software from BioRad (Hercules, CA, USA). Fluorescence imaging of Arabidopsis root tips overexpressing CBF1 :GFP:GUS transgenes. For cellular localization of CBF1 :GFP:GUS proteins, Arabidopsis seeds were sterilized, stratified for 3 days at 4°C, and plated on Gamborg agar plates supplemented 116 with sucrose as described above. The Petri dishes were oriented vertically and the seedlings grown for 6-7 days before being imaged. Laser confocal images were collected using an upright LSM Zeiss 510 META microscope equipped with a 40X oil immersion objective. To visualize the nuclei, Arabidopsis seedlings were incubated with a solution of 50 ug/ml Propidium Iodide (PI) and 5 rig/ml RNase, in the dark for 30 min — 1 hr. Samples were rinsed with distilled water a couple of times and mounted in tap water. GF P fluorescence images were obtained using Argon ion laser excitation of 488 nm with a 505/530 nm bandpass filter. PI fluorescence images were collected using an excitation line of 543 nm with a 560 nm longpass filter. Postacquisition image processing was done with the LSM 5 Image Browser (Zeiss) and Adobe Photoshop 5.0 software (Mountain View, CA). 117 LITERATURE CITED Abel, S., Oeller, P., and Theologis, A. (1994). Early auxin-induced genes encode short- lived nuclear proteins. PNAS 91, 326-330. Adamczak, Porollo A, and J, M. (2005). Combining prediction of secondary structure and solvent accessibility in proteins. Proteins: Structure, Function, and Bioinforrnatics 59, 467-475. Allen, M., Yamasaki, K., Ohme—Takagi, M., Tateno, M., and Suzuki, M. (1998). A novel mode of DNA recognition by a beta-sheet revealed by the solution structure of the GCC-box binding domain in complex with DNA. EMBO J 17, 5484-5496. An, G., Ebert, P., Mitra, A., and Ha, S. (1988). Binary vectors. (Dordrecht: Kluwer Academic Publishers). Birkenbihl, R., Jach, G., Saedler, H., and Huijser, P. (2005). Functional dissection of the plant-specific SBP-domain: overlap of the DNA binding and nuclear localization domain. JMB 352, 585-596. Biswas, T., Aihara, H., Radman-Livaja, M., Filman, D., Landy, A., and Ellenberger, T. (2005). A structural basis for allosteric control of DNA recombination by [lambda] integrase. Nature 435, 1059. Carey, M., Kakidani, H., Leatherwood, J., Mostashari, F., and Ptashne, M. (1989). An amino-terminal fragment of GAL4 binds DNA as a dimer. Journal of Molecular Biology 209, 423. Carrington, J.C., Freed, D.D., and Leinicke, A.J. (1991). Bipartite signal sequence mediates nuclear translocation of the plant potyviral Nla protein. Plant Cell 3, 953-962. Chakrabartty, A., Kortemme, T., and Baldwin, R.L. (1994). Helix propensities of the amino acids measured in alanine-based peptides without helix-stabilizing side- chain interactions. Protein Sci 3, 843-852. Chen, H.A., Simpson, P., Huyton, T., Roper, D., and Matthews, S. (2005). Solution structure and interactions of the Escherichia coli Cell Division Activator protein CedA. Biochemistry 44, 673 8-6744. Chou, P.Y., and Fasman, GD. (1978). Empirical predictions of protein conformation. Annual Review of Biochemistry 47, 251-276. Christophe-Hobertus, C., Duquesne, V., Pichon, B., Roger, RR, and Christophe, D. (1999). Critical residues of the homeodomain involved in contacting DNA bases also specify the nuclear accumulation of thyroid transcription factor-1. European Journal of Biochemistry 265, 491-497. 118 Church, G., and Walter, G. (1984). Genomic sequencing. PNAS 81, 1991-1995. Clough, S.J., and Bent, A.F. (1998). Floral dip: a simplified method for Agrobacterium- mediated transformation of Arabidopsis thaliana. The Plant Journal 16, 735-743. Collawn, J.F., Stangel, M., Kuhn, L.A., Esekogwu, V., Jing, S.Q., Trowbridge, 1.8., and Tainer, J.A. (1990). Transferrin receptor internalization sequence YXRF implicates a tight turn as the structural recognition motif for endocytosis. Cell 63, 1061-1072. _ Connolly, K.M., Wojciak, J.M., and Clubb, RT. (1998). Site-specific DNA binding using a variation of the double stranded RNA binding motif. Nat Struct Biol 5, 546-550. Connolly, K.M., Ilangovan, U., Wojciak, J.M., Iwahara, M., and Clubb, RT. (2000). Maj or groove recognition by three-stranded [beta]-sheets: afiinity determinants and conserved structural features. Journal of Molecular Biology 300, 841-856. Cormack, B.P., Valdivia, R.H., and Falkow, S. (1996). FACS-optimized mutants of the green fluorescent protein (GFP). Gene Flourescent Proteins and Applications 173, 33-3 8. Craig, L., Sanschagrin, P.C., Rozek, A., Lackie, S., Kuhn, L.A., and Scott, J.K. (1998). The role of structure in antibody cross-reactivity between peptides and folded proteins. Journal of Molecular Biology 281, 183-201. Crane-RobinsOn, C., Dragan, A.I., and Privalov, P.L. (2006). The extended arms of DNA-binding domains: a tale of tails. Trends in Biochemical Sciences 31, 547. DeWitt, N.D., Hong, B., Sussman, M.R., and Harper, J.F. (1996). Targeting of two Arabidopsis H+-ATPase isoforrns to the plasma membrane. Plant Physiol. 112, 833-844. Dragan, A., Read, C., Makeyeva, E., Milgotina, E., Churchill, M., Crane-Robinson, C., and Privalov, P. (2004). DNA binding and bending by HMG boxes: energetic determinants of specificity. Journal of Molecular Biology 343, 371. Eyrich, V.A., Marti-Renom, M.A., Przybylski, D., Madhusudhan, M.S., Fiser, A., Pazos, F., Valencia, A., Sali, A., and Rost, B. (2001). EVA: continuous automatic evaluation of protein structure prediction servers. Bioinforrnatics 17, 1242-1243. Gilmour, S.J., Fowler, S.G., and Thomashow, M.F. (2004). Arabidopsis transcriptional activators CBF1, CBF2, and CBF3 have matching functional activities. Plant Mol Biol 54, 768-781. Gilmour, S.J., Sebolt, A.M., Salazar, M.P., Everard, J .D., and Thomashow, M.F. (2000). Overexpression of the Arabidopsis CBF3 transcriptional activator mimics 119 multiple biochemical changes associated with cold acclimation. Plant Physiol. 124, 1854-1865. Gilmour, S.J., Zarka, D.G., Stockinger, E.J., Salazar, M.P., Houghton, J.M., and Thomashow, M.F. (1998). Low temperature regulation of the Arabidopsis CBF family of AP2 transcriptional activators as an early step in cold-induced COR gene expression. The Plant Journal 16, 433-442. Gilmour SJ, Fowler SG, and MF, T. (2004). Arabidopsis transcriptional activators CBF1, CBF2, and CBF3 have matching functional activities. Plant Mol Biol 54, 768-781. Grebenok, R., Pierson, E., Lambert, G., Gong, F., Afonso, C., Haldeman-Cahill, R., Carrington, J., and Galbraith, D. (1997). Green-fluorescent protein fusions for efficient characterization of nuclear targeting. Plant J 11, 573-5 86. Guana, C., Lib, P., Riggsa, P., and Inouyeb, H. (1987). Vectors that facilitate the expression and purification of foreign peptides in Escherichia coli by fusion to maltose-binding protein. Gene 67, 21-30. Haake, V., Cook, D., Riechmann, J.L., Pineda, O., Thomashow, M.F., and Zhang, J.Z. (2002). Transcription factor CBF4 is a regulator of drought adaptation in Arabidopsis. Plant Physiol. 130, 639-648. Hao, D., Ohme-Takagi, M., and Sarai, A. (1998). Unique mode of GCC box recognition by the DNA-binding domain of ethylene-responsive element-binding factor (ERF domain) in plant. J Biol Chem 273, 26857-26861. Haselofl', J., Siemering, K.R., Prasher, D.C., and Hodge, S. (1997). Removal of a cryptic intron and subcellular localization of green fluorescent protein are required to mark transgenic Arabidopsis plants brightly. PNAS 94, 2122-2127. Holm, L., and Sander, C. (1993). Protein structure comparison by alignment of distance matrices. J Mol. Biol. 233, 123-138. Jaglo-Ottosen, K.R., Gilmour, S.J., Zarka, D.G., Schabenberger, 0., and Thomashow, M.F. (1998). Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 280, 104-106. Jaglo, K.R., Kleff, S., Amundsen, K.L., Zhang, X., Haake, V., Zhang, J.Z., Deits, T., and Thomashow, M.F. (2001). Components of the Arabidopsis C- Repeat/Dehydration-Responsive Element Binding Factor cold-response pathway are conserved in Brassica napus and other plant species. Plant Physiol. 127, 910- 917. Jofuku, K.D., Boer, B.G., Montagu, M.V., and Okamuro, J.K. (1994). Control of Arabidopsis flower and seed development by the homeotic gene APETALA2. PlantCell6, 1211-1225. 120 Kaffman, A., and O'Shea, E. (1999). Regulation of nuclear localization: a key to a door. Annual Review of Cell and Developmental Biology 15, 291-339. Karplus, K., Draper, J., Casper, J., Mandel-Gutfreund, Y., Diekhans, M., and Hughey, R. (2003). Combining local-structure, fold-recognition, and new fold methods for protein structure prediction. Proteins: Structure, Function, and Genetics 53, 491-496. Kim, J., Mao, J., and Gunner, M. (2005). Are acidic and basic groups in buried proteins predicted to be ionized? Journal of Molecular Biology 348, 1283. Kim, Y., Geiger, J.H., Hahn, S., and Sigler, PB. (1993). Crystal structure of a yeast TBP/TATA-box complex. Nature 365, 512. Kleber-Janke, T., and Becker, W.-M. (2000). Use of modified BL21(DE3) Escherichia coli cells for high-level expression of recombinant peanut allergens affected by poor codon usage. Protein Expression and Purification 19, 419. Klucher, K.M., Chow, H., Reiser, L., and Fischer, R.L. (1996). The AINTEGUMENTA gene of Arabidopsis required for ovule and female gametophyte development is related to the floral homeotic gene APETALA2. Plant Cell 8, 137-153. Koroleva, 0., Tomlinson, M., Leader, D., Shaw, P., and Doonan, J. (2005). High- throughput protein localization in Arabidopsis using Agrobacterium-mediated transient expression of GFP-ORF fusions. Plant J 41, 162-174. Krizek, B., and Sulli, C. (2006). Mapping sequences required for nuclear localization and the transcriptional activation function of the Arabidopsis protein AINTEGUMENTA. Planta 224, 612-621. LaCasse, E., and Lefebvre, Y. (1995). Nuclear localization signals overlap DNA- or RNA-binding domains in nucleic acid-binding proteins. Nucleic Acids Res. 23, 1647-1656. Li, S., and Wilkinson, M. (1997). Site-directed mutagenesis: a two-step method using PCR and DpnI. Biotechniques 23, 588-590. Liu, Q., Kasuga, M., Sakuma, Y., Abe, H., Miura, S., Yamaguchi-Shinozaki, K., and Shinozaki, K. (1998). Two transcription factors, DREBI and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10, 1391-1406. Lnenicek-Allen, M., Read, C., and Crane-Robinson, C. (1996). The DNA bend angle and binding affinity of an HMG box increased by the presence of short terminal arms. Nucl. Acids Res. 24, 1047-1051. 121 Luscombe, N., and Thornton, J. (2002). Protein-DNA interactions: amino acid conservation and the effects of mutations on binding specificity. Jom'nal of Molecular Biology 320, 991 . Luscombe, N., Laskowski, R., and Thornton, J. (2001). Amino acid-base interactions: a three-dimensional analysis of protein-DNA interactions at an atomic level. Nucl. Acids Res. 29, 2860-2874. Lyck, R., Harmening, U., Hohfeld, 1., Trenter E, Scharf KD, and L, N. (1997). Intracellular distribution and identification of the nuclear localization signals of two plant heat-stress transcription factors. Planta 202, 117-125. Magome, H., Yamaguchi, S., Hanada, A., Kamiya, Y., and Ode, K. (2004). dwarf and delayed-flowering 1, a novel Arabidopsis mutant deficient in gibberellin biosynthesis because of overexpression of a putative AP2 transcription factor. The Plant Journal 37, 720—729. Maruyama, K., Sakuma, Y., Kasuga, M., Ito, Y., Seki, M., Gods, H., Shimada, Y., Yoshida, S., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2004). Identification of cold-inducible downstream genes of the Arabidopsis DREBlA/CBF3 transcriptional factor using two microarray systems. The Plant Journal 38, 982-993. McGuffin, L.J., Bryson, K., and Jones, D.T. (2000). The PSIPRED protein structure prediction server. Bioinformatics 16, 404—405. Molar/ck, 2., Neidle, 8., and Schneider, B. (2002). Protein and drug interactions in the minor groove ofDNA. 30, 1182-1191. Nadassy, K., Wodak, S., and Janin, J. (1999). Structural features of protein-nucleic acid recognition sites. Biochemistry 38, 1999-2017. Nagadoi, A., Nakazawa, K., Uda, H., Oknno, K., Maekawa, T., Ishii, S., and Nishimura, Y. (1999). Solution structure of the transactivation domain of ATF-2 comprising a zinc finger-like subdomain and a flexible subdomain. J.Mol.Biol. 287, 593-607. O'Neil, K., and DeGrado, W. (1990). A thermodynamic scale for the helix-forming tendencies of the commonly occurring amino acids. 250, 646-651. Ohme-Takagi, M., and Shinshi, H. (1995). Ethylene-inducible DNA binding proteins that interact with an ethylene-responsive element. Plant Cell 7, 173-182. Overington, J., Donnelly, D., Johnson, M.S., Sali, A., and Blundell, T.L. (1992). Environment-specific amino acid substitution tables: Tertiary templates and prediction of protein folds. Protein Sci 1, 216-226. 122 Poon, I.K.H., and Jans, D.A. (2005). Regulation of nuclear transport: central role in development and transformation? Traffic 6, 173-186. Prieve, M.G., Guttridge, K.L., Mnnguia, J., and Waterman, ML. (1998). Differential importin-alpha recognition and nuclear transport by nuclear localization signals within the High-Mobility-Group DNA binding domains of Lymphoid Enhancer Factor 1 and T-Cell Factor 1. Mol. Cell. Biol. 18, 4819-4832. Rashotte, A.M., Mason, M.G., Hutchison, C.E., Femira, F.J., Schaller, G.E., and Kieber, J.J. (2006). A subset of Arabidopsis AP2 transcription factors mediates cytokinin responses in concert with a two-component pathway. PNAS 103, 11081-11085. Ren, J., Esnouf, R., German, E., Somers, D., Ross, C., Kirby, 1., Keeling, J., Darby, G., Jones, Y., and Stuart, D. (1995). High resolution structures of HIV -1 RT from four RT-inhibitor complexes. Nat Struct Biol. 2, 293-302. Riechmann, J.L., and Meyerowitz, EM. (1998). The AP2/EREBP family of plant transcription factors. Biol Chem 379, 633-646. Rivas, S., Romeis, T., and Jones, J.D.G. (2002). The Cf-9 disease resistance protein is present in an ~420—Kilodalton heteromultimeric membrane-associated complex at one molecule per complex. Plant Cell 14, 689-702. Rost, B., and Liu, J. (2003). The PredictProtein server. Nucl. Acids Res. 31, 3300-3304. Ryter, J.M., and Schultz, S.C. (1998). Molecular basis of double-stranded RNA-protein interactions: structure of a dsRNA-binding domain complexed with dsRNA. Embo J 17, 7505-7513. Sakuma, Y., Lin, 0., Dubouzet, J.G., Abe, H., Shinozaki, K., and Yamaguchi- Shinozaki, K. (2002). DNA-binding specificity of the ERF/AP2 domain of Arabidopsis DREBs, transcription factors involved in dehydration- and cold- inducible gene expression. Biochem Biophys Res Commun 290, 998-1009. Sakuma,Y., Maruyama,K.,Osahbe,Y.,Qin,F.,Seki,M.,Shinozaki,K.,and Yamaguchi—Shinozaki, K. (2006). Functional analysis of an Arabidopsis transcription factor, DREBZA, involved in drought-responsive gene expression. Plant Cell 18, 1292-1309. Sambrook, J., Fritsch, E., and Maniatis, T. (1989). (Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press). Schena, M., Pieard, D., and Yamamoto, K. (1991). Vectors for constitutive and inducible gene expression in yeast. Methods Enzymology 194. Silver, P.A., Keegan, L.P., and Ptashne, M. (1984). Amino terminus of the yeast GAL4 gene product is sufficient for nuclear localization. PNAS 81, 5951-5955. 123 Stockinger, E.J., Gilmour, S.J., and Thomashow, M.F. (1997). Arabidopsis thaliana CBF1 encodes an AP2 domain-containing transcriptional activator that binds to the C—repeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in response to low temperature and water deficit. PNAS 94, 1035- 1040. Sudbeck, P., and Scherer, G. (1997). Two independent nuclear localization signals are present in the DNA-binding High-mobility Group domains of SRY and SOX9. J. Biol. Chem. 272, 27848-27852. Tang, Y., Ramakrishnan, C., Thomas, J., and Deme DB. (1997). A role for HDJ-2/HSDJ in correcting subnuclear trafficking, transactivation, and transrepression defects of a glucocorticoid receptor zinc finger mutant. Mol Biol Cell 8, 795-809. Tansey, W.P. (2001). Transcriptional activation: risky business. Genes Dev. 15, 1045- 1050. Vlaclronasios, K.E., Thomashow, M.F., and Triezenberg, SJ. (2003). Disruption mutations of ADA2b and GCN5 transcriptional adaptor genes dramatically affect Arabidopsis growth, development, and gene expression. Plant Cell 15, 626-638. Vogel, J.T., Zarka, D.G., Van Buskirk, H.A., Fowler, S.G., and Thomashow, M.F. (2005). Roles of the CBF2 and ZAT12 transcription factors in configuring the low temperature transcriptome of Arabidopsis. The Plant Journal 41, 195-211. von Arnim, A.G., and Deng, X.-W. (1994). Light inactivation of Arabidopsis photomorphogenic repressor COPl involves a cell-specific regulation of its nucleocytoplasmic partitioning. Cell 79, 1035-1045. Walther, R.F., Atlas, E., Carrigan, A., Rouleau, Y., Edgeeombe, A., Visentin, L., Lamprecht, C., Addielrs, G.C., Hache, R.J.G., and Lefebvre, Y.A. (2005). A Serine/I‘hreonine-rich motif is one of three nuclear localization signals that determine unidirectional transport ofthe mineralocorticoid receptor to the nucleus. J. Biol. Chem. 280, 17549-17561. Wang, W., and Malcolm, B. (1999). Two-Stage PCR protocol allowing introduction of multiple mutations, deletions and insertions using QuikChange site-directed mutagenesis. BioTechniques 26, 680-682. Wang, Z., Triezenberg, S., Thomashow, M., and EJ., S. (2005). Multiple hydrophobic motifs in Arabidopsis CBF1 COOH-terminus provide frmctional redundancy in trans-activation. Plant Mol Biol 58, 543-559. Wojciak, J., Sarkar, D., Landy, A., and Clubb, R. (2002). Arm-site binding by lambda -integrase: solution structure and functional characterization of rts amino-terminal domain. PNAS 99, 3434-3439. 124 Wojciak, J.M., Connolly, K.M., and Clubb, R.T. (1999).‘NMR structm'e of the Tn916 integrase-DNA complex 6, 366-373. Wolberger, C. (1996). Homeodomain interactions. Current Opinion in Structural Biology. 6, 62-68. Zhang, R.-g., Kim, Y., Skarina, T., Beasley, S., Laskowski, R., Arrowsmith, C., Edwards, A., Joachimiak, A., and Savchenko, A. (2002). Crystal structure of Themotoga maritima 0065, a member of the IclR transcriptional factor family. J. Biol. Chem. 277, 19183-19190. Zhang, X., Basso, W., and Matthews, B. (1991). Toward a simplification of the protein folding problem: a stabilizing polyalanine alpha-helix engineered in T4 lysozyme. Biochemistry 30, 2012-2017. 125 lllllllllllllllllllllll LIBRAFilES Willi/lllill/fl!!!W!!!1fill/Illll/lI/MW 31293 02956 4006