A Necessary Good: Nuclear Hormone Receptors and Their Chromatin Templates

Fyodor D. Urnov and Alan P. Wolffe

Sangamo Biosciences Point Richmond Technical Center Richmond, California 94804


    INTRODUCTION
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
The eukaryotic genome is exquisitely responsive to a great variety of internal, developmental, and environmental stimuli and yet is most commonly visualized in packaged form: each 180 bp of DNA are bound to histones into a nucleosome, chains of nucleosomes coalesce into a necklace, and the necklace twists into a chromosome of undetermined structure that is somewhat inert in its regularity. This seeming conundrum is easily resolved in vivo, because, rather than treat the histones and chromatin as some pernicious obstacle that ought to be swept out of the way, eukaryotic transcriptional regulators clearly see a powerful ally in the nucleoprotein architecture of chromatin and make extensive use of its disruption, modification, and assembly to control the genome’s behavior. In this review, we focus on a well researched example of such a symbiosis: the intimate functional relationship that nuclear hormone receptors (NHRs) have with their chromatin templates. Its history in biological scholarship is a distinguished one and includes the pioneering study by Klever and Carlson (1) in the 1950s on the insect steroid hormone ecdysone that revealed a functional connection between hormone action, gene activation, and remodeling of chromatin, studies in the 1980s on the glucocorticoid receptor that took such analysis to a molecular level (2), and subsequent work in the 1990s from a number of laboratories that characterized the extraordinarily diverse spectrum of chromatin-modifying and -remodeling factors that partner with the NHRs to effect gene regulation (3).

In this review, we address some textbook schematic-inspired misconceptions about the structure of the nucleosome and of chromatin that are relevant to transcriptional control and elaborate on the structural impediments created by chromatin for access to DNA by nonhistone factors. We highlight mechanisms whereby such NHRs as the glucocorticoid, estrogen, and thyroid hormone receptors overcome this impediment and present a hypothetical scenario for gene activation by NHRs in vivo. Various types of chromatin disruption and modification phenomena that occur in vivo and the chromatin-modifying machines thought to be responsible for such modifications are described. We list the evidence implicating such machines in NHR function and conclude by reviewing recent data that illuminate the remarkably dynamic nature of in vivo gene regulation by NHRs.


    THE MISREPRESENTED NUCLEOSOME
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
The elementary particle of chromatin—the union of 8 molecules of core histone, 1 molecule of linker histone, and approximately 160 bp of DNA—is often represented as a featureless spool, the DNA’s thin thread winding around its surface. The inevitable oversimplification inherent in such textbook-style schematics creates three common misconceptions.

1. From a structural standpoint, the histone octamer is not a spool—a more adequate analogy would be to imagine DNA wound around a Rubik’s cube: a dynamic, pliable, internally nonhomogeneous particle (4). All four core histones, H2A, H2B, H3, and H4, contain a distinctive lysine- and arginine-rich COOH domain that forms a histone fold motif (an extended central {alpha}-helix flanked on either side by a loop and two shorter {alpha}-helices). Histones H3 and H4 heterodimerize via a handshake interaction of their histone fold motifs, and this dimer then associates with a second such entity to form an H3/H4 tetramer. In vivo, this particle is deposited onto nascent DNA by dedicated molecular chaperones (5, 6) and binds to approximately 130 bp of DNA. The resulting seminucleosome has interesting properties (7) immediately relevant to transcriptional regulation by NHRs (see below) and normally exists only on newly replicated DNA (i.e. when proliferating cells are in S phase). Histones H2A and H2B are added, by a distinct set of chaperones, to the DNA-bound (H3/H4)2, thus completing the assembly process and compacting approximately 146 bp of DNA (in the nucleosome, H2A and H2B also interact via a handshake). The x-ray crystal structure of the core histone octamer has been solved (8) and reveals the extended network of electrostatic and, surprisingly, hydrophobic interactions between the DNA and the octamer.

2. The junction between (H3/H4)2 and (H2A/H2B)2 is a delicate one, a feature most likely exploited not only in assembly (6), but also, possibly, in the disruption of chromatin by regulators (9). The consequences of such disruptions are not immediately apparent from thread on a spool drawings because they obscure the extraordinary structural stress that B-form DNA undergoes when assembled onto the histone octamer. The 1.7 turns made by DNA in a nucleosome feature the sharpest curvature of the DNA backbone that is thermodynamically feasible, coupled with an underwinding of the double helix (10); thus, the DNA is not complacently wound on the spool—it is extensively bent and twisted out of its familiar B-form shape. Thus, while the union of DNA and core histones in a structurally intact nucleosome is a stable one, it is easy to conceptualize how disruptions or modifications of the histone octamer that destabilize the protein-DNA contacts offer the DNA a window of opportunity to spring back into a less torsionally stressful conformation, a phenomenon with regulatory consequences (see below).

3. Limitations of experimental technique as well as an understandable desire to present simplified drawings routinely omit from the nucleosome one of its least understood and most important structural features: the NH2-terminal tails of the core histones. External to the histone-fold domain that lies inside the DNA superhelix, the histone tails account for approximately 25% of the overall protein mass of the nucleosome and are exceedingly lysine rich. While known to reside outside of the core nucleosome, they are, unfortunately, invisible to current x-ray crystallographic analysis (8). Thus, while nothing is known about the tails’ secondary structure, two things are certain: the tails are quite long and in fully extended form could, potentially, project far beyond the core particle (11); in addition, they contain a total of 44 lysine residues per nucleosome (the {epsilon}-NH2 group of lysine has a pKa of 10.8, i.e. at physiological pH, the average nucleosome is surrounded by 43.99 charged lysine side-chains). In a tail-centric view of the nucleosome, therefore, it resembles an octopus, with eight long tentacles/tails enveloping the DNA wound around its body in an web of positive charge.

As discovered in the 1960s (12), however, this charge can be neutralized via acetylation of the lysine residue: R-(CH2)3-NH3+ + AcCoA -> R-(CH2)3-NH-CO-CH3 Some 40 yr later, we know that every eukaryotic taxon that has been studied contains a vast number of enzymes, histone acetyltransferases (HATs), capable of performing this reaction, and have extensive data connecting such enzymes—and their functional antagonists, histone deacetylases (HDACs)—to every aspect of genomic function (13), including transcriptional control effected by NHRs (3). The invisible tails, therefore, are quite functionally conspicuous, as discussed in detail below.

Before describing how NHRs make use of chromatin modification and disruption to effect gene control, it is useful to briefly consider how these regulators are able to interact with their target loci in the context of mature (i.e. unremodeled) chromatin.


    ACCESS GRANTED: HOW NHRs ASSOCIATE WITH CHROMATIN
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
The twisted necklace drawings of chromatin illustrate an important problem: if mature, inactive chromatin is, in some sense, a tightly packed and wound-up nucleoprotein entity, how can a nonhistone factor gain access to its response element? While no data are available as to the interaction of NHRs with higher-order chromatin structures, we know that NHRs can bind to their response elements directly even when they are assembled into nucleosomes.

Two technical terms are commonly used to describe the exact way in which a particular DNA stretch is assembled into nucleosomes. The core histone octamer forms the most stable contacts with approximately 146 bp of DNA; the addition of linker histone compacts an additional approximately 15 bp, and about 20 bp of DNA are free (not associated with core histones). This internucleosomal gap (also known as "linker DNA") creates an inherent aperiodicity in the way the genome is bound into chromatin: some sequences are associated with core histones, and others are located in the linker. The term "translational positioning" is used to describe where exactly the gaps are in a particular locus. For example, the following 520-bp segment stretch of DNA (each letter corresponds to 20 bp) ABCDEFGHIJKLMNOPQRSTUVWXYZ can form nucleosomes in many different translational frames (DNA bound to histones is double underlined, while DNA in the linker is in boldface):

It is important to note that while some loci in the genome are associated with a specific translational frame of histone octamer occupancy, on many loci nucleosomes assume relatively random translational positions (14, 15). For our hypothetical alphabet locus, however, it is clear that frame 1 is a welcoming environment for a regulator seeking to bind sites B and K, but frames 2 and 3 are not (unless, of course, the nucleosome is no obstacle to the regulator; see below). There is direct evidence that a significant number of transcriptional regulators that operate in eukaryotic genomes cannot associate with their binding sites in the context of a mature nucleosome: such, for instance, is the case for TATA box-binding protein (TBP) (16), where the requirement for a significant structural distortion in the width of the DNA minor groove effected by TBP binding (17) is incompatible with the structure of the DNA in the nucleosome. A less severe impairment is presented by the well studied Zn-finger regulator TFIIIA (18, 19): it cannot bind to its site if the translational frame puts it in the center of a nucleosome (e.g. site J in frame 2 above), can stably bind if the DNA is in a linker (site J, frame 3) and can bind with reduced affinity if the DNA is on the edge of the nucleosome (site J, frame 1).

Even if the translational frame is roughly the same on two different chromosomes, the exact way in which the DNA is wound around the histones also has major consequences for the capacity of nonhistone regulators to access their sites. Consider the following fragment of the Xenopus laevis TRßA gene promoter (20):

The underlined portion of this DNA is a classical thyroid hormone receptor response element (TRE): two AGGA/TCA half-sites (boldface) spaced by 4 bp and arranged in the form of a direct repeat, with the left half-site bound by the retinoid X receptor (RXR), and the right half-site bound by the thyroid hormone receptor (TR)(21). The inevitable consequence of winding DNA around a histone octamer is that the major groove side of certain sequences will face toward the histones, and some will face toward the solution, thus yielding a particular rotational frame, i.e. a specific orientation of sequences in the double helix relative to the octamer. Here, for example, is the TRßA sequence in three distinct rotational frames (sequences located in a major groove rotated toward solution are underlined, while sequences located in a major groove that is rotated toward the histones are not; the TRE half-sites are in boldface):

A remarkable property of many NHRs is their ability to stably bind nucleosomal DNA in vitro and in vivo in a manner largely irrespective of translational frame; thus, quite unlike TBP or TFIIIA, NHRs experience only a modest reduction in affinity for their response elements when they are associated with core histones. For example, glucocorticoid receptor (GR) can associate with its response element on the mouse mammary tumor virus long terminal repeat (MMTV LTR) presented in multiple translational frames in vitro and in vivo (22, 23, 24); an important point, however, is that the rotational frame has to be such that the core residues required for GR binding to DNA are facing toward solution (25); for instance, frames 1 and 3 above expose the response element to the receptor, while frame 2 rotates it toward the histones.

When a favorable rotational frame exists, an association of the NHR with its response element both in vitro and in vivo has been demonstrated for GR action on the MMTV LTR (22), for TR on the TRßA promoter (26, 27), for estrogen receptor (ER) on the pS2 promoter (28), and for retinoic acid receptor (RAR) on the RARß promoter (29, 30). It is thought that the Zn-finger DNA-binding domain of one NHR molecule invades the major groove (31) that lies exposed on the surface of the octamer, and the second molecule of the receptor in the dimer tolerates some inevitable distortion of base-protein contacts both in the case of homodimers for class I NHRs (31) and heterodimers in the case of class II NHRs (21). It is not surprising, therefore, that the known structural anisotropy of the DNA within the nucleosome (10) leads to distinct effects of altering rotational phasing of GR response elements on GR binding when the location of the response element is shifted within the nucleosome (25); it is quite likely that the two neighboring double helices of DNA that come into proximity over segments of the nucleosome also contribute to such effects via steric hindrance.

The past 2 yr have seen the recapitulation of transcriptional control by GR (32), progesterone receptor (PR) (33), ER (34), vitamin D receptor (VDR) (35), and RAR (36) in various in vitro systems with purified receptor on chromatin templates. While these systems are not fully biochemically defined— Drosophila embryo extract is used to assemble chromatin, and HeLa nuclear extract is added to endow the system with transcriptional competency—they represent an exciting and important advance in the field and will undoubtedly allow a fine-resolution dissection of receptor action on chromatin (see below), including the initial step of receptor interaction with a nucleosomal template.


    THE HOUSE THAT NHRs BUILD
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
The average NHR complex bound to DNA, e.g. the TR/RXR heterodimer, has a molecular mass of approximately 100 kDa, comparable to that of a core histone octamer; a common NHR ligand, e.g. thyroid hormone, is tiny (molecular mass = 651 Da). Such slight agents, nevertheless, command an army of truly epic proportions, as most of the coactivator and corepressor complexes implicated in NHR action dwarf the receptors that target them: for instance, a complex containing the corepressor SMRT (silencing mediator for retinoid and thyroid hormone receptors) (37), the chromatin remodeling complex SWI/SNF (38), and the related coactivator complexes TRAP and DRIP (35, 39) are all well over 1 MDa in molecular mass. These complexes, together with the DNA-bound receptor, the basal transcription factors (40), and the RNA polymerase II holoenzyme (41), assemble at an NHR-regulator promoter and form an entity that exceeds the ribosome in size and complexity.

It is unlikely that a general scenario can be envisaged that will account for the behavior of all NHR-regulated genes. A partial synthesis, however, would be helpful before a discussion of the major partners of the NHRs in transcriptional control. The model presented here is based on biochemical studies from a large number of laboratories (3, 42) and from structural and functional studies on the budding yeast HO endonuclease gene (43, 44), the MMTV LTR (2, 32, 45, 46, 47), the mouse serum albumin gene (48, 49, 50), and the Xenopus TRßA gene (27, 51).

1. The receptor accesses its binding site in the chromatinized promoter either due to an existing favorable rotational frame (see above), due to intrinsic nucleosome mobility relative to the DNA that stochastically presents such a frame to the receptor (52), due to the action of a factor that enforces the maintenance of a particular rotational frame within a given locus (49), or in the aftermath of DNA replication when chromatin is incompletely assembled and the nascent DNA is in an accessible conformation (5, 6, 53).

2. Certain chromatin-bound class II NHRs, when in the unliganded state, target corepressor complexes that effect transcriptional repression through deacetylation of chromatin (42); unliganded TR also exploits an auxiliary repression pathway with an undetermined mechanistic foundation that is not HDAC dependent (54).

3. Upon addition of ligand, the NHR recruits an ATP-dependent chromatin remodeling complex that actively reorganizes histone-DNA contacts in the vicinity of the receptor binding site. On a macroscopic level, such targeting leads to the generation of a stretch of chromatin ("DNAse I hypersensitive site") in which the DNA is significantly more accessible to non-histone-regulatory factors than DNA in mature chromatin.

4. The liganded receptor shuttles on and off the DNA (47) while the remodeled state persists, in part, perhaps, due to the binding of additional regulatory factors within the DNAse I-hypersensitive site that exert stimulatory effects on the basal transcriptional machinery. The liganded receptor, as well as these additional nonreceptor factors, both target complexes that possess HAT activity; this recruitment stimulates transcription.

5. In addition to HAT-containing complexes, NHRs also target large coactivator complexes that engage the basal transcriptional machinery (35, 39) and are also involved in transcriptional activation by non-NHR regulators (55).

We will presently review recent progress in our understanding of some of these steps of gene regulation by various NHRs.


    WHAT IS WITHOUT Sin3: REPRESSION REVISITED
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
In contrast to class I NHRs (e.g. GR and PR) that are sequestered in the cytoplasm in the absence of ligand, some class II NHRs, including RAR and TR, are largely nuclear; ligand is known to have some effect on the intracellular distribution of TR (56). A considerable fraction of cellular TR is a constitutive component of chromatin (57), and it is possible that the same holds for RAR, although questions exists as to the stable occupancy by RAR of target gene promoters in the absence of ligand (29).

A physiological rationalization of such receptor behavior lies in the known biological function of thyroid hormone (58, 59) and various retinoids (60) as inducers of cell cycle arrest and differentiation. It is thought that unliganded TR/RXR and RAR/RXR silence genes that are required for execution of this program, and that ligand causes a relief of transcriptional repression and a concomitant drastic change in cell phenotype; interestingly, vitamin D and its receptor, VDR, have a well established analogous role in such processes as osteogenesis (61), but unliganded VDR is not known to act as a transcriptional repressor. An attractive evolutionary parallel exists with the major NHR in insects, the ecdysone receptor, that effects an analogous function in insect metamorphosis (62). The target genes in vertebrates are largely unknown, with a few exceptions (63), but evidence from studies of mutated versions of TR and RAR found in various leukemias are fully consistent with this working model (64, 65).

Unliganded TR and RAR act as potent transcriptional repressors on model promoters (65), and studies with in vitro systems demonstrated that such action could occur on naked DNA templates (66): for instance, unliganded TR efficiently interferes with preinitiation complex formation (67). An alternate or complementary pathway for repression by TR and RAR was proposed when yeast two-hybrid assays identified two large, related polypeptides termed N-CoR (nuclear receptor corepressor) (68, 69) and SMRT (70). Their mechanism of action remained obscure until work from several laboratories (71, 72) offered the provocative finding that the transcriptional repressor Mad/Max associates with the histone deacetylase RPD3 (also called HDAC1) via the adapter molecule Sin3. This observation provided a direct connection between targeted transcriptional repression and the modification of chromatin.

It is useful to note that a correlative relationship between levels of acetylation and transcriptional activity of specific loci had by then been well established: for instance, active domains of chromatin were known to be hyperacetylated (73, 74, 75), while the inactive X in human females was known to be hypoacetylated (76). Thus, it was tempting to speculate that the targeted deacetylation of chromatin could contribute to transcriptional repression in mammals; evidence that N-CoR (77) and SMRT (78) could interact with Sin3, and that Sin3 could associate with HDAC1/RPD3 led to a model in which a hypothetical complex of the corepressor with Sin3-HDAC1 was recruited by the unliganded NHR to specific loci in the genome.

It remained unclear, however, what fraction of cellular N-CoR was associated in vivo with Sin3A, whether the HDAC activity targeted by N-CoR was due to HDAC1/RPD3 or some other HDAC, or whether the hypothetical N-CoR-Sin3-HDAC1 complex can associate in vivo or in vitro with unliganded NHRs (77). A similar state of uncertainty existed with regard to transcriptional repression effected by SMRT; most significantly, it was unknown whether endogenous SMRT associates with Sin3 or HDAC in vivo (78). It also remained to be demonstrated that unliganded NHRs recruit HDAC activity to target promoters; for instance, while the HDAC inhibitor TSA was shown to up-regulate transcriptional activity of a promoter repressed by unliganded RAR/RXR (78), this up-regulation was equally robust in the presence of all-trans retinoic acid (78), i.e. after the receptor released both the corepressor and the attending hypothetical HDAC. These findings were consistent with the notion that TSA effects on transcription were unrelated to the hypothetical HDAC targeted by RAR/RXR (78). To further complicate matters, stable occupancy of target promoters by unliganded RAR/RXR heterodimers has been very difficult to detect in vivo (79); while explanations for such action in absentia are currently lacking, hit-and-run targeting of corepressor action via transient occupancy of target promoters by RAR may be invoked.

Recent data (37, 54) directly argue against the involvement of a hypothetical (SMRT/N-CoR)-Sin3-HDAC1/RPD3 complex in transcriptional repression by unliganded NHRs. For instance, the application of two different biochemical purification strategies to characterize an endogenous SMRT-containing complex in HeLa cells yielded the identical observation that Sin3 or HDAC1 fail to associate with this corepressor (37). The 2% of endogenous SMRT and N-CoR that are not contained in this complex are not associated with Sin3-HDAC1, either (80). Similarly, targeting of N-CoR and HDAC activity to unliganded TR and its mutated derivative, the oncoprotein v-ErbA, occurs without an association with Sin3 and HDAC1 (54). Thus, at the present time, the existence of an in vivo complex between Sin3-HDAC1 and any NHR corepressors or a role of this hypothetical entity in NHR function both await experimental support.

What, then, are the corepressors’ partners, and are any of them HDACs? Recent biochemical analysis (37, 81) describes a large (~ 2 mDa) SMRT-containing complex in HeLa cells. Remarkably robust [it is reported to remain stable in 0.5 M NaCl (37)], this entity contains multiple polypeptides, including histone deacetylase 3 (HDAC3). An association between HDAC3 and N-CoR/SMRT was also discovered in Xenopus oocytes (54, 81) as well as in a distinct biochemical purification scheme applied to HeLa cells (82). Several HDACs other than HDAC1 have been biochemically connected to N-CoR and SMRT (80, 83), but no evidence is available as to whether these associations are relevant to NHR function.

In contrast, the novel Sin-free (SMRT/N-CoR)-HDAC3 complex has been implicated in repression by NHRs by several lines of evidence. In human cells, one of its integral components, the histone-binding protein TBL1 (see below), potentiates transcriptional repression by Gal4-TRLBD when overexpressed (37). Importantly, an allelic version of TR that fails to recruit corepressors (84) also fails to respond to TBL1 overexpression (37). In Xenopus, unliganded TR associates with N-CoR-HDAC3, and treatment with thyroid hormone both relieves transcriptional repression effected by TR and eliminates any detectable association between the receptor and N-CoR or HDAC3 (54). Finally, microinjection of purified antibodies directed against HDAC3 or against N-CoR partly alleviates repression effected by unliganded TR in Xenopus oocytes (81); importantly, such treatment failed to affect basal transcription driven by the reporter construct used, directly linking antibody action to receptor-dependent phenomena.

Any speculation on the significance of targeting of HDAC3, rather than HDAC1, to promoters repressed by unliganded NHRs is complicated by the paucity of information on functional distinctions between these HDACs; their numbers are growing faster than our insight into the selective advantage reaped by the cell from possessing multiple HDACs. In interesting contrast to the HATs, which exhibit a wide spectrum of histone tail lysine specificity for action (see below), HDACs target all lysines with seemingly equal efficiency. The choice by unliganded TR of HDAC3 in particular is an interesting one, however, because this enzyme was reported to be more efficient than its sister protein HDAC1/RPD3 in deacetylating histones assembled into nucleosomes (85). Such capacity may have functional utility for TR when it associates with nascent DNA in the aftermath of DNA replication fork passage [a common occurrence, considering that unliganded TR maintains a variety of cell types in the proliferative state (64, 86)]. On such newly replicated loci, chromatin is gradually assembled de novo by molecular chaperones that deposit histones onto DNA in hyperacetylated form. The newborn hyperacetylated nucleosomes are converted into a mature, deacetylated state over about 30 min by unknown HDACs; on loci bound by unliganded TR, however, such deacetylation—and attending transcriptional repression—could be accelerated by N-CoR-targeted HDAC3.

Several lines of evidence indicate, however, that transcriptional repression by unliganded TR is mediated via both an HDAC-dependent as well as an -independent pathway (54); other eukaryotic transcriptional repressors, for instance, Ikaros (87, 88), display similar properties. An important challenge for future experiments will be to provide a currently lacking experimental connection between in vitro data that TR can directly interfere with basal transcription machinery function (67) and in vivo behavior of the receptor. An additional significant, and currently unaddressed, gap in the literature is a lack of evidence that transcriptional repression by NHRs leads to histone tail deacetylation over target promoters; at present, such data are only available for transcriptional repressors that operate in budding yeast (89, 90). While chromatin immunoprecipitation (91) is an exceptionally powerful and useful technique that has yielded profound insights on various aspects of genomic function in budding yeast (43, 92, 93, 94), it is unclear whether the size of mammalian genomes and the relatively small changes in levels of histone acetylation over target promoters are likely to collude in preventing the acquisition of a robust and interpretable dataset on this issue.

A final possibility that deserves investigation is that unliganded TR and RAR drive transcriptional repression via pathways not previously considered; it is possible, for instance, that some structural feature of the DNA-bound unliganded NHR leads to the relocalization of the entire locus into a transcriptionally inactive nuclear compartment (95). While an earlier hypothesis (96) that unliganded TR targets a ubiquitous chromatin remodeling machine termed Mi-2/NURD (97) lacks both biochemical (54) and functional (96) support, recent work offered a novel, and unexpected, connection between unliganded TR and repression-related chromatin assembly, when a histone-binding protein called transducin ß-like 1 (TBL1) was identified as a stable component of a large SMRT- or N-CoR-containing complex (37, 81). This observation is exciting for two reasons. First, mutations in TBL1 have been implicated in late-onset sensorineural deafness in human patients (98), and strong genetic evidence implicates TRß in auditory cortex development in mice (99). Second, from a mechanistic standpoint, TBL1 has the intriguing capacity to bind histone H3; a distant relative of human TBL1, the budding yeast transcriptional repressor Tup1p, exerts its function, in part, via directing the assembly of a repressive nucleosomal array (100, 101). It is tempting to speculate, therefore, that when bound to target loci in proliferating cells, where chromatin is repeatedly destroyed by DNA replication, and spurious gene activation is therefore much more likely than in quiescent cells, unliganded TR sustains transcriptional repression by non-chromatin-based pathways (54), as well as promoting the assembly of repressive chromatin both by interacting with the histones via the TBL1 moiety of the corepressor complex (37) and accelerating their deacetylation via HDAC3 action (37, 54).


    LENDING A HELPING TAIL: HISTONES AND TRANSCRIPTIONAL ACTIVATION BY NHRs
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
A review of chromatin involvement in transcriptional activation by liganded NHRs written only a few years ago would have been a relatively simple matter: one would describe the association of liganded GR with the MMTV LTR and the generation of a DNAse I-hypersensitive site over the promoter (2, 46), the subsequent binding by non-receptor-regulatory factors such as NF-1, and hypothesize on the role that interactions with the basal transcription machinery play in regulating the activity of RNA polymerase II. In the summer of 2000, few things appear straightforward, and even the best-studied experimental system—the action of GR on the MMTV LTR—is also the most complicated and, in some sense, the least well understood.

This complexity results from two circumstances. Extraordinary progress in the biochemical characterization of transcriptional cofactors implicated enough of them in NHR signaling to take up a five-page table in a review published in this journal earlier this year (3). For students of chromatin, it is comforting to see that the eukaryotic nucleus is populated with chromatin-modifiying and -remodeling machines, but such abundance does not allow itself easily to being confined into parsimonious models. In particular, there continues to be a significant gap between our understanding of biochemical activities of such machines in vitro and our knowledge of in vivo structural and functional consequences of their presumed action. In addition, we have evidence that the actual assembly of target promoters into chromatin potentiates and attenuates the response to NHR action, but the mechanistic underpinnings of such a synergy between chromatin and the receptors are emerging only very slowly.

Recent reviews describe in detail the various types of chromatin disruption and modification that occur in vivo (11) and also the coactivators that abet NHR signaling by effecting such alterations in chromatin structure (3, 42, 102). We will discuss these issues briefly, focusing on specific examples of NHR action on chromatin templates.

It is helpful to make a distinction between the two major types of chromatin structure disruption that occur in vivo and are known to be associated with transcriptional activation effected by liganded NHRs. The first type is often called "chromatin remodeling" and refers to a dramatic, localized alteration in the fiber of chromatin in which a particular nucleosome, or several adjacent nucleosomes, undergo a receptor-controlled structural change (2). The end result is the creation of a stretch of DNA that is significantly more accessible to nucleases (e.g. DNAse I); hence, its technical name, "DNAse I hypersensitive site." It is quite likely, although a formal demonstration in vivo is currently lacking, that such remodeling effected by liganded NHRs occurs via the recruitment of large ATP-utilizing complex. In the case of GR, the best candidate is a multisubunit complex called SWI/SNF (pronounced "switch-sniff") (45, 102, 103). In budding yeast, powerful genetic data indicate that SWI/SNF controls gene expression via a chromatin-based pathway (104, 105), and chromatin remodeling at specific promoters is known to be SWI/SNF-dependent (106, 107, 108). In vertebrates, however, it is not known whether remodeling at any NHR-regulated promoter in vivo requires SWI/SNF. In purified form, the SWI/SNF complex has a variety of ATP-dependent activities, including the rearrangement of purified nucleosomes to potentiate regulatory factor access to DNA (16, 109, 110), and the induction of histone octamer sliding relative to the DNA (111). It is unclear how the remodeled and mobilized nucleosomes observed in vitro relate in structure to the entity that is revealed in vivo as a DNAse I-hypersensitive site; thus, while many NHRs are known to effect such remodeling (2, 112), their biochemical accomplices in such action remain obscure, a gap that well deserves to be addressed.

Biochemical analysis of eukaryotic nuclei yielded a large number of other ATP-dependent chromatin remodeling machines (38, 113), many of which contain a core catalytic subunit called "imitation switch" (ISWI); as of August, 2000, however, there is no evidence connecting these complexes to gene regulation in vivo in organisms other than budding yeast (and even in yeast, targeted action of any ISWI-containing complex at a specific promoter has not been demonstrated), although PR and ISWI have been reported to synergize in vitro (114).

Whatever the specific ATP-dependent complex that effects such remodeling in vivo, it is useful to note that such a nucleosome-scale disruption is not a formal prerequisite for gene activation and that specific promoters can undergo significant transcriptional up-regulation without concomitant remodeling (e.g. Refs. 112, 115); some hormonally responsive promoters, for example, the ER-regulated pS2 gene (28), do not undergo a macroscopic change in chromatin structure upon ligand-driven transcriptional activation. A further mystery involves the actual fate of the histone octamer during remodeling; while original models proposed that the octamer is physically evicted, subsequent in vitro studies showed that much of the DNA remains in some contact with the octamer (116) and that transcription factors and histones can share the same remodeled nucleosome (117). Analysis of chromatin structure in vivo suggested that specific remodeled loci can remain associated with histones (48), and similar observations were reported for RAR- and GR-regulated genes (22, 29). It is possible that the persistence of the octamer in the vicinity of the receptor-binding site is related to negative feedback mechanisms of attenuating transcriptional response to hormone (see below), since the receptor shuttles on and off the DNA (47), and since continued maintenance of DNA occupancy by a nonhistone factor is required to exclude histone octamers from its binding site (118).

To complicate matters somewhat beyond our current ability to put forward rational explanations, a meticulous high-resolution analysis of chromatin structure at the MMTV LTR led to the revision of two previously held concepts. Earlier studies indicated that this promoter is assembled into an array of translationally and rotationally positioned nucleosomes, one of which (nucleosome B) presents on its surface the glucocorticoid response element (GRE) to the liganded receptor. Subsequent mapping demonstrated that, in fact, chromatin in this region is assembled into multiple translational frames, but that some frames are more common than others (also known as "frequency-biased distribution"; see Fig. 1Go) (119); thus, previous mapping data indicated the existence of a very specific chromatin organization because the methods used generated data that averaged across a heterogeneous sample (a common complication). It is possible that in vivo, nucleosome mobility, whether intrinsic or facilitated (52, 118, 120), can lead to transition from one frame to another, and that the receptor then selects those templates that at a particular point in time are found in a correct rotational frame. Interestingly, a high-resolution mapping study on GR interaction with the MMTV LTR in Xenopus oocytes (121) demonstrated the ability of liganded GR to actively reorganize chromatin not only around its binding site, but elsewhere in the promoter; thus, GR can enforce a particular translational frame on the templates that it encounters, a property thought to be important in the regulation of mammalian genes by the winged-helix transcription factor HNF3 (48, 49, 50, 122). It will be of great interest to investigate whether such active reorganization is simply a matter of freezing chromatin in a particular frame via a bookend mechanism, or if liganded GR does, indeed, target a remodeling machine that can effect the active organization of chromatin into a particular frame.



View larger version (10K):
[in this window]
[in a new window]
 
Figure 1. An Array of Precisely Translationally Positioned Nucleosomes in the MMTV LTR Visualized by Low-Resolution Mapping Is Dissolved into a "Frequency-Biased Distribution of Multiple Frames" by High-Resolution Analysis

The top half shows the approximate position of histone octamers (ovals) relative to the transcription start site ("+1") of the MMTV LTR, the major GRE (solid bar), and the NF1 binding site. The lower half shows the positions of specific octamers as mapped by high resolution techniques. The thickness of the line designating each frame corresponds to its frequency; for some frames, the exact position is indicated in base pairs relative to the transcription start site. The array visualized on top results from superimposing signal derived from all the nucleosome positions actually observed in this DNA fragment (bottom). [This figure summarizes data in Ref. 22 and is based on Fig. 8 in that paper.]

 
Not only did the MMTV LTR lose its previously positioned nucleosome, we now know that the remodeling effected by liganded GR is not a single-nucleosome-based phenomenon (23, 32); while sequences surrounding the GRE do become more accessible to nucleases and transcription factors, the length of this accessible DNA stretch is greater than one nucleosome but shorter than two nucleosomes. Importantly, such a fractional transition occurs on individual DNA molecules and is not a consequence of superimposing signal from two remodeled nucleosomes that occupy different frames (23). While additional GREs located elsewhere in the LTR are thought to contribute to this phenomenon, current models are insufficient to explain the nature of the remodeling event; it is possible that some currently uncharacterized feature of internucleosomal association (8) is perturbed by the complexes being targeted by liganded GR (32).

The second type of chromatin structure disruption that occurs in vivo is commonly referred to as "chromatin modification": the core histones are the targets of many posttranslational covalent modifications (123), the most prominent of which is the acetylation of the lysines effected by HATs and reversed by HDACs (13, 124). For certain transcriptional activators in budding yeast, powerful data exist that formally prove a causative role between targeted histone acetylation and transcriptional activation: for instance, the HAT Gcn5p is required for activated transcription of specific genes, this requirement is dependent on the biochemical activity of Gcn5p as a HAT, Gcn5p action leads to the hyperacetylation of histones at target promoters, and the abolition of charged lysines in histone tails via their mutation to a noncharged amino acid relieves the requirement for Gcn5p in transcriptional activation (125, 126). It is important to note that the firm positive correlation between extent of acetylation and level of transcriptional activity does not always hold: for example, treatment with HDAC inhibitors has been, rather counterintuitively, shown to repress transcriptional activation driven by liganded GR on the MMTV LTR (127, 128), and genetic data from yeast and Drosophila also point to an unexpected role for HDAC in gene activation (129). In most cases, however, deacetylation correlates with repression, and hyperacetylation correlates with activation.

Mammalian genomes contain a large number of HATs (130), and a variety of biochemical evidence (e.g., two-hybrid interactions, GST pull-down assays with in vitro synthesized proteins, coimmunoprecipitations, etc.) connect many of these enzymes to signaling by liganded NHRs (3). In vivo evidence for a role of specific HATs in transcriptional activation effected by NHRs is, inevitably, less robust, with some exceptions. For example, the clinical syndrome of resistance to thyroid hormone in human patients (131) results from mutations in the ligand binding domain of TRß. Biochemical analysis indicated that many of these mutations impair the receptor’s capacity to target such HAT coactivators as steroid receptor coactivator-1 (SRC-1) (132, 133). In powerful support of these observations, genetic ablation of the gene for SRC-1 in mice was shown to lead to thyroid hormone resistance in the knockout animals (134). SRC-1 was shown to potentiate PR signaling in an in vitro system with chromatin templates, and such action was obviated by use of the HDAC inhibitor trichostatin A (33); it will be of great interest to determine whether the targeting of coactivator HATs by liganded NHRs in such in vitro systems leads to targeted hyperacetylation of chromatin on a scale similar to that effected by chimeric activators that target yeast coactivators to in vitro assembled chromatin (135). This will be particularly important because of a mysterious discrepancy between the known capacity of such NHR HAT coactivators as CBP/p300 (136) to efficiently hyperacetylate both histones H3 and H4 (137) with the observation that in vivo, certain NHR-regulated promoters undergo a change in the acetylation status of histone H4, while such changes in histone H3 acetylation are rather insignificant (138). It is possible that HATs other than CBP/p300 potentiate NHR signaling at those promoters, or that the acetylation status of histone H4 over those promoters is affected not only by NHR-targeted HATs, but other cell-wide effects of NHR ligands [for instance, H4 acetylation is expected to increase in proliferating cells (139)]. In addition, many HATs are known to have nonhistone targets (13), including transcription factors, other HATs, and components of the basal transcription machinery; it is quite possible that changes in acetylation status of such polypeptides also account for some of the transcriptional effects of liganded NHRs.

The wealth of biochemical evidence connecting the multitude of HATs to liganded NHRs is not, unfortunately, paralleled by insight of how hyperacetylation of chromatin leads to transcriptional activation; to some extent this is due to a lack of structural information about the NH2-terminal histone tails (8). In vitro data have indicated that hyperacetylation of chromatin facilitates transcription elongation by RNA polymerase (120), and recent evidence from budding yeast offers very strong support to this notion (140, 141). Since liganded NHRs are not known to move along their target genes with RNA polymerase, however, it is unlikely that the HATs they are presumed to target affect transcriptional elongation. Several other hypotheses have been put forward and confirmed through in vitro studies, most prominently, that a hyperacetylated nucleosome is more accessible to binding by nonhistone transcriptional regulators (19, 142). It has been difficult until very recently to investigate this issue in vivo, in part because NHRs not only target HATs, but also ATP-remodeling machines; the latter also promote accessibility of DNA to nonhistone regulators, which makes a dissection of the two pathways’ contribution to overall accessibility very difficult.

Recent data concerning ER action in vitro and in vivo on the pS2 promoter (28) provide a novel line of support for the hypothesis of acetylation-driven "facilitated access." Robustly responsive to estradiol, this promoter contains an ER binding site that is located at the edge of a translationally positioned nucleosome, and a TATA box bound at the edge of another positioned nucleosome approximately 380 bp downstream (i.e. exactly two nucleosome lengths away; see Fig. 2Go); the stable and unique translational frame of these two nucleosomes is likely due to intrinsic positioning properties of the DNA sequence (28). Liganded ER was shown to directly engage its ERE on the surface of the octamer and, without mediating a significant alteration in histone-DNA contacts, to stimulate transcription of the pS2 gene (28). Considering that large-scale remodeling similar to that seen over the MMTV LTR does not occur on this promoter, an important question was the mechanism of such stimulation. Chromatin immunoprecipitation assays demonstrated that treatment with estradiol led to hyperacetylation of histones H3 and H4 within both nucleosomes in vivo; concomitant with this chromatin modification, TBP binding was observed over the TATA element (143). Importantly, and in full agreement with earlier observations (16), while TBP could not access its binding site in the promoter when it was assembled in vitro into mature, deacetylated chromatin (143), an identical in vitro binding assay performed with hyperacetylated chromatin showed robust binding of TBP to the pS2 TATA box on the surface of a hyperacetylated, translationally positioned nucleosome (143). Such potentiation of binding occurred on a highly purified chromatin template, thereby excluding the possibility that contaminating nonhistone components are responsible for enhanced TBP access (143). These data extend earlier in vitro observations (19, 142) and offer a comforting parallel to data obtained in budding yeast (144, 145), according to which transcriptional activators increase TBP occupancy of target promoters in vivo.



View larger version (16K):
[in this window]
[in a new window]
 
Figure 2. Three Examples of Specific Chromatin Architecture Associated with NHR-Regulated Promoters

On the vitellogenin B1 promoter, a static loop formed by a translationally positioned nucleosome facilitates interactions between DNA-bound ER and the basal transcriptional machinery (146 ). On the pS2 promoter, a translationally positioned nucleosome excludes TBP from the TATA box, until ER-targeted histone hyperacetylation enables TBP binding (28 143 ). On the Xenopus TRßA promoter, unliganded TR binds to four sites spread over seven positioned nucleosomes; the TREs are located both in linker DNA and on the surface of nucleosomes, and chromatin assembly potentiates TR occupancy of the promoter (27, 51).

 
An interesting convergence of various chromatin disruption pathways was recently observed in vitro, when glycerol gradient purification of chromatin templates engaged by the chimeric activator Gal4-VP16 in the presence of ATP, highly purified chromatin remodeling complexes, the HAT p300, and histone chaperones yielded the highly unexpected observation that action by p300 promotes an ATP-dependent transfer of the H2A/H2B tetramer from the nucleosome to the histone chaperone (9). While it was not directly investigated, whether a bona fide H3/H4 tetramer remains on the DNA, this observation offers an interesting perspective on earlier studies of (H3/H4)2 complexed with DNA (7), according to which this entity is remarkably like naked DNA in terms of accessibility to nonhistone regulators and nucleases. Thus, it is possible that ATP-driven remodeling of histone-DNA contacts and hyperacetylation of histones driven by HATs may synergize at particular NHR-regulated promoters to generate a more severe state of chromatin disruption that can be achieved by either agent acting alone.

It is useful to appreciate, however, that none of the disruption, remodeling, or modification events (summarized in Table 1Go) that have been observed in vivo or in vitro during NHR action result in the generation of naked DNA. In fact, there are many well documented cases [e.g. the action of ER on the vitellogenin B1 (146) and the pS2 (28) promoters, of TR on the TRßA promoter (27, 51), GR on the MMTV LTR (147)] in which the assembly of target DNA into chromatin potentiates NHR function (Fig. 2Go); importantly, this potentiation occurs not only by expanding the range within which transcriptional levels vary, but also by endowing the system with regulatory properties that cannot be recapitulated on naked DNA (148). An excellent example is the regulation of the MMTV LTR by GR and other factors; on transiently transfected (i.e. not physiologically chromatinized) templates, such cellular factors as NF1 and Oct-1 can gain promiscuous access to the promoter (46). On endogenous chromatinized templates, however, such access, and the attending subtle alterations in promoter activity, are unequivocally dependent on GR (149, 150).


View this table:
[in this window]
[in a new window]
 
Table 1. NHRs and Chromatin: A Summary

 

    IN LIVING COLOR: NHRs IN ACTION
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
It is quite likely that the intranuclear environment endows transcriptional regulation by NHRs with functional properties that are not currently "dreamt of in our philosophy." As in other empirical fields, however, many hypotheses in the study of transcriptional control are based on a "presumption of common sense": phenomena that have not been investigated directly (due to limitations of experimental technique or other circumstances) are not shrouded with a veil of agnosticism, but rather are by default assumed to occur in a manner most consistent with common sense [in evolutionary biology, such scenarios are by necessity more ubiquitous and are referred to as "Just So Stories" (151) after R. Kipling’s famously inventive fairytales], until an actual experiment shows otherwise. Recent investigations of in vivo dynamics of GR in living cells offer an interesting illustration of the discrepancy between what common sense hypotheses maintained, and the actual state of affairs.

It is nontrivial to study the dynamics of the interaction of NHRs with their templates: a cell population large enough to generate experimental signal will be asynchronous with respect to transcriptional state of their template, and while it has been possible to partly circumvent this problem in elegant in vitro experiments (34), transcription in an intact mammalian cell does not lend itself to synchronization easily. In addition, most experimental techniques that investigate protein-DNA interactions (genomic footprinting with DNAse I, in vivo footprinting with DMS, chromatin immunoprecipitation, exonuclease III boundary measurements, etc.) all generate signal that represents an averaged sample of a potentially heterogenous population: for instance, the occupancy of >50% of target molecules by a nonhistone factor is sufficient to generate a genomic or in vivo footprint; similarly, accurately quantitating results from chromatin immunoprecipitation experiments to meaningfully estimate the percentage of templates bound by a protein at a given time is a nontrivial challenge that has yet to be adequately met.

In a technical tour de force, it has recently become possible to visualize, on a timescale of seconds, the dynamics of NHR-DNA interactions in chromatin of living cells (47). The significant technical obstacle of light microscope detection limits was overcome by creating a cell line that contains, in integrated form, approximately 200 reiterated copies of a 9-kb reporter construct in which the MMTV LTR is fused to the v-ras gene (152); this approach using lac repressor and operator sequences has been successfully used to track chromosome behavior in vivo (153, 154). A clone for a chimeric NHR—GR fused to green fluorescent protein (GFP)—was then introduced into the cells’ genome under the control of an inducible promoter; this experimental setup allowed precise control over the onset of NHR expression and, via the use of GR agonists, over the intracellular location of GR once it is expressed. Laser confocal microscopy in dexamethasone-treated cells revealed a large (~ 1–2 µM) intensely fluorescent structure in the nucleus, whose location corresponded precisely to the site of v-ras transcription as gauged by in situ hybridization.

The rate of exchange between GR bound to its response elements and GR elsewhere in the nucleus was then measured using two techniques that have been very successfully used to investigate the dynamics of membrane-based phenomena in cell biology: fluorescence recovery after photobleaching (FRAP) (155, 156, 157) and fluorescence loss in photobleaching (FLIP) (158). In FRAP, a laser beam is pointed at a defined location in the living nucleus (in this case, the 1.8 Mb GR-bound array) and thereby bleaches GFPs fluorescence for a considerable length of time. Recovery of fluorescence in the bleached spot can only occur if nonbleached molecules invade the target area from elsewhere in the nucleus; by definition, the rate of exchange is equal to the rate of recovery. In remarkable contradiction to common-sense notions of stable target site occupancy by NHRs, recovery of fluorescence occurred as rapidly as it could be measured, and in 5–10 sec following bleaching of all the GR molecules bound to the entire 1.8 Mbp array, these nonfluorescent NHR molecules were replaced by non-bleached GFP-GR (47). These data strongly suggest that liganded, DNA-bound GR undergoes extremely rapid "ejection" and reassociation with its DNA template in a hit-and-run mechanism of transcriptional activation (159). Equally powerful evidence to that effect was obtained using FLIP: GR located elsewhere in the nucleus was bleached, and subsequently, over approximately 60 sec, invaded the 1.8-Mb MMTV array.

This highly unexpected and convincing dataset sheds new light on observations earlier obtained by K. Nasmyth and colleagues (43) on the sequence of events that occur during transcriptional activation of the budding yeast HO endonuclease gene. Genetic observations indicated that, in full functional analogy to regulation by GR in mammals, the activation of this gene is dependent on a DNA-bound activator (Swi5p), the recruitment of a chromatin remodeling machine (SWI/SNF), and the subsequent targeting of a HAT complex (SAGA). Remarkably, chromatin immunoprecipitation assays in synchronized cells indicated that the activator, Swi5p, resides on the HO endonuclease promoter for a very brief period of time (~5–10 min) and disappears from the DNA at approximately the same time as it recruits SWI/SNF. Most significantly, SWI/SNF residency at the promoter then continues for at least 1 h; thus, epigenetic memory of activator occupancy persists long after it has been ejected.

What is the mechanism of this ejection and what functional benefit can it bring to the cell? No data exist to directly address either issue; by way of speculation, it is possible that ATP-driven remodeling action of SWI/SNF displaces from the DNA the very activator that recruits it, via direct action on the receptor-DNA contacts, or on the histones underlying the receptor-bound nucleosome. On the TRßA promoter in Xenopus, for instance, unliganded TR binds to multiple nonconsensus TREs (Fig. 2Go), and such binding is potentiated by chromatin assembly (27), perhaps though the structural distortion of chromatinized DNA that makes TREs more palatable to the receptor. Addition of ligand leads to the targeted disruption of chromatin around the receptor binding sites (27, 112) and is expected to therefore create a template that has less affinity for the receptor (i.e. lead to receptor ejection). Data obtained with an in vitro system that recapitulates regulation by purified GR on a chromatin template indicate that in the absence of ATP, GR can stably bind, but not remodel, chromatin over the promoter (32). The addition of ATP leads to chromatin remodeling coupled with an increased accessibility of DNA over the receptor binding sites (32), in full agreement with the hit-and-run model for GR function.

An obvious suggestion for the functional utility of such a drifter model of NHR action is in the attenuation of hormonal response. The MMTV LTR becomes refractory to dexamethasone treatment after prolonged exposure to ligand; it is possible that the continuous active elimination of GR from the DNA is partly responsible for this phenomenon. In addition, TR is less stable in the presence of ligand than in its absence (54, 160), which is known to be a consequence of liganded TR being preferentially targeted by the ubiquitin-proteasome protein degradation pathway (160), an example of a general phenomenon of activator degradation by proteolysis (161). It is possible, for instance, that such liganded NHR ejection from the DNA as demonstrated by studies of G. Hager and co-workers contributes to a more efficient targeting of NHRs for degradation and leads to a more effective attenuation of response to hormone.


    CONCLUSION
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 
The study of NHR action on chromatin occurs today in a time of great technical and conceptual progress in the field, when new in vitro and in vivo systems offer much promise in that regard. The timing of such developments is an exciting one, since the availability of genomic sequence data from most major taxa, including the complete Drosophila genome and large portion of the human and murine genomes, and new technologies for genome-wide expression profiling (162), will undoubtedly allow for analysis of NHR regulation and behavior in new contexts, while, one hopes, simultaneously providing solutions to old mysteries.


    ACKNOWLEDGMENTS
 
We thank Gordon Hager, Ulla Hansen, and Alexander Strunnikov for sharing results before publication, and Jennifer Lippincott-Schwartz for advice regarding in vivo imaging methods. We thank the anonymous referees for helpful comments and suggestions.


    FOOTNOTES
 
Address requests for reprints to: Alan P. Wolffe, Sangamo Biosciences, Point Richmond Technical Center, 501 Canal Boulevard, Suite A100, Richmond, California 94804. E-mail: awolffe{at}sangamo.com

{smhd3}Note Added in Proof.

Goldmark et al. (Mol Cell 103:423–433) demonstrate by genetic analysis of ISWI in budding yeast that in vivo it functions as a transcriptional corepressor for Ume6p to effect silencing of meiosis-specific genes in a pathway parallel to that driven by Sin3p-Rpd3p. These are the first data on ISWI action at a specific promoter in vivo; they offer an important perspective on in vitro studies proposing that mammalian ISWI is a transcriptional coactivator for such NHRs as PR (Di Croce et al., Mol Cell 4:45) and RAR (Dilworth et al., Mol Cell 6:1049).

Received for publication August 16, 2000. Revision received October 3, 2000. Accepted for publication October 16, 2000.


    REFERENCES
 TOP
 INTRODUCTION
 THE MISREPRESENTED NUCLEOSOME
 ACCESS GRANTED: HOW NHRs...
 THE HOUSE THAT NHRs...
 WHAT IS WITHOUT Sin3:...
 LENDING A HELPING TAIL:...
 IN LIVING COLOR: NHRs...
 CONCLUSION
 REFERENCES
 

  1. Clever U, Karlson P 1960 Induktion von Puff-Veraenderungen in den Speicheldruesenchromosomen von Chironomus tentans durch Ecdyson. Exp Cell Res 20:623–626
  2. Zaret KS, Yamamoto KR 1984 Reversible and persistent changes in chromatin structure accompany activation of a glucocorticoid-dependent enhancer element. Cell 38:29–38[Medline]
  3. Robyr D, Wolffe AP, Wahli W 2000 Nuclear hormone receptor coregulators in action: diversity for shared tasks. Mol Endocrinol 14:329–347[Free Full Text]
  4. Wolffe AP, Guschin D 2000 Chromatin structural features and targets that regulate transcription. J Struct Biol 129:102–122[CrossRef][Medline]
  5. Adams CR, Kamakaka RT 1999 Chromatin assembly: biochemical identities and genetic redundancy. Curr Opin Genet Dev 9:185–190[CrossRef][Medline]
  6. Verreault A 2000 De novo nucleosome assembly: new pieces in an old puzzle. Genes Dev 14:1430–1438[Free Full Text]
  7. Tse C, Fletcher TM, Hansen JC 1998 Enhanced transcription factor access to arrays of histone H3/H4 tetramer. DNA complexes in vitro: implications for replication and transcription. Proc Natl Acad Sci USA 95:12169–12173[Abstract/Free Full Text]
  8. Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ 1997 Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389:251–260[CrossRef][Medline]
  9. Ito T, Ikehara T, Nakagawa T, Kraus WL, Muramatsu M 2000 p300-Mediated acetylation facilitates the transfer of histone H2A-H2B dimers from nucleosomes to a histone chaperone. Genes Dev 14:1899–1907[Abstract/Free Full Text]
  10. Hayes JJ, Tullius TD, Wolffe AP 1990 The structure of DNA in a nucleosome. Proc Natl Acad Sci USA 87:7405–7409[Abstract]
  11. Wolffe AP, Hayes JJ 1999 Chromatin disruption and modification. Nucleic Acids Res 27:711–720[Abstract/Free Full Text]
  12. Allfrey V, Faulkner RM, Mirsky AE 1964 Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proc Natl Acad Sci USA 51:786–794[Medline]
  13. Sterner DE, Berger SL 2000 Acetylation of histones and transcription-related factors. Microbiol Mol Biol Rev 64:435–459[Abstract/Free Full Text]
  14. Wallrath LL, Lu Q, Granok H, Elgin SC 1994 Architectural variations of inducible eukaryotic promoters: preset and remodeling chromatin structures. Bioessays 16:165–170[Medline]
  15. Wolffe AP 1994 Nucleosome positioning and modification: chromatin structures that potentiate transcription. Trends Biochem Sci 19:240–244[CrossRef][Medline]
  16. Imbalzano AN, Kwon H, Green MR, Kingston RE 1994 Facilitated binding of TATA-binding protein to nucleosomal DNA. Nature 370:481–485[CrossRef][Medline]
  17. Kim Y, Geiger JH, Hahn S, Sigler PB 1993 Crystal structure of a yeast TBP/TATA-box complex. Nature 365:512–520[CrossRef][Medline]
  18. Hayes JJ, Wolffe AP 1992 The interaction of transcription factors with nucleosomal DNA. Bioessays 14:597–603[Medline]
  19. Lee DY, Hayes JJ, Pruss D, Wolffe AP 1993 A positive role for histone acetylation in transcription factor access to nucleosomal DNA. Cell 72:73–84[Medline]
  20. Shi YB, Yaoita Y, Brown DD 1992 Genomic organization and alternative promoter usage of the two thyroid hormone receptor ß genes in Xenopus laevis. J Biol Chem 267:733–738[Abstract/Free Full Text]
  21. Rastinejad F, Perlmann T, Evans RM, Sigler PB 1995 Structural determinants of nuclear receptor assembly on DNA direct repeats. Nature 375:203–211[CrossRef][Medline]
  22. Fragoso G, John S, Roberts MS, Hager GL 1995 Nucleosome positioning on the MMTV LTR results from the frequency-biased occupancy of multiple frames. Genes Dev 9:1933–1947[Abstract]
  23. Fragoso G, Pennie WD, John S, Hager GL 1998 The position and length of the steroid-dependent hypersensitive region in the mouse mammary tumor virus long terminal repeat are invariant despite multiple nucleosome B frames. Mol Cell Biol 18:3633–3644[Abstract/Free Full Text]
  24. Li Q, Wrange O 1993 Translational positioning of a nucleosomal glucocorticoid response element modulates glucocorticoid receptor affinity. Genes Dev 7:2471–2482[Abstract]
  25. Li Q, Wrange O 1995 Accessibility of a glucocorticoid response element in a nucleosome depends on its rotational positioning. Mol Cell Biol 15:4375–4384[Abstract]
  26. Wong J, Li Q, Levi BZ, Shi YB, Wolffe AP 1997 Structural and functional features of a specific nucleosome containing a recognition element for the thyroid hormone receptor. EMBO J 16:7130–7145[Abstract/Free Full Text]
  27. Wolffe AP, Urnov FD, Guschin D 2000 Co-repressor complexes and remodelling chromatin for repression. Biochem Soc Trans 28:379–386[Medline]
  28. Sewack GF, Hansen U 1997 Nucleosome positioning and transcription-associated chromatin alterations on the human estrogen-responsive pS2 promoter. J Biol Chem 272:31118–31129[Abstract/Free Full Text]
  29. Bhattacharyya N, Dey A, Minucci S, Zimmer A, John S, Hager G, Ozato K 1997 Retinoid-induced chromatin structure alterations in the retinoic acid receptor ß2 promoter. Mol Cell Biol 17:6481–6490[Abstract]
  30. Minucci S, Wong J, Blanco JC, Shi YB, Wolffe AP, Ozato K 1998 Retinoid receptor-induced alteration of the chromatin assembled on a ligand-responsive promoter in Xenopus oocytes. Mol Endocrinol 12:315–324[Abstract/Free Full Text]
  31. Luisi BF, Xu WX, Otwinowski Z, Freedman LP, Yamamoto KR, Sigler PB 1991 Crystallographic analysis of the interaction of the glucocorticoid receptor with DNA. Nature 352:497–505[CrossRef][Medline]
  32. Fletcher TM, Ryu B-W, Baumann CT, Warren BS, Fragoso G, Hager G 2000 Structure and dynamic properties of a glucocorticoid receptor-induced chromatin transition. Mol Cell Biol 20:6466–6475[Abstract/Free Full Text]
  33. Liu Z, Wong J, Tsai SY, Tsai MJ, O’Malley BW 1999 Steroid receptor coactivator-1 (SRC-1) enhances ligand-dependent and receptor-dependent cell-free transcription of chromatin. Proc Natl Acad Sci USA 96:9485–9490[Abstract/Free Full Text]
  34. Kraus WL, Kadonaga JT 1998 p300 and estrogen receptor cooperatively activate transcription via differential enhancement of initiation and reinitiation. Genes Dev 12:331–342[Abstract/Free Full Text]
  35. Rachez C, Lemon BD, Suldan Z, Bromleigh V, Gamble M, Naar AM, Erdjument-Bromage H, Tempst P, Freedman LP 1999 Ligand-dependent transcription activation by nuclear receptors requires the DRIP complex. Nature 398:824–828[CrossRef][Medline]
  36. Dilworth FJ, Fromental-Ramain C, Remboutsika E, Benecke A, Chambon P 1999 Ligand-dependent activation of transcription in vitro by retinoic acid receptor {alpha}/retinoid X receptor {alpha} heterodimers that mimics transactivation by retinoids in vivo. Proc Natl Acad Sci USA 96:1995–2000[Abstract/Free Full Text]
  37. Guenther MG, Lane WS, Fischle W, Verdin E, Lazar MA, Shiekhattar R 2000 A core SMRT corepressor complex containing HDAC3 and TBL1, a WD40-repeat protein linked to deafness. Genes Dev 14:1048–1057[Abstract/Free Full Text]
  38. Kingston RE, Narlikar GJ 1999 ATP-dependent remodeling and acetylation as regulators of chromatin fluidity. Genes Dev 13:2339–2352[Free Full Text]
  39. Ito M, Yuan CX, Malik S, Gu W, Fondell JD, Yamamura S, Fu ZY, Zhang X, Qin J, Roeder RG 1999 Identity between TRAP and SMCC complexes indicates novel pathways for the function of nuclear receptors and diverse mammalian activators. Mol Cell 3:361–370[Medline]
  40. Roeder RG 1996 The role of general initiation factors in transcription by RNA polymerase II. Trends Biochem Sci 21:327–335[CrossRef][Medline]
  41. Cramer P, Bushnell DA, Fu J, Gnatt AL, Maier-Davis B, Thompson NE, Burgess RR, Edwards AM, David PR, Kornberg RD 2000 Architecture of RNA polymerase II and implications for the transcription mechanism. Science 288:640–649[Abstract/Free Full Text]
  42. Zhang J, Lazar MA 2000 The mechanism of action of thyroid hormones. Annu Rev Physiol 62:439–466[CrossRef][Medline]
  43. Cosma MP, Tanaka T, Nasmyth K 1999 Ordered recruitment of transcription and chromatin remodeling factors to a cell cycle- and developmentally regulated promoter. Cell 97:299–311[Medline]
  44. Krebs JE, Kuo MH, Allis CD, Peterson CL 1999 Cell cycle-regulated histone acetylation required for expression of the yeast HO gene. Genes Dev 13:1412–1421[Abstract/Free Full Text]
  45. Fryer CJ, Archer TK 1998 Chromatin remodelling by the glucocorticoid receptor requires the BRG1 complex. Nature 393:88–91[CrossRef][Medline]
  46. Archer TK, Lefebvre P, Wolford RG, Hager GL 1992 Transcription factor loading on the MMTV promoter: a bimodal mechanism for promoter activation. Science 255:1573–1576[Medline]
  47. McNally JG, Muller WG, Walker D, Wolford R, Hager GL 2000 The glucocorticoid receptor: rapid exchange with regulatory sites in living cells. Science 287:1262–1265[Abstract/Free Full Text]
  48. McPherson CE, Shim EY, Friedman DS, Zaret KS 1993 An active tissue-specific enhancer and bound transcription factors existing in a precisely positioned nucleosomal array. Cell 75:387–398[Medline]
  49. Shim EY, Woodcock C, Zaret KS 1998 Nucleosome positioning by the winged helix transcription factor HNF3. Genes Dev 12:5–10[Abstract/Free Full Text]
  50. Cirillo LA, Zaret KS 1999 An early developmental transcription factor complex that is more stable on nucleosome core particles than on free DNA. Mol Cell 4:961–969[Medline]
  51. Wong J, Shi YB, Wolffe AP 1995 A role for nucleosome assembly in both silencing and activation of the Xenopus TR ß A gene by the thyroid hormone receptor. Genes Dev 9:2696–2711[Abstract]
  52. Pennings S, Meersseman G, Bradbury EM 1991 Mobility of positioned nucleosomes on 5 S rDNA. J Mol Biol 220:101–110[Medline]
  53. Felsenfeld G 1992 Chromatin as an essential part of the transcriptional mechanism. Nature 355:219–224[CrossRef][Medline]
  54. Urnov FD, Yee J, Sachs L, Bauer A, Beug H, Shi YB, Wolffe AP 2000 Targeting of N-CoR and histone deacetylase 3 by the oncoprotein v-ErbA yields a chromatin infrastructure-dependent transcriptional repression pathway. EMBO J 15:1–17[Abstract]
  55. Naar AM, Beaurang PA, Zhou S, Abraham S, Solomon W, Tjian R 1999 Composite co-activator ARC mediates chromatin-directed transcriptional activation. Nature 398:828–832[CrossRef][Medline]
  56. Zhu XG, Hanover JA, Hager GL, Cheng SY 1998 Hormone-induced translocation of thyroid hormone receptors in living cells visualized using a receptor green fluorescent protein chimera. J Biol Chem 273:27058–27063[Abstract/Free Full Text]
  57. Samuels HH, Perlman AJ, Raaka BM, Stanley F 1982 Organization of the thyroid hormone receptor in chromatin. Recent Prog Horm Res 38:557–599[Medline]
  58. Tata JR 1999 Amphibian metamorphosis as a model for studying the developmental actions of thyroid hormone. Biochimie 81:359–366[CrossRef][Medline]
  59. Bauer A, Mikulits W, Lagger G, Stengl G, Brosch G, Beug H 1998 The thyroid hormone receptor functions as a ligand-operated developmental switch between proliferation and differentiation of erythroid progenitors. EMBO J 17:4291–4303[Abstract/Free Full Text]
  60. Sucov HM, Evans RM 1995 Retinoic acid and retinoic acid receptors in development. Mol Neurobiol 10:169–184[Medline]
  61. Haussler MR, Haussler CA, Jurutka PW, Thompson PD, Hsieh JC, Remus LS, Selznick SH, Whitfield GK 1997 The vitamin D hormone and its nuclear receptor: molecular actions and disease states. J Endocrinol 154[Suppl]:S57–S73
  62. Kozlova T, Thummel CS 2000 Steroid regulation of postembryonic development and reproduction in Drosophila. Trends Endocrinol Metab 11:276–280[CrossRef][Medline]
  63. Liu M, Lee MH, Cohen M, Bommakanti M, Freedman LP 1996 Transcriptional activation of the Cdk inhibitor p21 by vitamin D3 leads to the induced differentiation of the myelomonocytic cell line U937. Genes Dev 10:142–153[Abstract]
  64. Stunnenberg HG, Garcia-Jimenez C, Betz JL 1999 Leukemia: the sophisticated subversion of hematopoiesis by nuclear receptor oncoproteins. Biochim Biophys Acta 1423:F15–F33
  65. Hu I, Lazar MA 2000 Transcriptional repression by nuclear hormone receptors. Trends Endocrinol Metab 11:6–10[CrossRef][Medline]
  66. Fondell JD, Roy AL, Roeder RG 1993 Unliganded thyroid hormone receptor inhibits formation of a functional preinitiation complex: implications for active repression. Genes Dev 7:1400–1410[Abstract]
  67. Fondell JD, Brunel F, Hisatake K, Roeder RG 1996 Unliganded thyroid hormone receptor {alpha} can target TATA-binding protein for transcriptional repression. Mol Cell Biol 16:281–287[Abstract]
  68. Seol W, Choi HS, Moore DD 1995 Isolation of proteins that interact specifically with the retinoid X receptor: two novel orphan receptors. Mol Endocrinol 9:72–85[Abstract]
  69. Hoerlein AJ, Naar AM, Heinzel T, Torchia J, Gloss B, Kurokawa R, Ryan A, Kamei Y, Soderstrom M, Glass CK, Rosenfeld MG 1995 Ligand-independent repression by the thyroid hormone receptor mediated by a nuclear receptor co-repressor. Nature 377:397–404[CrossRef][Medline]
  70. Chen JD, Evans RM 1995 A transcriptional co-repressor that interacts with nuclear hormone receptors. Nature 377:454–457[CrossRef][Medline]
  71. Alland L, Muhle R, Hou Jr H, Potes J, Chin L, Schreiber-Agus N, DePinho RA 1997 Role for N-CoR and histone deacetylase in Sin3-mediated transcriptional repression. Nature 387:49–55[CrossRef][Medline]
  72. Laherty CD, Yang WM, Sun JM, Davie JR, Seto E, Eisenman RN 1997 Histone deacetylases associated with the mSin3 corepressor mediate mad transcriptional repression. Cell 89:349–356[Medline]
  73. Hebbes TR, Thorne AW, Crane-Robinson C 1988 A direct link between core histone acetylation and transcriptionally active chromatin. EMBO J 7:1395–1402[Abstract]
  74. Hebbes TR, Clayton AL, Thorne AW, Crane-Robinson C 1994 Core histone hyperacetylation co-maps with generalized DNase I sensitivity in the chicken ß-globin chromosomal domain. EMBO J 13:1823–1830[Abstract]
  75. Bone JR, Lavender J, Richman R, Palmer MJ, Turner BM, Kuroda MI 1994 Acetylated histone H4 on the male X chromosome is associated with dosage compensation in Drosophila. Genes Dev 8:96–104[Abstract]
  76. Jeppesen P, Turner BM 1993 The inactive X chromosome in female mammals is distinguished by a lack of histone H4 acetylation, a cytogenetic marker for gene expression. Cell 74:281–289[Medline]
  77. Heinzel T, Lavinsky RM, Mullen TM, Soderstrom M, Laherty CD, Torchia J, Yang WM, Brard G, Ngo SD, Davie JR, Seto E, Eisenman RN, Rose DW, Glass CK, Rosenfeld MG 1997 A complex containing N-CoR, mSin3 and histone deacetylase mediates transcriptional repression. Nature 387:43–48[CrossRef][Medline]
  78. Nagy L, Kao HY, Chakravarti D, Lin RJ, Hassig CA, Ayer DE, Schreiber SL, Evans RM 1997 Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase. Cell 89:373–380[Medline]
  79. Minucci S, Ozato K 1996 Retinoid receptors in transcriptional regulation. Curr Opin Genet Dev 6:567–574[CrossRef][Medline]
  80. Huang EY, Zhang J, Miska EA, Guenther MG, Kouzarides T, Lazar MA 2000 Nuclear receptor corepressors partner with class II histone deacetylases in a Sin3-independent repression pathway. Genes Dev 14:45–54[Abstract/Free Full Text]
  81. Li J, Wang J, Wang J, Nawaz Z, Liu JM, Qin J, Wong J 2000 Both corepressor proteins SMRT and N-CoR exist in large protein complexes containing HDAC3. EMBO J 19:4342–4350[Abstract/Free Full Text]
  82. Wen YD, Perissi V, Staszewski LM, Yang WM, Krones A, Glass CK, Rosenfeld MG, Seto E 2000 The histone deacetylase-3 complex contains nuclear receptor corepressors. Proc Natl Acad Sci USA 97:7202–7207[Abstract/Free Full Text]
  83. Kao HY, Downes M, Ordentlich P, Evans RM 2000 Isolation of a novel histone deacetylase reveals that class I and class II deacetylases promote SMRT-mediated repression. Genes Dev 14:55–66[Abstract/Free Full Text]
  84. Hu X, Lazar MA 1999 The CoRNR motif controls the recruitment of corepressors by nuclear hormone receptors. Nature 402:93–96[CrossRef][Medline]
  85. Emiliani S, Fischle W, Van Lint C, Al-Abed Y, Verdin E 1998 Characterization of a human RPD3 ortholog, HDAC3. Proc Natl Acad Sci USA 95:2795–2800[Abstract/Free Full Text]
  86. Brown DD, Wang Z, Kanamori A, Eliceiri B, Furlow JD, Schwartzman R 1995 Amphibian metamorphosis: a complex program of gene expression changes controlled by the thyroid hormone. Recent Prog Horm Res 50:309–315[Medline]
  87. Koipally J, Renold A, Kim J, Georgopoulos K 1999 Repression by Ikaros and Aiolos is mediated through histone deacetylase complexes. EMBO J 18:3090–3100[Abstract/Free Full Text]
  88. Koipally J, Georgopoulos K 2000 Ikaros interactions with CtBP reveal a repression mechanism that is independent of histone deacetylase activity. J Biol Chem 275:19594–19602[Abstract/Free Full Text]
  89. Kadosh D, Struhl K 1998 Targeted recruitment of the Sin3-Rpd3 histone deacetylase complex generates a highly localized domain of repressed chromatin in vivo. Mol Cell Biol 18:5121–5127[Abstract/Free Full Text]
  90. Rundlett SE, Carmen AA, Suka N, Turner BM, Grunstein M 1998 Transcriptional repression by UME6 involves deacetylation of lysine 5 of histone H4 by RPD3. Nature 392:831–835[CrossRef][Medline]
  91. Solomon MJ, Larsen PL, Varshavsky A 1988 Mapping protein-DNA interactions in vivo with formaldehyde: evidence that histone H4 is retained on a highly transcribed gene. Cell 53:937–947[Medline]
  92. Aparicio OM, Weinstein DM, Bell SP 1997 Components and dynamics of DNA replication complexes in S. cerevisiae: redistribution of MCM proteins and Cdc45p during S phase. Cell 91:59–69[CrossRef][Medline]
  93. Meluh PB, Koshland D 1997 Budding yeast centromere composition and assembly as revealed by in vivo cross-linking. Genes Dev 11:3401–3412[Abstract/Free Full Text]
  94. Tanaka T, Knapp D, Nasmyth K 1997 Loading of an Mcm protein onto DNA replication origins is regulated by Cdc6p and CDKs. Cell 90:649–660[Medline]
  95. Dernburg AF, Broman KW, Fung JC, Marshall WF, Philips J, Agard DA, Sedat JW 1996 Perturbation of nuclear architecture by long-distance chromosome interactions. Cell 85:745–759[Medline]
  96. Xue Y, Wong J, Moreno GT, Young MK, Cote J, Wang W 1998 NURD, a novel complex with both ATP-dependent chromatin-remodeling and histone deacetylase activities. Mol Cell 2:851–861[Medline]
  97. Wade PA, Jones PL, Vermaak D, Wolffe AP 1998 A multiple subunit Mi-2 histone deacetylase from Xenopus laevis cofractionates with an associated Snf2 superfamily ATPase. Curr Biol 8:843–846[Medline]
  98. Bassi MT, Ramesar RS, Caciotti B, Winship IM, De Grandi A, Riboni M, Townes PL, Beighton P, Ballabio A, Borsani G 1999 X-linked late-onset sensorineural deafness caused by a deletion involving OA1 and a novel gene containing WD-40 repeats. Am J Hum Genet 64:1604–1616[CrossRef][Medline]
  99. Forrest D, Erway LC, Ng L, Altschuler R, Curran T 1996 Thyroid hormone receptor beta is essential for development of auditory function. Nat Genet 13:354–357[Medline]
  100. Roth SY, Dean A, Simpson RT 1990 Yeast {alpha}2 repressor positions nucleosomes in TRP1/ARS1 chromatin. Mol Cell Biol 10:2247–2260[Medline]
  101. Roth SY, Shimizu M, Johnson L, Grunstein M, Simpson RT 1992 Stable nucleosome positioning and complete repression by the yeast {alpha} 2 repressor are disrupted by amino-terminal mutations in histone H4. Genes Dev 6:411–425[Abstract]
  102. Sudarsanam P, Winston F 2000 The Swi/Snf family: nucleosome-remodeling complexes and transcriptional control. Trends Genet 16:345–351[CrossRef][Medline]
  103. Wallberg AE, Neely KE, Hassan AH, Gustafsson JA, Workman JL, Wright AP 2000 Recruitment of the SWI-SNF chromatin remodeling complex as a mechanism of gene activation by the glucocorticoid receptor tau1 activation domain. Mol Cell Biol 20:2004–2013[Abstract/Free Full Text]
  104. Hirschhorn JN, Brown SA, Clark CD, Winston F 1992 Evidence that SNF2/SWI2 and SNF5 activate transcription in yeast by altering chromatin structure. Genes Dev 6:2288–2298[Abstract]
  105. Recht J, Osley MA 1999 Mutations in both the structured domain and N-terminus of histone H2B bypass the requirement for Swi-Snf in yeast. EMBO J 18:229–240[Abstract/Free Full Text]
  106. Wu L, Winston F 1997 Evidence that Snf-Swi controls chromatin structure over both the TATA and UAS regions of the SUC2 promoter in Saccharomyces cerevisiae. Nucleic Acids Res 25:4230–4234[Abstract/Free Full Text]
  107. Gavin IM, Simpson RT 1997 Interplay of yeast global transcriptional regulators Ssn6p-Tup1p and Swi-Snf and their effect on chromatin structure. EMBO J 16:6263–6271[Abstract/Free Full Text]
  108. Gregory PD, Schmid A, Zavari M, Munsterkotter M, Horz W 1999 Chromatin remodelling at the PHO8 promoter requires SWI-SNF and SAGA at a step subsequent to activator binding. EMBO J 18:6407–6414[Abstract/Free Full Text]
  109. Cote J, Quinn J, Workman JL, Peterson CL 1994 Stimulation of GAL4 derivative binding to nucleosomal DNA by the yeast SWI/SNF complex. Science 265:53–60[Medline]
  110. Kwon H, Imbalzano AN, Khavari PA, Kingston RE, Green MR 1994 Nucleosome disruption and enhancement of activator binding by a human SW1/SNF complex. Nature 370:477–481[CrossRef][Medline]
  111. Whitehouse I, Flaus A, Cairns BR, White MF, Workman JL, Owen-Hughes T 1999 Nucleosome mobilization catalysed by the yeast SWI/SNF complex. Nature 400:784–787[CrossRef][Medline]
  112. Wong J, Shi YB, Wolffe AP 1997 Determinants of chromatin disruption and transcriptional regulation instigated by the thyroid hormone receptor: hormone-regulated chromatin disruption is not sufficient for transcriptional activation. EMBO J 16:3158–3171[Abstract/Free Full Text]
  113. Vignali M, Hassan AH, Neely KE, Workman JL 2000 ATP-dependent chromatin-remodeling complexes. Mol Cell Biol 20:1899–1910[Free Full Text]
  114. Di Croce L, Koop R, Venditti P, Westphal HM, Nightingale KP, Corona DF, Becker PB, Beato M 1999 Two-step synergism between the progesterone receptor and the DNA- binding domain of nuclear factor 1 on MMTV minichromosomes. Mol Cell 4:45–54[Medline]
  115. Pile LA, Cartwright IL 2000 GAGA factor-dependent transcription and establishment of DNase hypersensitivity are independent and unrelated events in vivo. J Biol Chem 275:1398–1404[Abstract/Free Full Text]
  116. Cote J, Peterson CL, Workman JL 1998 Perturbation of nucleosome core structure by the SWI/SNF complex persists after its detachment, enhancing subsequent transcription factor binding. Proc Natl Acad Sci USA 95:4947–4952[Abstract/Free Full Text]
  117. Steger DJ, Workman JL 1997 Stable co-occupancy of transcription factors and histones at the HIV-1 enhancer. EMBO J 16:2463–2472[Abstract/Free Full Text]
  118. Pazin MJ, Bhargava P, Geiduschek EP, Kadonaga JT 1997 Nucleosome mobility and the maintenance of nucleosome positioning. Science 276:809–812[Abstract/Free Full Text]
  119. John S, Howe L, Tafrov ST, Grant PA, Sternglanz R, Workman JL 2000 The something about silencing protein, Sas3, is the catalytic subunit of NuA3, a yTAF(II)30-containing HAT complex that interacts with the Spt16 subunit of the yeast CP (Cdc68/Pob3)-FACT complex. Genes Dev 14:1196–1208[Abstract/Free Full Text]
  120. Ura K, Kurumizaka H, Dimitrov S, Almouzni G, Wolffe AP 1997 Histone acetylation: influence on transcription, nucleosome mobility and positioning, and linker histone-dependent transcriptional repression. EMBO J 16:2096–2107[Abstract/Free Full Text]
  121. Belikov S, Gelius B, Almouzni G, Wrange O 2000 Hormone activation induces nucleosome positioning in vivo. EMBO J 19:1023–1033[Abstract/Free Full Text]
  122. Cirillo LA, McPherson CE, Bossard P, Stevens K, Cherian S, Shim EY, Clark KL, Burley SK, Zaret KS 1998 Binding of the winged-helix transcription factor HNF3 to a linker histone site on the nucleosome. EMBO J 17:244–254[Abstract/Free Full Text]
  123. Strahl BD, Allis CD 2000 The language of covalent histone modifications. Nature 403:41–45[CrossRef][Medline]
  124. Ng HH, Bird A 2000 Histone deacetylases: silencers for hire. Trends Biochem Sci 25:121–126[CrossRef][Medline]
  125. Kuo MH, Zhou J, Jambeck P, Churchill ME, Allis CD 1998 Histone acetyltransferase activity of yeast Gcn5p is required for the activation of target genes in vivo. Genes Dev 12:627–639[Abstract/Free Full Text]
  126. Zhang W, Bone JR, Edmondson DG, Turner BM, Roth SY 1998 Essential and redundant functions of histone acetylation revealed by mutation of target lysines and loss of the Gcn5p acetyltransferase. EMBO J 17:3155–3167[Abstract/Free Full Text]
  127. Bresnick EH, John S, Berard DS, LeFebvre P, Hager GL 1990 Glucocorticoid receptor-dependent disruption of a specific nucleosome on the mouse mammary tumor virus promoter is prevented by sodium butyrate. Proc Natl Acad Sci USA 87:3977–3981[Abstract]
  128. Bartsch J, Truss M, Bode J, Beato M 1996 Moderate increase in histone acetylation activates the mouse mammary tumor virus promoter and remodels its nucleosome structure. Proc Natl Acad Sci USA 93:10741–10746[Abstract/Free Full Text]
  129. De Rubertis F, Kadosh D, Henchoz S, Pauli D, Reuter G, Struhl K, Spierer P 1996 The histone deacetylase RPD3 counteracts genomic silencing in Drosophila and yeast. Nature 384:589–591[CrossRef][Medline]
  130. Kouzarides T 1999 Histone acetylases and deacetylases in cell proliferation. Curr Opin Genet Dev 9:40–48[CrossRef][Medline]
  131. Chatterjee VK 1997 Resistance to thyroid hormone. Horm Res 48:43–46[Medline]
  132. Collingwood TN, Rajanayagam O, Adams M, Wagner R, Cavailles V, Kalkhoven E, Matthews C, Nystrom E, Stenlof K, Lindstedt G, Tisell L, Fletterick RJ, Parker MG, Chatterjee VK 1997 A natural transactivation mutation in the thyroid hormone ß receptor: impaired interaction with putative transcriptional mediators. Proc Natl Acad Sci USA 94:248–253[Abstract/Free Full Text]
  133. Collingwood TN, Wagner R, Matthews CH, Clifton-Bligh RJ, Gurnell M, Rajanayagam O, Agostini M, Fletterick RJ, Beck-Peccoz P, Reinhardt W, Binder G, Ranke MB, Hermus A, Hesch RD, Lazarus J, Newrick P, Parfitt V, Raggatt P, de Zegher F, Chatterjee VK 1998 A role for helix 3 of the TRß ligand-binding domain in coactivator recruitment identified by characterization of a third cluster of mutations in resistance to thyroid hormone. EMBO J 17:4760–4770[Abstract/Free Full Text]
  134. Xu J, Qiu Y, DeMayo FJ, Tsai SY, Tsai MJ, O’Malley BW 1998 Partial hormone resistance in mice with disruption of the steroid receptor coactivator-1 (SRC-1) gene. Science 279:1922–1925[Abstract/Free Full Text]
  135. Vignali M, Steger DJ, Neely KE, Workman JL 2000 Distribution of acetylated histones resulting from Gal4-VP16 recruitment of SAGA and NuA4 complexes. EMBO J 19:2629–2640[Abstract/Free Full Text]
  136. Goodman RH, Smolik S 2000 CBP/p300 in cell growth, transformation, and development. Genes Dev 14:1553–1577[Free Full Text]
  137. Schiltz RL, Mizzen CA, Vassilev A, Cook RG, Allis CD, Nakatani Y 1999 Overlapping but distinct patterns of histone acetylation by the human coactivators p300 and PCAF within nucleosomal substrates. J Biol Chem 274:1189–1192[Abstract/Free Full Text]
  138. Chen H, Lin RJ, Xie W, Wilpitz D, Evans RM 1999 Regulation of hormone-induced histone hyperacetylation and gene activation via acetylation of an acetylase. Cell 98:675–686[Medline]
  139. Shimamura A, Worcel A 1989 The assembly of regularly spaced nucleosomes in the Xenopus oocyte S-150 extract is accompanied by deacetylation of histone H4. J Biol Chem 264:14524–14530[Abstract/Free Full Text]
  140. Wittschieben BO, Otero G, de Bizemont T, Fellows J, Erdjument-Bromage H, Ohba R, Li Y, Allis CD, Tempst P, Svejstrup JQ 1999 A novel histone acetyltransferase is an integral subunit of elongating RNA polymerase II holoenzyme. Mol Cell 4:123–128[Medline]
  141. Wittschieben BO, Fellows J, Du W, Stillman DJ, Svejstrup JQ 2000 Overlapping roles for the histone acetyltransferase activities of SAGA and Elongator in vivo. EMBO J 19:3060–3068[Abstract/Free Full Text]
  142. Vettese-Dadey M, Grant PA, Hebbes TR, Crane-Robinson C, Allis CD, Workman JL 1996 Acetylation of histone H4 plays a primary role in enhancing transcription factor binding to nucleosomal DNA in vitro. EMBO J 15:2508–2518[Abstract]
  143. Sewack GF, Ellis TW, Hansen U, Binding of TATA-binding protein to a naturally positioned nucleosome is facilitated by histone acetylation. Mol Cell Biol, in press
  144. Li XY, Virbasius A, Zhu X, Green MR 1999 Enhancement of TBP binding by activators and general transcription factors. Nature 399:605–609[CrossRef][Medline]
  145. Kuras L, Struhl K 1999 Binding of TBP to promoters in vivo is stimulated by activators and requires Pol II holoenzyme. Nature 399:609–613[CrossRef][Medline]
  146. Schild C, Claret FX, Wahli W, Wolffe AP 1993 A nucleosome-dependent static loop potentiates estrogen-regulated transcription from the Xenopus vitellogenin B1 promoter in vitro. EMBO J 12:423–433[Abstract]
  147. Hager GL 2001 Understanding nuclear receptor function: from DNA to chromatin to the interphase nucleus. Prog Nucleic Acids Res Mol Biol 66:279–305[Medline]
  148. Smith CL, Hager GL 1997 Transcriptional regulation of mammalian genes in vivo. A tale of two templates. J Biol Chem 272:27493–27496[Free Full Text]
  149. Archer TK, Lee HL, Cordingley MG, Mymryk JS, Fragoso G, Berard DS, Hager GL 1994 Differential steroid hormone induction of transcription from the mouse mammary tumor virus promoter. Mol Endocrinol 8:568–576[Abstract]
  150. Lee HL, Archer TK 1994 Nucleosome-mediated disruption of transcription factor-chromatin initiation complexes at the mouse mammary tumor virus long terminal repeat in vivo. Mol Cell Biol 14:32–41[Abstract]
  151. Lewontin RC 1991 Biology as Ideology: The Doctrine of DNA. Harper Collins, New York, NY, pp 87–104
  152. Kramer PR, Fragoso G, Pennie W, Htun H, Hager GL, Sinden RR 1999 Transcriptional state of the mouse mammary tumor virus promoter can affect topological domain size in vivo. J Biol Chem 274:28590–28597[Abstract/Free Full Text]
  153. Robinett CC, Straight A, Li G, Willhelm C, Sudlow G, Murray A, Belmont AS 1996 In vivo localization of DNA sequences and visualization of large-scale chromatin organization using lac operator/repressor recognition. J Cell Biol 135:1685–1700[Abstract]
  154. Aragon-Alcaide L, Strunnikov A 2000 Functional dissection of in vivo interchromosome association in S. cerevisiae. Nat Cell Biol 2:812–818[CrossRef][Medline]
  155. Poo M, Cone RA 1974 Lateral diffusion of rhodopsin in the photoreceptor membrane. Nature 247:438–441[Medline]
  156. Ellenberg J, Siggia ED, Moreira JE, Smith CL, Presley JF, Worman HJ, Lippincott-Schwartz J 1997 Nuclear membrane dynamics and reassembly in living cells: targeting of an inner nuclear membrane protein in interphase and mitosis. J Cell Biol 138:1193–1206[Abstract/Free Full Text]
  157. Zaal KJ, Smith CL, Polishchuk RS, Altan N, Cole NB, Ellenberg J, Hirschberg K, Presley JF, Roberts TH, Siggia E, Phair RD, Lippincott-Schwartz J 1999 Golgi membranes are absorbed into and reemerge from the ER during mitosis. Cell 99:589–601[Medline]
  158. Cole NB, Smith CL, Sciaky N, Terasaki M, Edidin M, Lippincott-Schwartz J 1996 Diffusional mobility of Golgi proteins in membranes of living cells. Science 273:797–801[Abstract]
  159. Rigaud G, Roux J, Pictet R, Grange T 1991 In vivo footprinting of rat TAT gene: dynamic interplay between the glucocorticoid receptor and a liver-specific factor. Cell 67:977–986[Medline]
  160. Dace A, Zhao L, Park KS, Furuno T, Takamura N, Nakanishi M, West BK, Hanover JA, Cheng SY 2000 Hormone binding induces rapid proteasome-mediated degradation of thyroid hormone receptors. Proc Natl Acad Sci USA 97:8985–8990[Abstract/Free Full Text]
  161. Molinari E, Gilman M, Natesan S 1999 Proteasome-mediated degradation of transcriptional activators correlates with activation domain potency in vivo. EMBO J 18:6439–6447[Abstract/Free Full Text]
  162. Young RA 2000 Biomedical discovery with DNA arrays. Cell 102:9–15[Medline]
  163. Li J, O’Malley BW, Wong J 2000 p300 requires its histone acetyltransferase activity and SRC-1 interaction domain to facilitate thyroid hormone receptor activation in chromatin. Mol Cell Biol 20:2031–2042[Abstract/Free Full Text]
  164. Wang W, Xue Y, Zhou S, Kuo A, Cairns BR, Crabtree GR 1996 Diversity and specialization of mammalian SWI/SNF complexes. Genes Dev 10:2117–2130[Abstract]