Nuclear Matrix Proteins and Osteoblast Gene ... - Wiley Online Library

6 downloads 74676 Views 894KB Size Report
1Department of Periodontics, Indiana University School of Dentistry, Indianapolis, Indiana, U.S.A.. 2Department of .... attractive candidate for transmitting mechanical informa- tion between ..... ase II and specific transcription factors.(122,123) A ...
JOURNAL OF BONE AND MINERAL RESEARCH Volume 13, Number 2, 1998 Blackwell Science, Inc. © 1998 American Society for Bone and Mineral Research

Review Nuclear Matrix Proteins and Osteoblast Gene Expression JOSEPH P. BIDWELL,1,2 MARTA ALVAREZ,3 HILARY FEISTER,2 JUDE ONYIA,3,4 and JANET HOCK1,4

ABSTRACT The molecular mechanisms that couple osteoblast structure and gene expression are emerging from recent studies on the bone extracellular matrix, integrins, the cytoskeleton, and the nucleoskeleton (nuclear matrix). These proteins form a dynamic structural network, the tissue matrix, that physically links the genes with the substructure of the cell and its substrate. The molecular analog of cell structure is the geometry of the promoter. The degree of supercoiling and bending of promoter DNA can regulate transcriptional activity. Nuclear matrix proteins may render a change in cytoskeletal organization into a bend or twist in the promoter of target genes. We review the role of nuclear matrix proteins in the regulation of gene expression with special emphasis on osseous tissue. Nuclear matrix proteins bind to the osteocalcin and type I collagen promoters in osteoblasts. One such protein is Cbfa1, a recently described transcriptional activator of osteoblast differentiation. Although their mechanisms of action are unknown, some nuclear matrix proteins may act as “architectural” transcription factors, regulating gene expression by bending the promoter and altering the interactions between other trans-acting proteins. The osteoblast nuclear matrix is comprised of cell- and phenotype-specific proteins including proteins common to all cells. Nuclear matrix proteins specific to the osteoblast developmental stage and proteins that distinguish osteosarcoma from the osteoblast have been identified. Recent studies indicating that nuclear matrix proteins mediate bone cell response to parathyroid hormone and vitamin D are discussed. (J Bone Miner Res 1998;13:155– 167)

THE CONNECTION BETWEEN THE SHAPE OF A BONE CELL AND ITS GENETIC PROGRAM

T

HE COUPLING OF CELL SHAPE and gene expression through the linkage of dynamic skeletal networks provides a molecular framework for sensing and responding to mechanical stimuli, fluid flow, and diffusion-based chemical signaling pathways that induce changes in cell morphology.(1–10) This is particularly relevant to osteoblast gene expression. That the shape of a cell is somehow related to

what that cell does has been a venerable article of faith in pathology. However, data to support the concept that molecular mechanisms link cell structure to gene expression are only recently emerging from studies in cell and molecular biology.(1–15) In the tensegrity paradigm, a subcellular scaffolding consisting of proteins from the extracellular matrix (ECM), the integrin receptors, and the cyto- and nucleoskeletons, is physically linked or “hard-wired” to the genes.(16 –18) An outstanding question of this paradigm is how a mechanical tug on the DNA contributes to the regulation of

1

Department of Periodontics, Indiana University School of Dentistry, Indianapolis, Indiana, U.S.A. Department of Anatomy, Indiana University School of Medicine, Indianapolis, Indiana, U.S.A. 3 Department of Oral Biology, Indiana University School of Dentistry, Indianapolis, Indiana, U.S.A. 4 Endocrine Division, Eli Lilly & Co, Indianapolis, Indiana, U.S.A. 2

155

156

BIDWELL ET AL. TABLE 1. PROTEINS THAT REGULATE DNA STRUCTURE

1. Architectural transcription factors HMG proteins(22,128,130–132) LEF-1(129,134,135) c-jun/c-fos(25) POU proteins(27) YY1(30) runt domain proteins(32) 2. Nuclear matrix MAR-binding proteins Topoisomerase II-a and -b(116–118,122,123) SATB1(34) bright(36,74) nucleolin(35) 3. Osteoblast nuclear matrix proteins NMP1/YY1(37,142,164) NMP2/Cbfa1(28,75,142) NMP4(138,146) topoisomerase II-a and -b(105)

High mobility group proteins. Large family of proteins that bend DNA in a variety of promoter contexts. Regulates T-cell receptor a-gene enhancer by bending DNA. fos-jun heterodimers and Jun homodimers induce flexure at the AP-1 site. Oct-1, Oct-2A, Oct-6, and Pit-1 induce DNA bending. Ubiquitously expressed protein that regulates c-fos promoter through DNA bending. These transcriptional regulators bend DNA, perhaps stabilizing higher nucleoprotein complexes. Ubiquitously expressed. May regulate cell development by mediating changes in supercoiling. Predominantly expressed in thymus suggesting a role in gene regulation. B cell–specific protein required for IgH gene transcription. Ubiquitously expressed MAR-binding protein. Localized to both nuclear matrix and chromatin of osteoblasts. Binds to osteocalcin promoter and mediates repression of vitamin D induction. Binds to osteocalcin promoter. Transcriptional activator of osteoblast differentiation. Member of the Runt family. Binds to type I collagen promoter. PTH-responsive and bends DNA. Differentially expressed in immature and mature bone cells. Topo II-a expression is responsive to PTH.

promoter activity, i.e., what are the mechanisms that underlie the cytoskeletal-nuclear communication pathway?(18 –20) Recent studies on transcription indicate that promoter geometry, i.e., the degree of supercoiling and/or the amplitude of a directed bend or loop in the DNA, plays a significant role in regulating the activity of some genes.(21– 35) Architectural transcription factors mediate this promoter structure by inducing bends or twists in the DNA.(21– 36) Some of these architectural transcription factors have been localized to the nucleoskeleton or nuclear matrix,(30,32,34 –38) thus physically linking the cell substructure with promoter geometry and gene activity (Table 1). This model suggests that a change in the genetic program of the cell is more than the result of a redistribution of phosphate groups among proteins but includes a coordinated realignment of structural elements in the cell and nucleus, including the target gene promoter.(16 –18) Here, we review the role of nuclear matrix proteins in gene expression with particular emphasis on the osteoblast. The nuclear matrix will be described as an integral component of the cell’s substructure, linking the geometry of the gene with the periphery of the cell.

THE TISSUE MATRIX PHYSICALLY LINKS THE GENES WITH THE CELL SUBSTRUCTURE AND THE ECM The tissue matrix is the linkage between the structural networks of the ECM, the cytoplasm (cytoskeleton), and

the nucleus (nucleoskeleton or nuclear matrix), that link DNA to the cell periphery (Fig. 1).(39 – 42) Fibronectin, laminin, and the collagens of the ECM, the cell– cell and cell– substrate receptors (e.g., the integrins), the microfilaments, microtubules, and intermediate filaments, the lamins, and the myriad of fibrillogranular proteins of the nuclear matrix all contribute to this tissue–matrix complex.(39,41,42) Tissue matrix proteins play a vital role in cell signaling and underlie the application of the tensegrity theory to signal transduction.(43– 45) Tensegrity is defined as a structural system of discontinuous compression elements coupled by continuous tension cables forming a dynamic structure.(43– 45) Although a detailed discussion is beyond the scope of this review, the consideration of the cell in the context of an interacting tension and compression system confers a capacity for vectorial chemomechanical coupling to second messenger mobilization.(43– 45) In this paradigm, oscillating second messenger signals may propagate and pass amplitude- and frequency-dependent information along structural elements.(43– 45) Protein kinase CK2, a ubiquitously expressed protein putatively involved in growth control, associates with the nuclear matrix, thus integrating soluble and “solid-state” signaling pathways.(44) Conversely, the soluble signaling pathways can effect biochemical modifications, polymerization or depolymerization, cross-linking, and contractile and vibratory movements of the skeletal networks.(39) Does the tissue matrix control gene expression or does gene expression regulate the organization of the tissue ma-

OSTEOBLAST NUCLEAR MATRIX

157

FIG. 1. Osteoblast tissue matrix. The cell substructure is comprised of integrated structural networks of the extracellular matrix, the cytosol, and the nucleus.

trix? Recent evidence indicates a reciprocal relationship. Chen et al.(2) demonstrated that cell shape determined whether individual cells divided or died. Using microfabricated, nonadhesive surfaces coated with fibronectin in various geometric patterns, it was determined that human and bovine capillary endothelial cells switched from growth to apoptosis as the total cell–matrix contact area decreased. Additionally, while maintaining the total cell–matrix contact area constant but changing the degree of cell spreading by altering the spacing between multiple focal adhesionsized islands, cell growth increased and apoptosis decreased with an increase in projected cell area.(2) Similarly, a recent study has demonstrated that depolymerization of microtubules activates the transcription factor NF-kB and induces NF-kB– dependent gene expression in HeLa cells.(20) Colchicine, a disrupter of microtubules, activates mitogen-activated protein (MAP) kinase(46) and protein kinase A (PKA)(47) in rat 3Y1 cells and human monocytes, respectively. The recent characterization of the novel protein kinases misshapen and RON, a receptor protein kinase, is evidence that gene expression can, in turn, regulate cell shape.(48,49) The misshapen gene is required for the normal shape and orientation of Drosophila photoreceptor cells.(48) This protein is also expressed in the embryonic mesoderm, pole plasm, and other sites of cell shape change and may be part of a signal transduction pathway that mediates cytoskeletal organization. The activation of RON alters macrophage morphology and induces migra-

tion of macrophages and epithelial cells.(49) The signal transduction pathway that mediates this change in cell shape includes the binding of macrophage stimulating protein to RON, inducing a phosphorylation of the receptor and phosphatidylinositol-3 kinase (PI-3 kinase); the phosphorylated RON forms a complex with the PI-3 kinase critical to mediating the shape change.(49)

Osteoblast tissue matrix proteins The components of the osteoblast tissue matrix play a significant role in the regulation of gene expression.(50 –54) The interaction between the a2b1 integrin receptor with matrix type I collagen was determined to be critical for mediating osteoblastic differentiation in the bone cell line MC3T3-E1.(52) The progressive development of MC3T3-E1 cells toward the mature osteoblast phenotype includes an increase in the expression of alkaline phosphatase (ALP) and a decrease in the expression transforming growth factor b (TGF-b) receptors.(52) Exposure to inhibitors of collagen synthesis, or a neutralizing antibody to a2b1, abolished these changes in ALP activity and TGF-b receptor expression.(52) This is consistent with Coffey’s maxim that “what a cell touches has a major role in determining what a cell does.”(45) Similarly, neutralizing the binding activity of the integrin avb3, critical to the osteoclast-matrix association, inhibited bone resorption in vitro and prevented osteoporosis in oophorectomized rats.(53) The osteoblast actin cy-

158 toskeleton is involved in mediating the response to fluid flow.(54) Mechanotransduction in bone may occur via loading-dependent flow of interstitial fluid through the haversian-canalicular network.(54) Osteoblasts, as well as osteocytes, in culture responded to pulsating fluid flow by release of enhanced amounts of prostaglandin E2 (PGE2).(54) Treatment of these cell preparations with cytochalasin B, an actin poison, abrogated the release of PGE2 in response to fluid flow.(54)

THE SUBSTRUCTURE OF THE NUCLEUS IS THE NUCLEAR MATRIX, A UNIQUE BIOCHEMICAL FRACTION THAT MEDIATES A CELL TYPE-SPECIFIC ORGANIZATION OF THE GENOME The mechanisms that underlie cytoskeletal-nuclear communication are not known, but the nuclear matrix is an attractive candidate for transmitting mechanical information between the cytosol and nucleus. This nucleoskeleton is operationally defined as the proteinaceous substructure that resists nuclease digestion and high-salt extraction.(55) Transmission electronmicrographs of the nonchromatin structure of the nucleus and analysis of this biochemical fraction indicate that the nuclear matrix includes the lamin– nucleopore complex, a dense fibrillogranular lattice of ribonucleoproteins and 9 –13 nm heteronuclear RNA (hnRNA) filaments, and the chromosome scaffold.(56 –59) This unique biochemical fraction of ribonucleoproteins and lamins is typically represented as a filamentous nuclear scaffolding or a network of channels connected with the nuclear pores.(45,60) However, the nuclear matrix appears to be more than an extension of the tissue matrix into the nuclear space. Like the cytoskeleton, there are nuclear matrix proteins common to all cells.(61– 64) These include proteins like NuMA, a component of the nuclear core filaments of the interchromatin net in interphase cells but an organizer of the microtubules of the mitotic spindle during mitogenesis.(65) The lamins, the outer shell of the nuclear matrix, form a meshwork of filaments that lines the inside surface of the inner nuclear membrane.(66) These proteins are involved in the assembly of the nuclear envelope and higher order chromosomal structures.(66,67) The germ cell–specific lamin B3 appears to mediate the reorganization of nuclear and chromosomal structures during meiotic divisions, and when the protein was expressed in somatic cells in culture, their nuclear morphology was transformed from spherical to hook-shaped.(68) The nuclear matrins are common proteins that are components of the internal nuclear matrix.(69) Also, like the cytoskeleton, the nuclear matrix is comprised of tissue- and phenotypespecific proteins.(57) Recently, numerous DNA-binding proteins unique to this biochemical fraction have been identified, and the nuclear matrix accommodates the transient localization of transcription factors from the chromatin.(39,70,71) Evidence from biochemical, molecular, and confocal microscopy studies have demonstrated roles for nuclear matrix proteins in DNA replication,(72) the processing and transport of RNA transcripts,(73) and transcrip-

BIDWELL ET AL. tional regulation.(74,75) The protein composition and organization of the nuclear matrix reflects and appears to mediate changes in cellular growth, differentiation, and transformation.(39,76 –78) Tumor-specific, nuclear matrix proteins have been characterized in prostate, breast, and bladder tissues.(79 – 81) The diagnostic and prognostic value of these proteins is under investigation and may be the molecular correlate to the pathologist’s observation that cancer is often heralded by changes in nuclear morphology.(82) The nucleoskeleton or nuclear matrix mediates the threedimensional organization of DNA.(39,72,83,84) Although every cell has a complete copy of the genome, the DNA is organized into topologically constrained loop domains which provide the structural basis for tissue- and phenotypic-specific gene expression (Fig. 2).(39,85–91) The eukaryotic nucleus contains over 2 m of DNA, requiring many levels of organization for the ordered packaging of the genome (Fig. 2). The sense and antisense strands of the B-form of the DNA molecule, which comprises the majority of the genome, are organized into the 2 nm right-handed double helix, consisting of two antiparallel strands of purines and pyrimidines. The base pairs (bp) are spaced 0.34 nm apart, and the helix makes a complete turn every 10 bp or 3.4 nm which results in 36° of helical twist.(39) A major and minor groove are formed by the B-form double helix, and these two grooves provide distinct surfaces with which proteins can interact. The double helix wraps around a core of histone proteins, the nucleosomes, to form the secondorder structure. In this “beads on a string” conformation, 200 bp of DNA is wound around each histone octamer core.(39) In the radial coil model of the metaphase chromosome, six nucleosomes coil in a solenoidal conformation to form the 30 nm filament.(39,85–91) This 30 nm filament is further organized into the loop domains, 60 –100 kbp in length, and are attached at their base to the chromosome scaffold(39) (Fig. 2). In the radial coil model, five loops (;300 kbp) are arranged into a radial array subunit. These radial arrays are then turned sideways and coiled to form 240 nm coils that stack to form the chromosome(39) (Fig. 2). It is the proteins of the nuclear matrix that mediate the formation of the topological and functionally distinct DNA loop domains.(39,59,72,83) These loops are fastened to the chromosome scaffold through matrix attachment regions (MARs), also known as scaffold attachment regions (SARs).(39,59,92) Although a variety of MARs have been described,(92) the archetypal MAR/SAR is 200 – 800 bp in length, often A/T-rich, and exhibits the characteristics of a chromatin boundary element. These gene boundaries act to insulate the transcriptional activity of a particular gene from positional influences or from the action of an enhancer from a neighboring gene.(93) Additionally, MARs and the proteins that associate with them participate in the activation of transcription.(74) The architecture of transcription includes the further organization of the nuclear matrix into an interchromatin net that orders splicing factors (snRNPs), hnRNP particles, and thin fibrils into interchromatin granule clusters (IGCs) connected by perichromatin fibrils.(94 –99) For the transcription of some genes, the proteins of the ICGs migrate to the

OSTEOBLAST NUCLEAR MATRIX

159

FIG. 2. The organization of nuclear DNA. The radial coil model. A transcriptionally active loop is shown extended along the nuclear matrix. The loops are attached to the chromosome scaffold through matrix attachment regions (MARs). Transient interactions with the nuclear matrix may be mediated by nuclear matrix elements (NMEs).

gene’s location on the chromosome loop, where transcription and RNA-processing take place.(99) Conversely, for other genes, it is the DNA that moves closer to the ICG for transcription.(97) Transcription is also regulated by the organization of the chromatin, and the nuclear matrix integrates with the chromatin via proteins like CHD1, a novel DNA-binding protein that associates with MARs, localizes to both chromatin and nuclear matrix subfractions, and plays a significant role in the determination of chromatin architecture.(100) The nonrandom topography of transcription reflects the higher order structure of the genome. Interphase chromosomes occupy broad but discrete nuclear territories,(101,102) and active genes are typically localized to the internal nucleus rather than the nuclear envelope.(103) Finally, the small fraction of protein-coding genomic DNA is localized in the G-light bands of the chromosome.(94)

The osteoblast nuclear matrix The protein composition and organization of the osteoblast nuclear matrix reflects the bone cell phenotype.(76 –78) The nuclear matrix protein composition of osteoblasts derived from fetal rat calvaria was specific to successive stages of differentiation in vitro.(76) The two-dimensional sodium dodecyl sulfide-polyacrylamide gel electrophoresis (SDSPAGE) profiles of nuclear matrix proteins from the calvarial cultures revealed proteins specific to the proliferative, matrix synthesis, and matrix mineralization stages of devel-

opment.(76) Additionally, nuclear matrix proteins common to each of these stages of osteoblast development were observed. This would suggest that the bone cell progression from an osteoprogenitor to a mature osteoblast involves a reorganization of nuclear architecture.(76) Major changes in the expression and organization of the major structural nuclear matrix proteins accompany osteoblast development.(104,105) Confocal laser scanning microscopy revealed that NuMA exhibited a diffuse distribution throughout the central planes of the nucleus, excluding the nucleolus, in interphase osteoblasts and osteosarcoma cells,(104) but, as observed in other cell types,(65,84,106,107) localized to the mitotic spindle apparatus during metaphase and anaphase of the dividing bone cell.(104) Similarly, osteoblast expression of the nuclear matrix proteins topoisomerase II (topo II)-a and -b were determined to be specific to the stage of development in bone.(105) These proteins are major structural components of the mitotic chromosome scaffold and interphase nuclear matrix of all cells.(108 –110) The isoforms topo II-a and -b are products of two different genes(110,111–113); topo II-a expression is upregulated during mitogenesis, and topo II-b is typically expressed when cells have plateaued in growth.(110) Immunohistochemical analysis of histologic sections of lumbar vertebrae from 4- to 5-week-old male Sprague-Dawley rats indicated that topo II-a was expressed in cells of the bone marrow cavity, where the osteoprogenitor cells reside, and was conspicuously absent in mature osteoblasts and osteo-

160

BIDWELL ET AL.

FIG. 3. Architectural transcription factors. Architectural transcription factors modulate promoter geometry. In the compass model, the architectural transcription factor bends the DNA and brings distal trans-acting proteins together, without interacting with the other regulatory proteins.(22) In the scaffold model, the DNA-bending proteins make contact with the other trans-acting proteins in the complex.(22) Architectural transcription factors can function by different mechanisms in different promoter contexts. cytes.(105) Topo II-b, however, was expressed in cells of the bone marrow cavity as well as mature osteoblasts and osteocytes.(105) The alteration of osteoblast DNA topology may precede and mediate the sweeping changes in gene expression that occur during the transition from an osteoprogenitor cell to a mature osteoblast. Consistent with a phenotype-specific osteoblast nuclear architecture, nuclear matrix and intermediate filament proteins specific to rat and human osteosarcomas have been characterized.(77,78) In the human studies, one-dimensional SDS-PAGE profiles of the major tissue matrix proteins revealed that the nuclear matrix and intermediate filament protein composition of normal osteoblasts did not vary with biopsy site, age, or gender of patients.(78) The loss of expression of an ;250 kDa intermediate filament protein designated as OB250 was observed in primary human osteosarcoma tumors and human osteosarcoma cell lines. A potential tumor-specific nuclear matrix protein (OT200) was also observed in the osteosarcoma tissue.(78) A redistribution of transcription factors between the osteoblast nuclear matrix and the chromatin, contributes to the change in the nucleoskeleton protein composition.(70) ROS 17/2.8 cells were observed to exhibit a cell typespecific distribution between the chromatin and nuclear matrix subfractions of the trans-acting proteins (or related DNA-binding activities) of SP-1, ATF, CCAAT, OCT-1, and AP-1.(70) The functional significance of the nuclear partitioning of trans-acting proteins has not been established but may mediate a localized concentration of these low-abundant proteins, thus favoring activation of transcription.(70,114)

NUCLEAR MATRIX, DNA-BINDING PROTEINS MAY TRANSDUCE CHANGES IN CELL SHAPE INTO CHANGES IN GENE ACTIVITY BY ALTERING PROMOTER GEOMETRY The cytoskeletal-nuclear communication pathway may involve a tissue matrix–induced alteration in genomic organization that ultimately leads to a change in promoter geometry. The molecular consequence of changing cell shape may be the twisting or bending of the promoters

coupled to the tissue matrix. This mechanical alteration of gene structure could contribute to a change in activity. DNA-binding, nuclear matrix proteins are in a unique position to couple cell and promoter structure. There are numerous DNA-binding, nuclear matrix proteins that associate with the MARs and appear to modulate gene expression through altering some aspect of DNA structure, putatively supercoiling.(34 –36,74) The AT-rich character of these sites confers an anomalous DNA structure that certain kinds of proteins recognize instead of a particular nucleotide sequence.(34,115) For example, ATC sequences are stretches of DNA in which one strand is A’s, T’s, and C’s, excluding G’s. These sequences of more than 20 bp are usually found in clusters within MARs and impart a strong unwinding ability to these sites.(34,35) Proteins such as SATB1 (special AT-rich sequence-binding protein),(34) Bright (B cell regulator of IgH transcription),(36,74) and nucleolin recognize the minor groove of these ATC/MAR sites and may mediate the relief of negative superhelical strain by stabilizing the base unpairing of these regions.(35) Bright, a B cell–specific protein, binds MAR sequences flanking the core of the intronic enhancer Em of the immunoglobulin heavy-chain (IgH) locus and is required for the activation of Em-driven IgH transcription.(36,74) The topo II enzymes are the most well characterized of the nuclear matrix, MAR-binding proteins.(108 –110) Topoisomerases alter DNA structure or topology by a concerted cleavage/religation reaction, which relaxes local supercoiling incurred during DNA replication and transcription.(108 –110) Changes in supercoiling influence the accessibility of DNA by transcription factors,(116 –118) and topo II cleavage sites are distributed throughout the genome, frequently in close proximity to gene enhancer regions.(119 –121) Topo II enzymes also interact directly with RNA polymerase II and specific transcription factors.(122,123) A recent hypothesis proposes the presence of multiple nuclear pools of topo II enzymes that regulate gene expression at the levels of the single and multiple genes.(124) Topo II may govern the expression of single-target genes by binding directly to their promoter regions, and, at the second level, MAR-bound topo II may exert a general effect on the transcriptional expression of clusters of genes by controlling supercoiling of chromosome loops.(124)

OSTEOBLAST NUCLEAR MATRIX Bending the DNA by architectural transcription factors is another mechanism by which gene activity is mechanically regulated.(21–33) The bending or looping of a promoter contributes to governing gene activity by altering the interactions between proximal and/or distal trans-acting proteins and the basic transcription machinery.(21–33) These architectural proteins often lack distinct activation domains, do not necessarily initiate transcription themselves, or interact with the other regulatory proteins.(21–33) As with MARbinding proteins, architectural transcription factors often recognize the minor groove of AT-rich regions of DNA which confers a local structural anomaly to the B-DNA helix.(115,125–127) High mobility group (HMG) proteins are an abundant, heterogeneous family of architectural transcription factors that comprise the nonhistone components of the chromatin.(22) LEF-1 and HMG-I(Y) are both HMG proteins that bind to the minor groove of DNA, induce DNA bending, and are required for enhancer activity but do not activate transcription on their own.(24,128 –135) LEF-1 binds to the T cell receptor a-gene enhancer and induces a sharp bend in the DNA, which brings together two nonadjacent sites, one binding to Ets-1, a lymphocyte-specific trans-acting protein, and the other to an ATF/CREB protein.(129,133–135) The bending protein does not interact with the other regulatory proteins in this “compass” structure(22) (Fig. 3). HMG-I(Y) binds to AT-rich sequences located in the center of a binding site for the dimeric NF-kB protein and adjacent to that for ATF-2 in the promoter of the human interferon b gene.(24,130 –132) Through DNA bending and protein–protein contacts, HMG-I(Y) mediates the formation of a higher order nucleoprotein complex between itself and these other trans-acting proteins, termed an enhanceosome.(24,130 –132) In this “scaffold” model,(22) the DNA-bending proteins make contact with the other proteins in the complex (Fig. 3).

Osteoblast, nuclear matrix, DNA-binding proteins Confocal laser scanning microscopy revealed that topo II enzymes exhibited varied and distinct nuclear distributions in bone cells,(105) consistent with this hypothesis of multifunctional nuclear pools.(124) Topo II-a exhibited a punctate distribution throughout the central portion of the nucleus of cycling ROS 17/2.8 rat osteosarcoma cells and rat primary spongiosa osteoblasts,(105) whereas the distribution of topo II-b was more diffuse, including a localization near the nuclear envelope in both cycling and noncycling cells.(105) Coupled with the differential expression of topo II-a and -b in the developing osteoblast, these nuclear matrix proteins may mediate distinct phenotype-specific MAR organizations, thus potentially regulating clusters of genes through supercoiling. Osteoblast, nuclear matrix, DNA-binding proteins, with a capacity for bending DNA, recognize elements on the osteocalcin and type I collagen promoters.(38,75,136 –138) This establishes a physical link between the regulatory regions of osteoblast ECM genes and cell structure. Osf2/Cbfa1, a transcriptional activator of osteoblast differentiation,(75,137,139 –141) was originally identified as NMP2, an osteoblast-specific, nuclear matrix protein.(38,142) Early

161

FIG. 4. Nuclear matrix protein binding sites on the rat osteocalcin and type I collagen promoters. Overlapping recognition sites for NMP1/YY1 and NMP2/Cbfa1 are located at 2609/2587 nt and at 2457/2430 nt. This latter site contains an AP1/VDRE that mediates NMP1/YY1 repression of vitamin D induction of osteocalcin. The type I collagen cis-regulatory promoter element that mediates osteoblast expression in vivo is located within 21683/21670 nt, but the region between 23600 and 22300 nt is critical for expression in cultured bone cells. NMP3 is between 22149 and 22106 nt and is in proximity to a putative VDRE (22221 nt) and an AP-1 site (22250 nt). NMP4 is located in two AT-rich regions between 23469/23450 nt and 21574/21555 nt. Abbreviations: OC-box, osteocalcin box; GRE, glucocorticoid response element. studies indicated that this DNA-binding protein recognized sites along the rat and mouse osteocalcin promoters (Fig. 4); however, more recent work suggests that this transcription factor may regulate numerous genes including osteopontin and type I collagen (COL1A1).(75) Mice with a homozygous mutation to Cbfa1 died just after birth and lacked mature osteoblasts and a calcified skeleton.(139 –141) NMP2/Osf2/Cbfa1 belongs to the AML/CBF/PEBP2/Runt family of transcription factors which have the capacity to bend DNA.(38) Likewise, YY1, a ubiquitous DNA-binding/ bending protein,(30) was identified as NMP1 in bone cells and was observed to localize in both the chromatin and nuclear matrix fractions of osteoblasts and associated with sequence specificity to the osteocalcin promoter.(37,142) The molecular mechanisms by which NMP2/Osf2/Cbfa1 and NMP1/YY1 mediate osteocalcin transcription have yet to be elucidated, but because these proteins have the capacity to bend DNA, they may act to modulate osteocalcin promoter topology. It has also been proposed that NMP2/Osf2/ Cbfa1 and NMP1/YY1 may also serve to transiently tether the osteocalcin promoter to the nuclear matrix, thereby mediating the local concentration of trans-acting proteins.(114) Two distinct, osteoblast nuclear matrix, DNA-binding proteins, NMP3 and NMP4, associate with sequence-spec-

162 ificity to the rat type I collagen promoter(138) (Fig. 4). Interestingly, the type I collagen cis-regulatory promoter element that mediates osteoblast expression in vivo is located within 21719 and 21670 nucleotides (nt), but the region between 23600 and 22300 nt is critical for expression in cultured bone cells.(143,144) This peculiar dichotomy between transcriptional regulatory regions in vivo and in vitro has been attributed to differences in osteoblast structure within bone and on plastic.(145) NMP3 and NMP4 bind within or in close proximity to these regulatory regions.(138,146) NMP4 has many of the properties of an architectural transcription factor.(146) This nuclear matrix protein, along with an HMG-like, soluble nuclear protein, NP, was determined to bind within the minor groove of two poly(dT) sites between 23469/23450 nt and 21574/21555 nt of the COL1A1 promoter(138,146) (Fig. 4). Circular permutation analysis indicated that protein binding to these sites induced a bend in the DNA.(146) These poly(dT) sites are in proximity to both the in vivo and in vitro regulatory regions of the COL1A1 promoter and consist of a succession of nine uninterrupted thymidines, which are critical to mediating the nuclear matrix protein binding.(146) This stretch of thymidines (and adenines on the antisense strand) form a very narrow DNA minor groove which is attractive to architectural transcription factors.(115) NMP3 binding activity was observed between 22149 and 22106 nt. This region contains a CarG box, an element that has been implicated in DNA bending,(147) and is in proximity to other potential regulatory elements including a putative vitamin D response element (VDRE) (22221 nt) and an AP-1 site (22250 nt).(148) Both the NMP3 and NMP4 proteins exhibited a tissue-restricted distribution and were observed to be selectively expressed in osteoblasts or osteoblast-like cells.(138) NMP3 and NMP4 may link the shift in cis-regulatory elements that regulate the basal transcription of type I collagen between osteoblasts in situ and in culture and may couple osteoblast shape and promoter geometry.

THE NUCLEAR MATRIX IS A TARGET FOR HORMONE ACTION Numerous steroid hormone/receptor complexes, including those that play a significant role in bone metabolism (e.g., 1,25-dihydroxyvitamin D3, and estrogen) localize to the nuclear matrix in a variety of endocrine tissues.(39,45,71,149 –152) Cooperative interactions between nuclear matrix and chromatin proteins may mediate tissue specificity of ligand-induced response by governing the accessibility of target cis elements.(39,45,71) Mammary gland differentiation and involution are mediated by a cooperative signaling pathway between the tissue matrix and the peptide hormone prolactin.(7) This “solid-state” hormonal signaling pathway involves the restructuring of the cytoskeleton and nuclear matrix that allows the prolactin signal to activate transcription of the b-casein gene.(7,8)

BIDWELL ET AL.

Parathyroid hormone, vitamin D, and the osteoblast tissue matrix Both 1,25-dihydroxyvitamin D3 (vitamin D) and parathyroid hormone (PTH) induce dramatic changes in osteoblast shape.(153–158) On the tissue level, these changes in osteoblast morphology may allow access of osteoclasts to the bone surface.(154) At the molecular level, the alteration of tissue matrix protein organization, including the nuclear matrix, may modulate promoter geometry and the accessibility to the regulatory elements of target genes.(71,159) The PTH-induced mobilization of the cyclic adenosine monophosphate (cAMP)-PKA pathway causes a rapid change in osteoblast shape, from a cuboidal to a stellate morphology.(153–158) This alteration in osteoblast shape includes major changes in the expression of structural proteins,(155,160 –163) a retraction of the cell’s peripheral lamellipodia, a massive reorganization of microfilaments, but not microtubules(160) and a decrease in the number of stress fibers.(155) In osteoblasts derived from fetal rat calvariae, PTH induced a rapid (2–5 minutes) dose-dependent (IC50 5 1 nM) depolymerization of actomyosin to about 50% of pretreatment levels attended by an equally rapid (2–3 minutes) dephosphorylation of myosin light chain.(155) These changes in cytoskeletal organization were transient, the levels of polymerized actin and phosphorylated myosin light chain returned to pretreatment levels within 30 minutes and 5 minutes, respectively.(155) Stable alterations in both the human and rat osteoblast cytoskeletons were observed with chronic exposure to PTH.(161–163) PTH selectively alters the bone cell expression of the nuclear matrix proteins topo II-a and NuMA.(104,163) Chronic exposure to hormone (10 nM) increased topo II-a and NuMA protein levels by a factor of two in the rat osteosarcoma cell cultures of ROS 17/2.8.(163) Topo II-b expression was not altered under the same experimental conditions.(163) Immunofluorescent and confocal laser scanning microscopy indicated that PTH increased the number of ROS 17/2.8 cells immunopositive for NuMA as opposed to an increase in NuMA protein/nucleus.(104) It was proposed that this change in NuMA expression, and thus nuclear organization, was coupled to the observed concomitant hormone-induced alterations in the bone cell cytoskeleton.(104) PTH also targets nuclear matrix proteins specific to osteoblast.(146,163) Acute and chronic exposure to PTH attenuated the expression of a 190 kDa nuclear matrix protein in both rat primary spongiosa osteoblasts and ROS 17/2.8 cells within 1 h.(163) This protein exhibited a restricted-tissue distribution and was not observed in opossum kidney (OK) cells, H4 hepatoma cells, or NIH3T3 mouse fibroblasts. Pulse-chase experiments indicated that the 190 kDa protein had a half-life of 1 h in the osteosarcoma cells.(163) These data are consistent with the hypothesis that the osteoblast nuclear matrix is a dynamic structure and is comprised, in part, of labile proteins that respond rapidly to hormonal signals. PTH modulated the DNA-binding activity of NMP4 along the type I collagen promoter.(146) Chronic exposure to hormone (10 nM, 72 h) up-regulated NMP4, but not

OSTEOBLAST NUCLEAR MATRIX NMP3, binding activity in ROS 17/2.8 cells and attenuated COL1A1 mRNA expression.(146) The nuclear matrix protein NMP1/YY1 repressed vitamin D–induced trans-activation of the osteocalcin gene in ROS 17/2.8 cells.(164) Osteosarcoma cells were cotransfected with an osteocalcin promoter-reporter construct (OC-CAT), containing VDRE, in combination with an NMP1/YY1 expression plasmid. NMP1/YY1 repressed vitamin D–induced activity of the OC-CAT construct but had no effect on the basal activity in the absence of vitamin D.(164) NMP1/YY1 repression of the OC promoter activity and vitamin D inducibility were lost using constructs lacking the VDRE. NMP1/YY1 recognition sequences were identified within the VDRE of the osteocalcin gene (Fig. 4). NMP1/YY1 and the vitamin D receptor/retinoid X receptor heterodimers competed for binding at the osteocalcin VDRE and competed for binding to TFIIB, an essential component of the transcriptional initiation complex.(164) How NMP1/YY1 interacted with TFIIB over a distance of 500 nt was not addressed in this study, but some form of DNA bending or looping would likely be required. Therefore, NMP1/YY1-induced changes in osteocalcin promoter geometry may contribute to the mechanisms underlying repression of vitamin D inducibility.

THE OSTEOBLAST NUCLEAR MATRIX PLAYS A REGULATORY ROLE IN GENE EXPRESSION AS CRITICAL AS THE CELL MEMBRANE The osteoblast nuclear matrix is a regulatory conduit through which changes in cell shape are translated into changes in gene activity, and reciprocally, the expression of the genetic program is rendered into three-dimensions. By altering DNA organization and structure, this complex nucleoskeleton is uniquely placed to convert bone cell morphology into promoter geometry (i.e., gene activity) as well as to orchestrate the genetic program of the developing osteoblast by coordinating global changes in DNA organization that mediate the transitions between stages in bone cell development. The nuclear matrix provides a functional paradigm that is particularly useful for an understanding of the cell mechanics of osteoblast response to loading and PTH. The structural continuum from the ECM to the geometry of the promoter may directly link osteoblast deformation, induced by loading or PTH, with a realignment of those genes coupled to the tissue matrix and literally “tug” genes into or out of the genetic program. The study of the osteoblast nuclear matrix brings a topography to the nuclear events that regulate bone cell biology, and a molecular basis to osteoblast structure.

ACKNOWLEDGMENTS This work was supported by National Institutes of Health (NIH) grant R55 DK48310 (J.P.B.) and NIH grant RO1 DE7272 (J.M.H.).

163

REFERENCES 1. Maniotis AJ, Chen CS, Ingber DE 1997 Demonstration of mechanical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure. Proc Natl Acad Sci USA 94:849 – 854. 2. Chen CS, Mrksich M, Huang S, Whitesides GM, Ingber DE 1997 Geometric control of cell life and death. Science 276:1425–1428. 3. Ingber DE, Prusty D, Sun Z, Betensky H, Wang N 1995 Cell shape, cytoskeletal mechanics, and cell cycle control in angiogenesis. J Biomech 28:1471–1484. 4. Malek AM, Izumo S 1996 Mechanism of endothelial cell shape change and cytoskeletal remodeling in response to fluid shear stress. J Cell Sci 109:713–726. 5. Malek AM, Jackman R, Rosenberg RD, Izumo S 1994 Endothelial expression of thrombomodulin is reversibly regulated by fluid shear stress. Circ Res 74:852– 860. 6. Thoumine O, Ziegler T, Girard PR, Nerem RM 1995 Elongation of confluent endothelial cells in culture: The importance of fields of force in the associated alterations of their cytoskeletal structure. Exp Cell Res 219:427– 441. 7. Lelievre S, Weaver VM, Bissell MJ 1996 Extracellular matrix signaling from the cellular membrane skeleton to the nuclear skeleton: A model of gene regulation. Recent Prog Horm Res 51:417– 432. 8. Roskelley CD, Srebrow A, Bissell MJ 1995 A hierarchy of ECM-mediated signalling regulates tissue-specific gene expression. Curr Opin Cell Biol 7:736 –747. 9. Sato M, Ohshima N 1994 Flow-induced changes in shape and cytoskeletal structure of vascular endothelial cells. Biorheology 31:143–153. 10. Goldblum SE, Ding X, Funk SE, Sage EH 1994 SPARC (secreted protein acidic and rich in cysteine) regulates endothelial cell shape and barrier function. Proc Natl Acad Sci USA 91:3448 –3452. 11. Simcha I, Geiger B, Yehuda-Levenberg S, Salomon D, BenZe’ev A 1996 Suppression of tumorigenicity by plakoglobin: An augmenting effect of N-cadherin. J Cell Biol 133:199 –209. 12. Mills JW, Mandell LJ 1994 Cytoskeletal regulation of membrane transport events. FASEB J 8:1161–1165. 13. Cowin P, Burke B 1996 Cytoskeleton-membrane interactions. Curr Opin Cell Biol 8:56 – 65. 14. Rana B, Mischoulon D, Xie Y, Bucher NLR, Farmer SR 1994 Cell-extracellular matrix interactions can regulate the switch between growth and differentiation in rat hepatocytes: Reciprocal expression of C/EBPa and immediate-early growth response transcription factors. Mol Cell Biol 14:5858 –5869. 15. DiPersio CM, Jackson DA, Zaret KS 1991 The extracellular matrix coordinately modulates liver transcription factors and hepatocyte morphology. Mol Cell Biol 11:4405– 4414. 16. Ingber DE 1997 Tensegrity: The architectural basis of cellular mechanotransduction. Ann Rev Physiol 59:575–599. 17. Stamenovic D, Fredberg JJ, Wang N, Butler JP, Ingber DE 1996 A microstructural approach to cytoskeletal mechanics based on tensegrity. J Theor Biol 181:125–136. 18. Ingber DE 1993 The riddle of morphogenesis: A question of solution chemistry or molecular cell engineering? Cell 75: 1249 –1252. 19. Hopkin K 1996 Turning on cells with a twist or a tug. J Natl Inst Health Res 8:23–25. 20. Rosette C, Karin M 1995 Cytoskeletal control of gene expression: depolymerization of microtubules activates NF-kB. J Cell Biol 128:1111–1119. 21. Wolffe AP 1994 Architectural transcription factors. Science 264:1100 –1101. 22. Grosschedl R, Giese K, Pagel J 1994 HMG domain proteins: Architectural elements in the assembly of nucleoprotein structures. Trends Genet 10:94 –100. 23. van der Vliet PC, Verrijzer CP 1993 Bending of DNA by transcription factors. BioEssays 15:25–32. 24. Thanos D, Maniatis T 1995 Virus induction of human IFNb

164

25. 26. 27. 28. 29. 30. 31. 32.

33. 34.

35.

36.

37.

38.

39. 40. 41. 42.

43.

44. 45. 46.

BIDWELL ET AL. gene expression requires the assembly of an enhanceosome. Cell 83:1091–1100. Kerppola TK, Curran T 1993 Selective DNA bending by a variety of bZIP proteins. Mol Cell Biol 13:5479 –5489. Paolella DN, Palmer CR, Schepartz A 1994 DNA targets for certain bZIP proteins distinguished by an intrinsic bend. Science 264:1130 –1133. Verrijzer CP, van Oosterhout JAWM, van Weperen WW, van der Vliet PC 1991 POU proteins bend DNA via the POUspecific domain. EMBO J 10:3007–3014. Oelgeschla¨ger T, Chiang CM, Roeder RG 1996 Topology and reorganization of a human TFIID-promoter complex. Nature 382:735–738. Cullen KE, Kladde MP, Seyfred MA 1993 Interaction between transcription regulatory regions of prolactin chromatin. Science 261:203–206. Natesan S, Gilman MZ 1993 DNA bending and orientationdependent function of YY1 in the c-fos promoter. Genes Dev 7:2497–2509. Bazett-Jones DP, Leblanc B, Herfort M, Moss T 1994 Shortrange DNA looping by the Xenopus HMG-Box transcription factor, xUBF. Science 264:1134 –1137. Golling G, Li LH, Pepling M, Stebbins M, Gergen JP 1996 Drosophila homologs of the protooncogene product PEBP2/ CBFb regulate the DNA-binding properties of runt. Mol Cell Biol 16:932–942. Cress WD, Nevins JR 1996 A role for a bent DNA structure in E2F-mediated transcription activation. Mol Cell Biol 16:2119 –2127. Nakagomi K, Kohwi Y, Dickinson LA, Kohwi-Shigematsu T 1994 A novel DNA-binding motif in the nuclear matrix attachment DNA-binding protein SATB1. Mol Cell Biol 14:1852–1860. Dickinson LA, Kohwi-Shigematsu T 1995 Nucleolin is a matrix attachment region DNA-binding protein that specifically recognizes a region with high base-unpairing potential. Mol Cell Biol 15:456 – 465. Forrester WC, Genderen CV, Jenuwein T, Grosschedl R 1994 Dependence of enhancer-mediated transcription of the immunoglobulin m gene on nuclear matrix attachment regions. Science 265:1221–1225. Guo B, Odgren PR, van Wijnen AJ, Last TJ, Nickerson J, Penman S, Lian JB, Stein JL, Stein GS 1995 The nuclear matrix protein NMP-1 is the transcription factor YY1. Proc Natl Acad Sci USA 92:10526 –10530. Merriman HL, van Wijnen AJ, Hiebert S, Bidwell JP, Fey E, Lian J, Stein J, Stein GS Tissue-specific nuclear matrix protein, NMP-2 is a member of the AML/PEBP2/runt domain transcription factor family: Interactions with the osteocalcin gene promoter. Biochemistry 34:13125–13132. Pienta KJ, Getzenberg RH, Coffey DH 1991 Cell structure and DNA organization. Crit Rev Eukaryot Gene Expr 1: 355–385. Vassy J, Beil M, Irinopoulou T, Rigaut JP 1996 Quantitative image analysis of cytokeratin filament distribution during fetal rat liver development. Hepatology 23:630 – 638. Pienta KJ, Hoover CN 1994 Coupling of cell structure to cell metabolism and function. J Cell Biochem 55:16 –21. Pienta KJ, Coffey DS 1992 Nuclear-cytoskeletal interactions: Evidence for physical connections between the nucleus and cell periphery and their alteration by transformation. J Cell Biochem 49:357–365. Ingber DE, Dike L, Hansen L, Karp S, Liley H, Maniotis A, McNamee H, Mooney D, Plopper G, Sims J, Wang N 1994 Cellular tenesgrity: Exploring how mechanical changes in the cytoskeleton regulate cell growth, migration, and tissue pattern during morphogenesis. Int Rev Cytol 150:173–224. Tawfic S, Faust RA, Gapany M, Ahmed K 1996 Nuclear matrix as an anchor for protein kinase CK2 nuclear signalling. J Cell Biochem 62:165–171. Getzenberg RH, Pienta K, Coffey DS 1990 The tissue matrix: Cell dynamics and hormone action. Endocr Rev 11:399 – 417. Shinohara-Gotoh Y, Nishida E, Hoshi M, Sakai H 1991 Ac-

47.

48. 49.

50. 51. 52.

53.

54.

55. 56. 57. 58.

59. 60. 61.

62.

63. 64. 65. 66. 67.

tivation of microtubule-associated protein kinase by microtubule disruption in quiescent rat 3Y1 cells. Exp Cell Res 193:161–166. Manie S, Schmid-Alliana A, Kubar J, Ferrua B, Rossi B 1993 Disruption of microtubule network in human monocytes induces expression of interleukin-1 but not that of interleukin-6 nor tumor necrosis factor-alpha: Involvement of protein kinase A stimulation. J Biol Chem 268:13675–13681. Treisman JE, Ito N, Rubin GM 1997 Misshapen encodes a protein kinase involved in cell shape control in Drosophila. Gene 186:119 –125. Wang M-H, Montero-Julian FA, Dauny I, Leonard EJ 1996 Requirement of phosphatidylinositol-3 kinase for epithelial cell migration activated by human macrophage stimulating protein. Oncogene 13:2167–2175. Diegelmann RF, Peterkofsky B 1972 Inhibition of collagen secretion from bone and cultured fibroblasts by microtubule disruptive drugs. Proc Natl Acad Sci USA 69:892– 896. Scherft JP, Heersche JNM 1975 Accumulation of collagen containing vacuoles in osteoblasts after administration of colchicine. Cell Tissue Res 157:353–365. Takeuchi Y, Nakayama K, Matsumoto T 1996 Differentiation and cell surface expression of transforming growth factor-b receptors are regulated by interaction with matrix collagen in murine osteoblastic cells. J Biol Chem 271:3938 –3944. Engleman VW, Nickols GA, Ross FP, Horton MA, Griggs DW, Settle SL, Ruminski PG, Teitelbaum SL 1997 A peptidomimetic antagonist of the aVb3 integrin inhibits bone resorption in vitro and prevents osteoporosis in vivo. J Clin Invest 99:2284 –2292. Ajubi NE, Klien-Nulend J, Nijweide PJ, Vrijheid-Lammers T, Alblas MJ, Burger EH 1996 Pulsating fluid flow increases prostaglandin production by cultured chicken osteocytes-a cytoskeleton-dependent process. Biochem Biophys Res Commun 225:62– 68. Berezney R, Coffey DS 1975 Nuclear protein matrix: Association with newly synthesized DNA. Science 189:291–292. Penman S 1995 Rethinking cell structure. Proc Natl Acad Sci USA 92:5251–5257. Nickerson JA, Blencowe BJ, Penman S 1995 The architectural organization of nuclear metabolism. Int Rev Cytol 162A:67–123. Fey EG, Wan J, Penman S 1984 The epithelial cytoskeletal framework and nuclear matrix-intermediate filament scaffold: Three-dimensional organization and protein composition. J Cell Biol 19:1973–1984. Mirkovitch J, Mirault M-E, Laemmli UK 1984 Organization of the higher order chromatin loop: Specific DNA attachment sites on nuclear scaffold. Cell 39:223–232. Razin SV, Gromova II 1995 The channels model of nuclear matrix structure. BioEssays 17:443– 450. Stuurman N, Meijne AM, van der Pol AJ, de Jong L, van Driel R, van Renswoude J 1990 The nuclear matrix from cells of different origin: Evidence for a common set of matrix proteins. J Biol Chem 265:5460 –5465. Holzmann K, Korosec T, Gerner C, Grimm R, Sauermann G 1997 Identification of human common nuclear-matrix proteins as heterogeneous nuclear ribonucleoproteins H and H’ by sequencing and mass spectrometry. Eur J Biochem 244: 479 – 486. De Carcer G, Lallena MJ, Correas I 1995 Protein 4.1 is a component of the nuclear matrix of mammalian cells. Biochem J 312:871– 877. Kallajoke M, Osborn M 1994 Gel electrophoretic analysis of nuclear matrix fractions isolated from different human cell lines. Electrophoresis 15:520 –528. Zeng C, He D, Brinkley BR 1994 Localization of NuMA protein isoforms in the nuclear matrix of mammalian cells. Cell Motil Cytoskel 29:167–176. Hutchison CJ, Bridger JM, Cox LS, Kill IR 1994 Weaving a pattern from disparate threads: Lamin function in nuclear assembly and DNA replication. J Cell Sci 107:3259 –3269. Moir RD, Spann TP, Goldman RD 1995 The dynamic prop-

OSTEOBLAST NUCLEAR MATRIX

68.

69. 70.

71. 72. 73. 74.

75. 76.

77.

78.

79. 80. 81.

82. 83. 84. 85. 86. 87. 88.

erties and possible functions of nuclear lamins. Int Rev Cytol 162B:141–182. Furukawa K, Hotta Y, 1993 cDNA cloning of a germ cell specific lamin B3 from mouse spermatocytes and analysis of its function by ectopic expression in somatic cells. EMBO J 12:97–106. Berezney R, Mortillaro MJ, Ma H, Wei X, Samarabandu J 1995 The nuclear matrix: A structural milieu for genomic function. Int Rev Cytol 162A:1– 65. van Wijnen AJ, Bidwell JP, Fey EG, Penman S, Lian JB, Stein JL, Stein GS 1993 Nuclear matrix association of multiple sequence-specific DNA binding activities related to SP-1, ATF, CCAAT, C/EBP, OCT-1, and AP-1. Biochemistry 32:8397– 8402. Ruh MF, Dunn R 2nd, Ruh TS 1996 Interrelationships between nuclear structure and ligand-activated intracellular receptors. Crit Rev Eukaryot Gene Expr 6:271–283. Dijkwel PA, Hamlin JL 1995 Origins of replication and the nuclear matrix: the DHFR domain as a paradigm. Int Rev Cytol 162A:455– 484. Xing Y, Johnson CV, Dobner PR, Lawrence JB 1993 Higher level organization of individual gene transcription and RNA splicing. Science 259:1326 –1330. Herrscher RF, Kaplan MH, Lelsz DL, Das C, Scheuermann R, Tucker PW 1995 The immunoglobin heavy-cahin matrixassociating regions are bound by Bright: A B cell-specific trans-activator that describes a new DNA-binding protein family. Genes Dev 9:3067–3082. Ducy P, Zhang R, Geoffroy V, Ridall AL, Karsenty G 1997 Osf2/Cbfa1: A transcriptional activator of osteoblast differentiation. Cell 89:747–754. Dworetzky SI, Fey EG, Penman S, Lian JB, Stein JL, Stein GS 1990 Progressive changes in the protein composition of the nuclear matrix during rat osteoblast differentiation. Proc Natl Acad Sci USA 87:4605– 4609. Bidwell JP, Fey EG, van Wijnen AJ, Penman S, Stein JL, Lian JB, Stein GS 1994 Nuclear matrix proteins distinguish normal diploid osteoblasts from osteosarcoma cells. Cancer Res 54:28 –32. Bidwell J, McCabe R, Rougraff B, Feister H, Fey E, Onyia J, Holden J, Hock J 1997 Tissue matrix protein expression in human osteoblasts, osteosarcoma tumors, and osteosarcoma cell lines. Mol Biol Rep 24:271–282. Getzenberg RH, Konety BR, Oeler TA, Quigley MM, Hakam A, Becich MJ 1996 Bladder cancer-associated nuclear matrix proteins. Cancer Res 56:1690 –1694. Khanuja PS, Lehr JE, Soule HD, Gehani SK, Noto AC, Choudhury S, Chen R, Pienta KJ 1993 Nuclear matrix proteins in normal and breast cancer cells. Cancer Res 53:3394–3398. Alberti I, Parodi S, Barboro P, Sanna P, Nicolo G, Allera C, Patrone E, Galli S, Balbi C 1996 Differential nuclear matrixintermediate filament expression in human prostate cancer in respect to benign prostatic hyperplasia. Cancer Lett 109: 193–198. Keesee SK, Briggman JV, Thill G, Wu YJ 1996 Utilization of nuclear matrix proteins for cancer diagnosis. Crit Rev Eukaryot Gene Expr 6:189 –214. Davie JR 1995 The nuclear matrix and the regulation of chromatin organization and function. Int Rev Cytol 162A: 191–250. He D, Zeng C, Brinkley BR 1995 Nuclear matrix proteins as structural and functional components of the mitotic apparatus. Int Rev Cytol 162B:1–74. Manuelidis L 1990 A view of interphase chromosomes. Science 250:1533–1540. Cook PR 1995 A chromomeric model for nuclear and chromosome structure. J Cell Sci 108:2927–2935. Saitoh Y, Laemmli UK 1994 Metaphase chromosome structure: Bands arise from a differential folding path of the highly AT-rich scaffold. Cell 76:609 – 622. Swedlow JR, Agard DA, Sedat JW 1993 Chromosome structure inside the nucleus. Curr Opin Cell Biol 5:412– 416.

165 89. Benbow RM 1992 Chromosome structures. Sci Prog 76: 425–450. 90. Rattner J, Lin CC 1985 Radial loops and helical coils coexist in metaphase chromosomes. Cell 42:291–296. 91. Manuelidis L, Chen TL 1990 A unified model of eukaryotic chromosomes. Cytometry 11:8 –25. 92. Boulikas T 1995 Chromatin domains and prediction of MAR sequences. Int Rev Cytol 162A:279 –388. 93. Eissenberg JC, Elgin SCR 1991 Boundary functions in the control of gene expression. Trends Genet 7:335–340. 94. Moen PT Jr, Smith KP, Lawrence JB 1995 Compartmentalization of specific pre-mRNA metabolism: An emerging view. Hum Mol Genet 4:1779 –1789. 95. Huang S, Spector DL 1996 Intron-dependent recruitment of pre-mRNA splicing factors to sites of transcription. J Cell Biol 133:719 –732. 96. Fakan S, Puvion E 1980 The ultrastructural visualization of nuclear and extranucleolar RNA synthesis and distribution. Int Rev Cytol 65:255–299. 97. Xing Y, Johnson CV, Moen PT Jr, NcNeil JA, Lawrence JB 1995 Nonrandom gene organization: Structrual arrangements of specific pre-mRNA transcription and splicing with SC-35 domains. J Cell Biol 131:1635–1647. 98. Spector DL 1990 Higher order nuclear organization: Threedimensional distribution of small nuclear ribonucleoprotein particles. Proc Natl Acad Sci USA 87:147–151. 99. Misteli T, Caceres JF, Spector DL 1997 The dynamics of a pre-mRNA splicing factor in living cells. Nature 387:523–527. 100. Stokes DG, Perry RP 1995 DNA-binding and chromatin localization properties of CHD1. Mol Cell Biol 15:2745–2753. 101. Manuelidis L, Borden J 1988 Reproducible compartmentalization of individual chromosome domains in human CNS cells revealed by in situ hybridization and three-dimensional reconstruction. Chromosoma 96:397– 410. 102. Cremer T, Cremer C, Schneider T, Baumann H, Hens L, Kirsch-Volders M 1982 Analysis of chromosome positions in the interphase nucleus of Chinese hamster cells by laser-UVmicroirradiaiton experiments. Hum Genet 62:201–209. 103. Lawrence JB, Carter KC, Xing X 1993 Probing functional organization within the nucleus: Is genome structure integrated with RNA metabolism? Cold Spring Harb Symp Quant Biol 58:807– 818. 104. Torrungruang K, Feister H, Swartz D, Hancock EB, Hock J, Bidwell J 1997 Parathyroid hormone (PTH) regulates the expression of nuclear mitotic apparatus protein (NUMA) in the osteoblast-like cells ROS 17/2.8. Bone (in press). 105. Feister HA, Swartz D, Odgren PR, Holden J, Hock J, Onyia J, Bidwell JP 1997 Topoisomerase II expression in osseous tissue. J Cell Biochem 67:451– 465. 106. Zeng C, He D, Berget SM, Brinkley BR 1994 Nuclear-mitotic apparatus protein: A structural protein interface between the nucleoskeleton and RNA splicing. Proc Natl Acad Sci USA 91:1505–1509. 107. Yang CH, Lambie EJ, Snyder M 1992 NuMA: An unusually long coiled-coil related protein in the mammalian nucleus. J Cell Biol 116:1303–1317. 108. Drolet M, Hai-Young W, Liu LF 1994 Role of DNA topoisomerases in transcription. Adv Pharmacol 29A:135–146. 109. Anderson HJ, Roberge M 1992 DNA topoisomerase II: A review of its involvement in chromosome structure, DNA replication, transcription and mitosis. Cell Biol Int Rep 16: 717–723. 110. Giaccone G 1994 DNA topoisomerases and topoisomerase inhibitors. Path Biol 42:346 –352. 111. Watanabe M, Tsutsui K, Tsutsui K, Inoue Y 1994 Differential expression of the topoisomerase IIa and IIb mRNAs in developing rat brain. Neurosci Res 19:51–57. 112. Juenke JM, Holden JA 1993 The distribution of DNA topoisomerase II isoforms in differentiated adult mouse tissues. Biochim Biophys Acta 1216:191–196. 113. Capranico G, Tinelli S, Austin CA, Fisher ML, Zunino F 1992 Different patterns of gene expression of topoisomerase II

166

114. 115. 116. 117. 118. 119.

120.

121. 122.

123.

124.

125. 126. 127. 128. 129.

130.

131. 132. 133. 134. 135.

BIDWELL ET AL. isoforms in differentiated tissues during murine development. Biochim Biophys Acta 1132:43– 48. Stein GS, van Wijnen AJ, Stein J, Lian JB, Montecino M 1995 Contributions of nuclear architecture to transcriptional control. Int Rev Cytol 162A:251–278. Churchill MEA, Travers AA 1991 Protein motifs that recognize structural features of DNA. Trends Biochem Sci 16:92–97. Riou JF, Gabillot M, Riou G 1993 Analysis of topoisomerase II-mediated DNA cleavage of the c-myc gene during HL60 differentiation. FEBS Lett 334:369 –372. Wang JC, Lynch AS 1993 Transcription and DNA supercoiling. Curr Opin Genet Dev 3:764 –768. Wang Z, Dro ¨ge P 1996 Differential control of transcriptioninduced and overall DNA supercoiling by eukaryotic topoisomerases in vitro. EMBO J 15:581–589. Cockerill PN, Garrard WT 1986 Chromosomal loop anchorage of the kappa immunoglobulin gene occurs next to the enhancer in a region containing topoisomerase II sites. Cell 44:273–282. Gasser SM, Laemmli UK 1986 Cohabitation of scaffold binding regions with upstream/enhancer elements of three developmentally regulated genes of D. melanogaster. Cell 46: 521– 530. Adachi Y, Kas E, Laemmli UK 1989 Preferential, cooperative binding of DNA topoisomerase II to scaffold-associated regions. EMBO J 8:3997– 4006. Kroll DJ, Sullivan DM, Gutierrez-Hartman A, Hoeffler JP 1993 Modification of DNA topoisomerase II activity via direct interactions with the cyclic adenosine-39,59-monophosphate response element-binding protein and related transcription factors. Mol Endocrinol 7:305–318. Brou C, Kuhn A, Staub A, Chaudhary S, Grummt I, Davidson I, Tora L 1993 Sequence-specific transactivators counteract topoisomerase II-mediated inhibition of in vitro transcription by RNA polymerases I and II. Nucleic Acids Res 21: 4011– 4018. Constantinou AI, Vaughan AT, Yamasaki H, Kamath N 1996 Commitment to erythroid differentiation in mouse erythroleukemia cells is controlled by alterations in topoisomerase II alpha phosphorylation. Cancer Res 56:4192– 4199. Sinden RR 1994 DNA Structure and Function. Academic Press, San Diego, CA, U.S.A., pp. 398. Wu HM, Crothers DM 1984 The locus of sequence-directed and protein-induced DNA bending. Nature 308:509 –513. Ulanovsky L, Bodner M, Trifonov EN, Choder M 1986 Curved DNA: Design, synthesis, and circularization. Proc Natl Acad Sci USA 83:862– 866. Ferrari S, Harley VR, Pontiggia A, Goodfellow PN, LovellBadge R, Bianchi ME 1992 SRY, like HMG1, recognizes sharp angles in DNA. EMBO J 11:4497– 4506. Giese K, Kingsley C, Kirshner JR, Grosschedl R 1995 Assembly and function of a TCRa enhancer complex is dependent on LEF-1-induced DNA bending and multiple protein-protein interactions. Genes Dev 9:995–1008. Falvo JV, Thanos D, Maniatis T 1995 Reversal of intrinsic DNA bends in the IFNb gene enhancer by transcription factors and the architectural protein HMG I(Y). Cell 83: 1101–1111. Du W, Thanos D, Maniatis T 1993 Mechanisms of transcriptional synergism between distinct virus-inducible enhancer elements. Cell 74:887– 898. Thanos D, Maniatis T 1992 The high mobility group protein HMG I(Y) is required for NF-kB– dependent virus induction of the human IFN-b gene. Cell 71:777–789. Giese K, Cox J, Grosschedl R 1992 The HMG domain of lymphoid enhancer factor 1 bends DNA and facilitates assembly of functional nucleoprotein structures. Cell 69:185–195. Giese K, Amsterdam A, Grosschedl R 1991 DNA-binding properties of the HMG domain of the lymphoid-specific transcriptional regulator LEF-1. Genes Dev 5:2567–2578. Waterman M, Fisher W, Jones K 1991 A thymus-specific member of the HMG protein family regulates the human T cell receptor alpha enhancer. Genes Dev 5:656 – 669.

136. Banerjee C, Hiebert SW, Stein JL, Lian JB, Stein GS 1996 An AML-1 consensus sequence binds an osteoblast-specific complex and transcriptionally activates the osteocalcin gene. Proc Nat Acad Sci USA 93:4968 – 4973. 137. Geoffroy V, Ducy P, Karsenty G 1995 A PEBP2 alpha/AML1-related factor increases osteocalcin promoter activity through its binding to an osteoblast-specific cis-acting element. J Biol Chem 270:30973–30979. 138. Alvarez M, Onyia J, Hock J, Long H, Xu W, Bidwell J 1997 Rat osteoblast and osteosarcoma nuclear matrix proteins bind with sequence specificity to the rat type I collagen promoter. Endocrinology 138:482– 489. 139. Mundlos S, Otto F, Mundlos C, Muliken JB, Aylsworth AS, Albright S, Lindhout D, Cole WG, Henn W, Know JHM, Owen MJ, Mertelsmann R, Zabel BU 1997 Mutations involving the transcription factor Cbfa1 cause cleidocranial dysplasia. Cell 89:773–779. 140. Otto F, Thornell AP, Crompton T, Denzel A, Gilmour KC, Rosewell IR, Stamp GWH, Beddington RSP, Mundlos S, Olsen BR, Selby PB, Owen MJ 1997 Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell 89:765–771. 141. Kormori T, Yagi H, Normura S, Yamaguchi A, Sasaki K, Deguchi K, Shimizu Y, Bronson RT, Gao Y-H, Inada M, Sato M, Okamoto R, Kitamura Y, Yoshiki S, Kishimoto T 1997 Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts. Cell 89:755–764. 142. Bidwell JP, van Wijnen AJ, Fey EG, Dworetzky S, Penman S, Stein JL, Lian JB, Stein GS 1993 Osteocalcin gene promoterbinding factors are tissue-specific nuclear matrix components. Proc Natl Acad Sci USA 90:3162–3166. 143. Bedalov A, Salvatori R, Dodig M, Kronenberg MS, Kapural B, Bogdanovic Z, Kream BE, Woody CO, Clark SH, Mack K, Rowe DW, Lichtler AC 1995 Regulation of COL1A1 expression in type I collagen producing tissues: Identification of a 49 base pair region which is required for transgene expression in bone of transgenic mice. J Bone Miner Res 10:1443–1451. 144. Pavlin D, Lichtler AC, Bedalov A, Kream BE, Harrison JR, Thomas HF, Gronowicz GA, Clark SH, Woody CO, Rowe DW 1992 Differential utilization of regulatory domains within the a 1(I) collagen promoter in osseous and fibroblastic cells. J Cell Biol 116:227–236. 145. Bogdanovic Z, Bedalov A, Krebsbach PH, Pavlin D, Woody CO, Clark SH, Thomas HF, Rowe DW, Kream BE, Lichtler AC 1994 Upstream regulatory elements necessary for expression of the rat COL1A1 promoter in transgenic mice. J Bone Miner Res 9:285–292. 146. Alvarez M, Thunyakittpisal P, Morrison P, Onyia J, Hock J, Bidwell J 1997 Parathyroid hormone (PTH) alters osteoblast, nuclear matrix protein binding to the type I collagen promoter. J Bone Miner Res 12:S157. 147. Gustafson TA, Taylor A, Kedes L 1989 DNA bending is induced by a transcription factor that interacts with the human c-fos and a-actin promoters. Proc Natl Acad Sci USA 86:2162–2166. 148. Kream BE, LaFrancis D, Petersen DN, Woody C, Clark S, Rowe DW, Lichtler A 1993 Parathyroid hormone represses a1(I) collagen promoter activity in cultured calvariae from neonatal transgenic mice. Mol Endocrinol 7:399 – 408. 149. Schuchard M, Landers JP, Sandhu NP, Spelsberg TC 1993 Steroid hormone regulation of nuclear proto-oncogenes. Endocr Rev 14:659 – 669. 150. Konety BR, Schwartz GG, Acierno JS Jr, Becich MJ, Getzenberg RH 1996 The role of vitamin D in normal prostate growth and differentiation. Cell Growth Differ 7:1563–1570. 151. Narvaez CJ, Vanweelden K, Byrne I, Welsh J 1996 Characterization of a vitamin D3-resistant MCF-7 cell line. Endocrinology 137:400 – 409. 152. Coutts AS, Davie JR, Dotzlaw H, Murphy LC 1996 Estrogen regulation of nuclear matrix-intermediate filament proteins in human breast cancer cells. J Cell Biochem 63:174 –184. 153. Lloyd QP, Kuhn MA, Gay CV 1995 Characterization of cal-

OSTEOBLAST NUCLEAR MATRIX

154. 155.

156.

157.

158.

159. 160.

cium translocation across the plasma membrane of primary osteoblasts using a lipophilic calcium-sensitive fluorescent dye, calcium green C18. J Biol Chem 270:22445–22451. Bronner F, Stein WD 1992 Modulation of bone calciumbinding sites regulates plasma calcium: An hypothesis. Calcif Tissue Int 50:483– 489. Egan JJ, Grononwicz G, Rodan G 1991 Parathyroid hormone promotes the disassembly of cytoskeletal actin and myosin in cultured osteoblastic cells: mediation by cyclic AMP. J Cell Biochem 45:101–111. Ali NN, Melhuish PB, Boyde A, Bennet A, Jones SJ 1990 Parathyroid hormone, but not protaglandin E2, changes the shape of osteoblasts maintained on bone in vitro. J Bone Miner Res 5:115–121. Ferrier J, Ward-Kesthely A, Heersche JN, Aubin JE 1988 Membrane potential changes, cAMP stimulation and contraction in osteoblast-like UMR 106 cells in response to calcitonin and parathyroid hormone. Bone Miner 4:133–145. Pockwinse SM, Stein JL, Lian JB, Stein GS 1995 Developmental stage-specific cellular responses to vitamin D and glucocorticoids during differentiation of the osteoblast phenotype: interrelationship of morphology and gene expression by in situ hybridization. Exp Cell Res 216:244 –260. Ingber DE 1993 Cellular tenesgrity: Defining new rules of biological design that govern the cytoskeleton. J Cell Sci 104:613– 627. Aubin JE, Alders E, Heersche JMN 1983 A primary role for microfilaments but not microtubules in hormone induced cytoplasmic retraction. Exp Cell Res 143:439 – 450.

167 161. Lormi A, Marie PJ 1990 Changes in cytoskeletal proteins in response to parathyroid hormone and 1,25-dihydroxyvitamin D in human osteoblastic cells. Bone Miner 10:1–12. 162. Bidwell JP, van Wijnen AJ, Banerjee C, Fey EG, Merriman H, Penman S, Stein JL, Lian J, Stein GS 1994 PTH-responsive modifications in the nuclear matrix of ROS 17/2.8 rat osteosarcoma cells. Endocrinology 134:1738 –1744. 163. Bidwell J, Feister H, Swartz D, Onyia J, Holden J, Hock J 1996 Parathyroid hormone regulates the expression of rat osteoblast and osteosarcoma nuclear matrix proteins. J Cell Biochem 63:374 –383. 164. Guo B, Aslam F, van Wijnen AJ, Roberts SG, Frenkel B, Green MR, DeLuca H, Lian JB, Stein GS, Stein JL 1997 YY1 regulates vitamin D receptor/retinoid X receptor mediated transactivation of the vitamin D responsive osteocalcin gene. Proc Natl Acad Sci USA 94:121–126.

Address reprint requests to: Joseph P. Bidwell Indiana University School of Dentistry 1121 West Michigan Street Indianapolis, IN 46202 U.S.A. Received in original form January 24, 1997; in revised form August 8, 1997; accepted October 8, 1997.