Chromatin remodeling

Last updated

Chromatin remodeling is the dynamic modification of chromatin architecture to allow access of condensed genomic DNA to the regulatory transcription machinery proteins, and thereby control gene expression. Such remodeling is principally carried out by 1) covalent histone modifications by specific enzymes, e.g., histone acetyltransferases (HATs), deacetylases, methyltransferases, and kinases, and 2) ATP-dependent chromatin remodeling complexes which either move, eject or restructure nucleosomes. [1] Besides actively regulating gene expression, dynamic remodeling of chromatin imparts an epigenetic regulatory role in several key biological processes, egg cells DNA replication and repair; apoptosis; chromosome segregation as well as development and pluripotency. Aberrations in chromatin remodeling proteins are found to be associated with human diseases, including cancer. Targeting chromatin remodeling pathways is currently evolving as a major therapeutic strategy in the treatment of several cancers.

Contents

Overview

Chromatin organization: The basic unit of chromatin organization is the nucleosome, which comprises 147 bp of DNA wrapped around a core of histone proteins. The level of nucleosomal packaging can have profound consequences on all DNA-mediated processes including gene regulation. Euchromatin (loose or open chromatin) structure is permissible for transcription whereas heterochromatin (tight or closed chromatin) is more compact and refractory to factors that need to gain access to the DNA template. Nucleosome positioning and chromatin compaction can be influenced by a wide range of processes including modification to both histones and DNA and ATP-dependent chromatin remodeling complexes. Sha-Boyer-Fig1-CCBy3.0.jpg
Chromatin organization: The basic unit of chromatin organization is the nucleosome, which comprises 147 bp of DNA wrapped around a core of histone proteins. The level of nucleosomal packaging can have profound consequences on all DNA-mediated processes including gene regulation. Euchromatin (loose or open chromatin) structure is permissible for transcription whereas heterochromatin (tight or closed chromatin) is more compact and refractory to factors that need to gain access to the DNA template. Nucleosome positioning and chromatin compaction can be influenced by a wide range of processes including modification to both histones and DNA and ATP-dependent chromatin remodeling complexes.

The transcriptional regulation of the genome is controlled primarily at the preinitiation stage by binding of the core transcriptional machinery proteins (namely, RNA polymerase, transcription factors, and activators and repressors) to the core promoter sequence on the coding region of the DNA. However, DNA is tightly packaged in the nucleus with the help of packaging proteins, chiefly histone proteins to form repeating units of nucleosomes which further bundle together to form condensed chromatin structure. Such condensed structure occludes many DNA regulatory regions, not allowing them to interact with transcriptional machinery proteins and regulate gene expression. To overcome this issue and allow dynamic access to condensed DNA, a process known as chromatin remodeling alters nucleosome architecture to expose or hide regions of DNA for transcriptional regulation.

By definition, chromatin remodeling is the enzyme-assisted process to facilitate access of nucleosomal DNA by remodeling the structure, composition and positioning of nucleosomes.

Classification

Access to nucleosomal DNA is governed by two major classes of protein complexes:

  1. Covalent histone-modifying complexes.
  2. ATP-dependent chromatin remodeling complexes.

Covalent histone-modifying complexes

Specific protein complexes, known as histone-modifying complexes catalyze addition or removal of various chemical elements on histones. These enzymatic modifications include acetylation, methylation, phosphorylation, and ubiquitination and primarily occur at N-terminal histone tails. Such modifications affect the binding affinity between histones and DNA, and thus loosening or tightening the condensed DNA wrapped around histones, e.g., Methylation of specific lysine residues in H3 and H4 causes further condensation of DNA around histones, and thereby prevents binding of transcription factors to the DNA that lead to gene repression. On the contrary, histone acetylation relaxes chromatin condensation and exposes DNA for TF binding, leading to increased gene expression. [3]

Known modifications

Well characterized modifications to histones include: [4]

Both lysine and arginine residues are known to be methylated. Methylated lysines are the best understood marks of the histone code, as specific methylated lysine match well with gene expression states. Methylation of lysines H3K4 and H3K36 is correlated with transcriptional activation while demethylation of H3K4 is correlated with silencing of the genomic region. Methylation of lysines H3K9 and H3K27 is correlated with transcriptional repression. [5] Particularly, H3K9me3 is highly correlated with constitutive heterochromatin. [6]

  • Acetylation - by HAT (histone acetyl transferase); deacetylation - by HDAC (histone deacetylase)

Acetylation tends to define the 'openness' of chromatin as acetylated histones cannot pack as well together as deacetylated histones.

However, there are many more histone modifications, and sensitive mass spectrometry approaches have recently greatly expanded the catalog. [7]

Histone code hypothesis

The histone code is a hypothesis that the transcription of genetic information encoded in DNA is in part regulated by chemical modifications to histone proteins, primarily on their unstructured ends. Together with similar modifications such as DNA methylation it is part of the epigenetic code.

Cumulative evidence suggests that such code is written by specific enzymes which can (for example) methylate or acetylate DNA ('writers'), removed by other enzymes having demethylase or deacetylase activity ('erasers'), and finally readily identified by proteins ('readers') that are recruited to such histone modifications and bind via specific domains, e.g., bromodomain, chromodomain. These triple action of 'writing', 'reading' and 'erasing' establish the favorable local environment for transcriptional regulation, DNA-damage repair, etc. [8]

The critical concept of the histone code hypothesis is that the histone modifications serve to recruit other proteins by specific recognition of the modified histone via protein domains specialized for such purposes, rather than through simply stabilizing or destabilizing the interaction between histone and the underlying DNA. These recruited proteins then act to alter chromatin structure actively or to promote transcription.

A very basic summary of the histone code for gene expression status is given below (histone nomenclature is described here):

Type of
modification
Histone
H3K4H3K9H3K14H3K27H3K79H4K20H2BK5
mono-methylation activation [9] activation [10] activation [10] activation [10] [11] activation [10] activation [10]
di-methylation repression [5] repression [5] activation [11]
tri-methylationactivation [12] repression [10] repression [10] activation, [11]
repression [10]
repression [5]
acetylation activation [12] activation [12]

ATP-dependent chromatin remodeling

ATP-dependent chromatin-remodeling complexes regulate gene expression by either moving, ejecting or restructuring nucleosomes. These protein complexes have a common ATPase domain and energy from the hydrolysis of ATP allows these remodeling complexes to reposition nucleosomes (often referred to as "nucleosome sliding") along the DNA, eject or assemble histones on/off of DNA or facilitate exchange of histone variants, and thus creating nucleosome-free regions of DNA for gene activation. [13] Also, several remodelers have DNA-translocation activity to carry out specific remodeling tasks. [14]

All ATP-dependent chromatin-remodeling complexes possess a sub unit of ATPase that belongs to the SNF2 superfamily of proteins. In association to the sub unit's identity, two main groups have been classified for these proteins. These are known as the SWI2/SNF2 group and the imitation SWI (ISWI) group. The third class of ATP-dependent complexes that has been recently described contains a Snf2-like ATPase and also demonstrates deacetylase activity. [15]

Known chromatin remodeling complexes

INO80 stabilizes replication forks and counteracts mislocalization of H2A.Z INO80 stabilizes replication forks and counteracts mislocalization of H2A.Z.png
INO80 stabilizes replication forks and counteracts mislocalization of H2A.Z

There are at least four families of chromatin remodelers in eukaryotes: SWI/SNF, ISWI, NuRD/Mi-2/CHD, and INO80 with first two remodelers being very well studied so far, especially in the yeast model. Although all of remodelers share common ATPase domain, their functions are specific based on several biological processes (DNA repair, apoptosis, etc.). This is due to the fact that each remodeler complex has unique protein domains (Helicase, bromodomain, etc.) in their catalytic ATPase region and also has different recruited subunits.

Specific functions

  • Several in-vitro experiments suggest that ISWI remodelers organize nucleosome into proper bundle form and create equal spacing between nucleosomes, whereas SWI/SNF remodelers disorder nucleosomes.
  • The ISWI-family remodelers have been shown to play central roles in chromatin assembly after DNA replication and maintenance of higher-order chromatin structures.
  • INO80 and SWI/SNF-family remodelers participate in DNA double-strand break (DSB) repair and nucleotide-excision repair (NER) and thereby plays crucial role in TP53 mediated DNA-damage response.
  • NuRD/Mi-2/CHD remodeling complexes primarily mediate transcriptional repression in the nucleus and are required for the maintenance of pluripotency of embryonic stem cells. [13]

Significance

Chromatin remodeling complexes in the dynamic regulation of transcription: In the presence of acetylated histones (HAT mediated) and absence of methylase (HMT) activity, chromatin is loosely packaged. Additional nucleosome repositioning by chromatin remodeler complex, SWI/SNF opens up DNA region where transcription machinery proteins, like RNA Pol II, transcription factors and co-activators bind to turn on gene transcription. In the absence of SWI/SNF, nucleosomes can not move farther and remain tightly aligned to one another. Additional methylation by HMT and deacetylation by HDAC proteins condenses DNA around histones and thus, make DNA unavailable for binding by RNA Pol II and other activators, leading to gene silencing. Luong LD SA F2.jpg
Chromatin remodeling complexes in the dynamic regulation of transcription: In the presence of acetylated histones (HAT mediated) and absence of methylase (HMT) activity, chromatin is loosely packaged. Additional nucleosome repositioning by chromatin remodeler complex, SWI/SNF opens up DNA region where transcription machinery proteins, like RNA Pol II, transcription factors and co-activators bind to turn on gene transcription. In the absence of SWI/SNF, nucleosomes can not move farther and remain tightly aligned to one another. Additional methylation by HMT and deacetylation by HDAC proteins condenses DNA around histones and thus, make DNA unavailable for binding by RNA Pol II and other activators, leading to gene silencing.

In normal biological processes

Chromatin remodeling plays a central role in the regulation of gene expression by providing the transcription machinery with dynamic access to an otherwise tightly packaged genome. Further, nucleosome movement by chromatin remodelers is essential to several important biological processes, including chromosome assembly and segregation, DNA replication and repair, embryonic development and pluripotency, and cell-cycle progression. Deregulation of chromatin remodeling causes loss of transcriptional regulation at these critical check-points required for proper cellular functions, and thus causes various disease syndromes, including cancer.

Response to DNA damage

Chromatin relaxation is one of the earliest cellular responses to DNA damage. [16] Several experiments have been performed on the recruitment kinetics of proteins involved in the response to DNA damage. The relaxation appears to be initiated by PARP1, whose accumulation at DNA damage is half complete by 1.6 seconds after DNA damage occurs. [17] This is quickly followed by accumulation of chromatin remodeler Alc1, which has an ADP-ribose–binding domain, allowing it to be quickly attracted to the product of PARP1. The maximum recruitment of Alc1 occurs within 10 seconds of DNA damage. [16] About half of the maximum chromatin relaxation, presumably due to action of Alc1, occurs by 10 seconds. [16] PARP1 action at the site of a double-strand break allows recruitment of the two DNA repair enzymes MRE11 and NBS1. Half maximum recruitment of these two DNA repair enzymes takes 13 seconds for MRE11 and 28 seconds for NBS1. [17]

Another process of chromatin relaxation, after formation of a DNA double-strand break, employs γH2AX, the phosphorylated form of the H2AX protein. The histone variant H2AX constitutes about 10% of the H2A histones in human chromatin. [18] γH2AX (phosphorylated on serine 139 of H2AX) was detected at 20 seconds after irradiation of cells (with DNA double-strand break formation), and half maximum accumulation of γH2AX occurred in one minute. [18] The extent of chromatin with phosphorylated γH2AX is about two million base pairs at the site of a DNA double-strand break. [18]

γH2AX does not, by itself, cause chromatin decondensation, but within seconds of irradiation the protein "Mediator of the DNA damage checkpoint 1" (MDC1) specifically attaches to γH2AX. [19] [20] This is accompanied by simultaneous accumulation of RNF8 protein and the DNA repair protein NBS1 which bind to MDC1 as MDC1 attaches to γH2AX. [21] RNF8 mediates extensive chromatin decondensation, through its subsequent interaction with CHD4 protein, [22] a component of the nucleosome remodeling and deacetylase complex NuRD. CHD4 accumulation at the site of the double-strand break is rapid, with half-maximum accumulation occurring by 40 seconds after irradiation. [23]

The fast initial chromatin relaxation upon DNA damage (with rapid initiation of DNA repair) is followed by a slow recondensation, with chromatin recovering a compaction state close to its pre-damage level in ~ 20 min. [16]

Cancer

Chromatin remodeling provides fine-tuning at crucial cell growth and division steps, like cell-cycle progression, DNA repair and chromosome segregation, and therefore exerts tumor-suppressor function. Mutations in such chromatin remodelers and deregulated covalent histone modifications potentially favor self-sufficiency in cell growth and escape from growth-regulatory cell signals - two important hallmarks of cancer. [24]

Cancer genomics

Rapid advance in cancer genomics and high-throughput ChIP-chip, ChIP-Seq and Bisulfite sequencing methods are providing more insight into role of chromatin remodeling in transcriptional regulation and role in cancer.

Therapeutic intervention

Epigenetic instability caused by deregulation in chromatin remodeling is studied in several cancers, including breast cancer, colorectal cancer, pancreatic cancer. Such instability largely cause widespread silencing of genes with primary impact on tumor-suppressor genes. Hence, strategies are now being tried to overcome epigenetic silencing with synergistic combination of HDAC inhibitors or HDI and DNA-demethylating agents. HDIs are primarily used as adjunct therapy in several cancer types. [36] [37] HDAC inhibitors can induce p21 (WAF1) expression, a regulator of p53's tumor suppressoractivity. HDACs are involved in the pathway by which the retinoblastoma protein (pRb) suppresses cell proliferation. [38] Estrogen is well-established as a mitogenic factor implicated in the tumorigenesis and progression of breast cancer via its binding to the estrogen receptor alpha (ERα). Recent data indicate that chromatin inactivation mediated by HDAC and DNA methylation is a critical component of ERα silencing in human breast cancer cells. [39]

Current front-runner candidates for new drug targets are Histone Lysine Methyltransferases (KMT) and Protein Arginine Methyltransferases (PRMT). [44]

Other disease syndromes

Senescence

Chromatin architectural remodeling is implicated in the process of cellular senescence, which is related to, and yet distinct from, organismal aging. Replicative cellular senescence refers to a permanent cell cycle arrest where post-mitotic cells continue to exist as metabolically active cells but fail to proliferate. [47] [48] Senescence can arise due to age associated degradation, telomere attrition, progerias, pre-malignancies, and other forms of damage or disease. Senescent cells undergo distinct repressive phenotypic changes, potentially to prevent the proliferation of damaged or cancerous cells, with modified chromatin organization, fluctuations in remodeler abundance, and changes in epigenetic modifications. [49] [50] [47] Senescent cells undergo chromatin landscape modifications as constitutive heterochromatin migrates to the center of the nucleus and displaces euchromatin and facultative heterochromatin to regions at the edge of the nucleus. This disrupts chromatin-lamin interactions and inverts of the pattern typically seen in a mitotically active cell. [51] [49] Individual Lamin-Associated Domains (LADs) and Topologically Associating Domains (TADs) are disrupted by this migration which can affect cis interactions across the genome. [52] Additionally, there is a general pattern of canonical histone loss, particularly in terms of the nucleosome histones H3 and H4 and the linker histone H1. [51] Histone variants with two exons are upregulated in senescent cells to produce modified nucleosome assembly which contributes to chromatin permissiveness to senescent changes. [52] Although transcription of variant histone proteins may be elevated, canonical histone proteins are not expressed as they are only made during the S phase of the cell cycle and senescent cells are post-mitotic. [51] During senescence, portions of chromosomes can be exported from the nucleus for lysosomal degradation which results in greater organizational disarray and disruption of chromatin interactions. [50]

Chromatin remodeler abundance may be implicated in cellular senescence as knockdown or knockout of ATP-dependent remodelers such as NuRD, ACF1, and SWI/SNP can result in DNA damage and senescent phenotypes in yeast, C. elegans, mice, and human cell cultures. [53] [50] [54] ACF1 and NuRD are downregulated in senescent cells which suggests that chromatin remodeling is essential for maintaining a mitotic phenotype. [53] [54] Genes involved in signaling for senescence can be silenced by chromatin confirmation and polycomb repressive complexes as seen in PRC1/PCR2 silencing of p16. [55] [56] Specific remodeler depletion results in activation of proliferative genes through a failure to maintain silencing. [50] Some remodelers act on enhancer regions of genes rather than the specific loci to prevent re-entry into the cell cycle by forming regions of dense heterochromatin around regulatory regions. [56]

Senescent cells undergo widespread fluctuations in epigenetic modifications in specific chromatin regions compared to mitotic cells. Human and murine cells undergoing replicative senescence experience a general global decrease in methylation; however, specific loci can differ from the general trend. [57] [52] [50] [55] Specific chromatin regions, especially those around the promoters or enhancers of proliferative loci, may exhibit elevated methylation states with an overall imbalance of repressive and activating histone modifications. [49] Proliferative genes may show increases in the repressive mark H3K27me3 while genes involved in silencing or aberrant histone products may be enriched with the activating modification H3K4me3. [52] Additionally, upregulating histone deacetylases, such as members of the sirtuin family, can delay senescence by removing acetyl groups that contribute to greater chromatin accessibility. [58] General loss of methylation, combined with the addition of acetyl groups results in a more accessible chromatin conformation with a propensity towards disorganization when compared to mitotically active cells. [50] General loss of histones precludes addition of histone modifications and contributes changes in enrichment in some chromatin regions during senescence. [51]

See also

Related Research Articles

<span class="mw-page-title-main">Histone</span> Family proteins package and order the DNA into structural units called nucleosomes.

In biology, histones are highly basic proteins abundant in lysine and arginine residues that are found in eukaryotic cell nuclei. They act as spools around which DNA winds to create structural units called nucleosomes. Nucleosomes in turn are wrapped into 30-nanometer fibers that form tightly packed chromatin. Histones prevent DNA from becoming tangled and protect it from DNA damage. In addition, histones play important roles in gene regulation and DNA replication. Without histones, unwound DNA in chromosomes would be very long. For example, each human cell has about 1.8 meters of DNA if completely stretched out; however, when wound about histones, this length is reduced to about 90 micrometers (0.09 mm) of 30 nm diameter chromatin fibers.

<span class="mw-page-title-main">Nucleosome</span> Basic structural unit of DNA packaging in eukaryotes

A nucleosome is the basic structural unit of DNA packaging in eukaryotes. The structure of a nucleosome consists of a segment of DNA wound around eight histone proteins and resembles thread wrapped around a spool. The nucleosome is the fundamental subunit of chromatin. Each nucleosome is composed of a little less than two turns of DNA wrapped around a set of eight proteins called histones, which are known as a histone octamer. Each histone octamer is composed of two copies each of the histone proteins H2A, H2B, H3, and H4.

RSC is a member of the ATP-dependent chromatin remodeler family. The activity of the RSC complex allows for chromatin to be remodeled by altering the structure of the nucleosome.

<span class="mw-page-title-main">SWI/SNF</span> Subfamily of ATP-dependent chromatin remodeling complexes

In molecular biology, SWI/SNF, is a subfamily of ATP-dependent chromatin remodeling complexes, which is found in eukaryotes. In other words, it is a group of proteins that associate to remodel the way DNA is packaged. This complex is composed of several proteins – products of the SWI and SNF genes, as well as other polypeptides. It possesses a DNA-stimulated ATPase activity that can destabilize histone-DNA interactions in reconstituted nucleosomes in an ATP-dependent manner, though the exact nature of this structural change is unknown. The SWI/SNF subfamily provides crucial nucleosome rearrangement, which is seen as ejection and/or sliding. The movement of nucleosomes provides easier access to the chromatin, allowing genes to be activated or repressed.

<span class="mw-page-title-main">Histone-modifying enzymes</span> Type of enzymes

Histone-modifying enzymes are enzymes involved in the modification of histone substrates after protein translation and affect cellular processes including gene expression. To safely store the eukaryotic genome, DNA is wrapped around four core histone proteins, which then join to form nucleosomes. These nucleosomes further fold together into highly condensed chromatin, which renders the organism's genetic material far less accessible to the factors required for gene transcription, DNA replication, recombination and repair. Subsequently, eukaryotic organisms have developed intricate mechanisms to overcome this repressive barrier imposed by the chromatin through histone modification, a type of post-translational modification which typically involves covalently attaching certain groups to histone residues. Once added to the histone, these groups elicit either a loose and open histone conformation, euchromatin, or a tight and closed histone conformation, heterochromatin. Euchromatin marks active transcription and gene expression, as the light packing of histones in this way allows entry for proteins involved in the transcription process. As such, the tightly packed heterochromatin marks the absence of current gene expression.

<span class="mw-page-title-main">SMARCA4</span> Protein-coding gene in the species Homo sapiens

Transcription activator BRG1 also known as ATP-dependent chromatin remodeler SMARCA4 is a protein that in humans is encoded by the SMARCA4 gene.

<span class="mw-page-title-main">SMARCA5</span> Protein-coding gene in the species Homo sapiens

SWI/SNF-related matrix-associated actin-dependent regulator of chromatin subfamily A member 5 is a protein that in humans is encoded by the SMARCA5 gene.

<span class="mw-page-title-main">ARID1A</span> Protein-coding gene in humans

AT-rich interactive domain-containing protein 1A is a protein that in humans is encoded by the ARID1A gene.

<span class="mw-page-title-main">CHD3</span> Protein-coding gene in the species Homo sapiens

Chromodomain-helicase-DNA-binding protein 3 is an enzyme that in humans is encoded by the CHD3 gene.

<span class="mw-page-title-main">CHD1</span> Chromatin remodeling protein that is widely conserved across many eukaryotic organisms

The Chromodomain-Helicase DNA-binding 1 is a protein that, in humans, is encoded by the CHD1 gene. CHD1 is a chromatin remodeling protein that is widely conserved across many eukaryotic organisms, from yeast to humans. CHD1 is named for three of its protein domains: two tandem chromodomains, its ATPase catalytic domain, and its DNA-binding domain.

Epigenomics is the study of the complete set of epigenetic modifications on the genetic material of a cell, known as the epigenome. The field is analogous to genomics and proteomics, which are the study of the genome and proteome of a cell. Epigenetic modifications are reversible modifications on a cell's DNA or histones that affect gene expression without altering the DNA sequence. Epigenomic maintenance is a continuous process and plays an important role in stability of eukaryotic genomes by taking part in crucial biological mechanisms like DNA repair. Plant flavones are said to be inhibiting epigenomic marks that cause cancers. Two of the most characterized epigenetic modifications are DNA methylation and histone modification. Epigenetic modifications play an important role in gene expression and regulation, and are involved in numerous cellular processes such as in differentiation/development and tumorigenesis. The study of epigenetics on a global level has been made possible only recently through the adaptation of genomic high-throughput assays.

ISWI is one of the five major DNA chromatin remodeling complex types, or subfamilies, found in most eukaryotic organisms. ISWI remodeling complexes place nucleosomes along segments of DNA at regular intervals. The placement of nucleosomes by ISWI protein complexes typically results in the silencing of the DNA because the nucleosome placement prevents transcription of the DNA. ISWI, like the closely related SWI/SNF subfamily, is an ATP-dependent chromatin remodeler. However, the chromatin remodeling activities of ISWI and SWI/SNF are distinct and mediate the binding of non-overlapping sets of DNA transcription factors.

In the field of molecular biology, the Mi-2/NuRDcomplex, is a group of associated proteins with both ATP-dependent chromatin remodeling and histone deacetylase activities. As of 2007, Mi-2/NuRD was the only known protein complex that couples chromatin remodeling ATPase and chromatin deacetylation enzymatic functions.

<span class="mw-page-title-main">Biomarkers of aging</span> Type of biomarkers

Biomarkers of aging are biomarkers that could predict functional capacity at some later age better than chronological age. Stated another way, biomarkers of aging would give the true "biological age", which may be different from the chronological age.

H3K4me3 is an epigenetic modification to the DNA packaging protein Histone H3 that indicates tri-methylation at the 4th lysine residue of the histone H3 protein and is often involved in the regulation of gene expression. The name denotes the addition of three methyl groups (trimethylation) to the lysine 4 on the histone H3 protein.

H3K27me3 is an epigenetic modification to the DNA packaging protein Histone H3. It is a mark that indicates the tri-methylation of lysine 27 on histone H3 protein.

H4K20me is an epigenetic modification to the DNA packaging protein Histone H4. It is a mark that indicates the mono-methylation at the 20th lysine residue of the histone H4 protein. This mark can be di- and tri-methylated. It is critical for genome integrity including DNA damage repair, DNA replication and chromatin compaction.

H4K16ac is an epigenetic modification to the DNA packaging protein Histone H4. It is a mark that indicates the acetylation at the 16th lysine residue of the histone H4 protein.

Robert E. Kingston is an American biochemist who studies the functional and regulatory role nucleosomes play in gene expression, specifically during early development. After receiving his PhD (1981) and completing post-doctoral research, Kingston became an assistant professor at Massachusetts General Hospital (1985), where he started a research laboratory focused on understanding chromatin's structure with regards to transcriptional regulation. As a Harvard graduate himself, Kingston has served his alma mater through his leadership.

Transgenerational epigenetic inheritance in plants involves mechanisms for the passing of epigenetic marks from parent to offspring that differ from those reported in animals. There are several kinds of epigenetic markers, but they all provide a mechanism to facilitate greater phenotypic plasticity by influencing the expression of genes without altering the DNA code. These modifications represent responses to environmental input and are reversible changes to gene expression patterns that can be passed down through generations. In plants, transgenerational epigenetic inheritance could potentially represent an evolutionary adaptation for sessile organisms to quickly adapt to their changing environment.

References

  1. Teif VB, Rippe K (September 2009). "Predicting nucleosome positions on the DNA: combining intrinsic sequence preferences and remodeler activities". Nucleic Acids Research. 37 (17): 5641–55. doi:10.1093/nar/gkp610. PMC   2761276 . PMID   19625488.
  2. Boyer (2009). "The chromatin signature of pluripotent cells". Stembook. doi: 10.3824/stembook.1.45.1 . PMID   20614601.
  3. Wang GG, Allis CD, Chi P (September 2007). "Chromatin remodeling and cancer, Part I: Covalent histone modifications". Trends in Molecular Medicine. 13 (9): 363–72. doi:10.1016/j.molmed.2007.07.003. PMID   17822958.
  4. Strahl BD, Allis CD (January 2000). "The language of covalent histone modifications". Nature. 403 (6765): 41–5. Bibcode:2000Natur.403...41S. doi:10.1038/47412. PMID   10638745. S2CID   4418993.
  5. 1 2 3 4 Rosenfeld JA, Wang Z, Schones DE, Zhao K, DeSalle R, Zhang MQ (March 2009). "Determination of enriched histone modifications in non-genic portions of the human genome". BMC Genomics. 10: 143. doi:10.1186/1471-2164-10-143. PMC   2667539 . PMID   19335899.
  6. Hublitz P, Albert M, Peters A (28 April 2009). "Mechanisms of Transcriptional Repression by Histone Lysine Methylation". The International Journal of Developmental Biology. 10 (1387): 335–354. doi: 10.1387/ijdb.082717ph . ISSN   1696-3547. PMID   19412890.
  7. Tan M, Luo H, Lee S, Jin F, Yang JS, Montellier E, Buchou T, Cheng Z, Rousseaux S, Rajagopal N, Lu Z, Ye Z, Zhu Q, Wysocka J, Ye Y, Khochbin S, Ren B, Zhao Y (September 2011). "Identification of 67 histone marks and histone lysine crotonylation as a new type of histone modification". Cell. 146 (6): 1016–28. doi:10.1016/j.cell.2011.08.008. PMC   3176443 . PMID   21925322.
  8. Jenuwein T, Allis CD (August 2001). "Translating the histone code". Science. 293 (5532): 1074–80. CiteSeerX   10.1.1.453.900 . doi:10.1126/science.1063127. PMID   11498575. S2CID   1883924.
  9. Benevolenskaya EV (August 2007). "Histone H3K4 demethylases are essential in development and differentiation". Biochemistry and Cell Biology. 85 (4): 435–43. doi:10.1139/o07-057. PMID   17713579.
  10. 1 2 3 4 5 6 7 8 Barski A, Cuddapah S, Cui K, Roh TY, Schones DE, Wang Z, Wei G, Chepelev I, Zhao K (May 2007). "High-resolution profiling of histone methylations in the human genome". Cell. 129 (4): 823–37. doi: 10.1016/j.cell.2007.05.009 . PMID   17512414. S2CID   6326093.
  11. 1 2 3 Steger DJ, Lefterova MI, Ying L, Stonestrom AJ, Schupp M, Zhuo D, Vakoc AL, Kim JE, Chen J, Lazar MA, Blobel GA, Vakoc CR (April 2008). "DOT1L/KMT4 recruitment and H3K79 methylation are ubiquitously coupled with gene transcription in mammalian cells". Molecular and Cellular Biology. 28 (8): 2825–39. doi:10.1128/MCB.02076-07. PMC   2293113 . PMID   18285465.
  12. 1 2 3 Koch CM, Andrews RM, Flicek P, Dillon SC, Karaöz U, Clelland GK, Wilcox S, Beare DM, Fowler JC, Couttet P, James KD, Lefebvre GC, Bruce AW, Dovey OM, Ellis PD, Dhami P, Langford CF, Weng Z, Birney E, Carter NP, Vetrie D, Dunham I (June 2007). "The landscape of histone modifications across 1% of the human genome in five human cell lines". Genome Research. 17 (6): 691–707. doi:10.1101/gr.5704207. PMC   1891331 . PMID   17567990.
  13. 1 2 Wang GG, Allis CD, Chi P (September 2007). "Chromatin remodeling and cancer, Part II: ATP-dependent chromatin remodeling". Trends in Molecular Medicine. 13 (9): 373–80. doi:10.1016/j.molmed.2007.07.004. PMC   4337864 . PMID   17822959.
  14. Saha A, Wittmeyer J, Cairns BR (June 2006). "Chromatin remodelling: the industrial revolution of DNA around histones". Nature Reviews Molecular Cell Biology. 7 (6): 437–47. doi:10.1038/nrm1945. PMID   16723979. S2CID   6180120.
  15. Vignali, M.; Hassan, A. H.; Neely, K. E.; Workman, J. L. (2000-03-15). "ATP-Dependent Chromatin-Remodeling Complexes". Molecular and Cellular Biology. 20 (6): 1899–1910. doi: 10.1128/mcb.20.6.1899-1910.2000 . ISSN   0270-7306. PMC   110808 . PMID   10688638.
  16. 1 2 3 4 Sellou H, Lebeaupin T, Chapuis C, Smith R, Hegele A, Singh HR, Kozlowski M, Bultmann S, Ladurner AG, Timinszky G, Huet S (2016). "The poly(ADP-ribose)-dependent chromatin remodeler Alc1 induces local chromatin relaxation upon DNA damage". Mol. Biol. Cell. 27 (24): 3791–3799. doi:10.1091/mbc.E16-05-0269. PMC   5170603 . PMID   27733626.
  17. 1 2 Haince JF, McDonald D, Rodrigue A, Déry U, Masson JY, Hendzel MJ, Poirier GG (2008). "PARP1-dependent kinetics of recruitment of MRE11 and NBS1 proteins to multiple DNA damage sites". J. Biol. Chem. 283 (2): 1197–208. doi: 10.1074/jbc.M706734200 . PMID   18025084.
  18. 1 2 3 Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM (1998). "DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139". J. Biol. Chem. 273 (10): 5858–68. doi: 10.1074/jbc.273.10.5858 . PMID   9488723.
  19. Mailand N, Bekker-Jensen S, Faustrup H, Melander F, Bartek J, Lukas C, Lukas J (2007). "RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins". Cell. 131 (5): 887–900. doi: 10.1016/j.cell.2007.09.040 . PMID   18001824. S2CID   14232192.
  20. Stucki M, Clapperton JA, Mohammad D, Yaffe MB, Smerdon SJ, Jackson SP (2005). "MDC1 directly binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks". Cell. 123 (7): 1213–26. doi: 10.1016/j.cell.2005.09.038 . PMID   16377563.
  21. Chapman JR, Jackson SP (2008). "Phospho-dependent interactions between NBS1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage". EMBO Rep. 9 (8): 795–801. doi:10.1038/embor.2008.103. PMC   2442910 . PMID   18583988.
  22. Luijsterburg MS, Acs K, Ackermann L, Wiegant WW, Bekker-Jensen S, Larsen DH, Khanna KK, van Attikum H, Mailand N, Dantuma NP (2012). "A new non-catalytic role for ubiquitin ligase RNF8 in unfolding higher-order chromatin structure". EMBO J. 31 (11): 2511–27. doi:10.1038/emboj.2012.104. PMC   3365417 . PMID   22531782.
  23. Smeenk G, Wiegant WW, Vrolijk H, Solari AP, Pastink A, van Attikum H (2010). "The NuRD chromatin-remodeling complex regulates signaling and repair of DNA damage". J. Cell Biol. 190 (5): 741–9. doi:10.1083/jcb.201001048. PMC   2935570 . PMID   20805320.
  24. Hanahan D, Weinberg RA (January 2000). "The hallmarks of cancer". Cell. 100 (1): 57–70. doi: 10.1016/S0092-8674(00)81683-9 . PMID   10647931. S2CID   1478778.
  25. Versteege I, Sévenet N, Lange J, Rousseau-Merck MF, Ambros P, Handgretinger R, Aurias A, Delattre O (July 1998). "Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer". Nature. 394 (6689): 203–6. Bibcode:1998Natur.394..203V. doi:10.1038/28212. PMID   9671307. S2CID   6019090.
  26. Shain AH, Pollack JR (2013). "The spectrum of SWI/SNF mutations, ubiquitous in human cancers". PLOS ONE. 8 (1): e55119. Bibcode:2013PLoSO...855119S. doi: 10.1371/journal.pone.0055119 . PMC   3552954 . PMID   23355908.
  27. 1 2 Hodges C, Kirkland JG, Crabtree GR (August 2016). "The Many Roles of BAF (mSWI/SNF) and PBAF Complexes in Cancer". Cold Spring Harbor Perspectives in Medicine. 6 (8): a026930. doi:10.1101/cshperspect.a026930. PMC   4968166 . PMID   27413115.
  28. Medina PP, Romero OA, Kohno T, Montuenga LM, Pio R, Yokota J, Sanchez-Cespedes M (May 2008). "Frequent BRG1/SMARCA4-inactivating mutations in human lung cancer cell lines". Human Mutation. 29 (5): 617–22. doi: 10.1002/humu.20730 . PMID   18386774. S2CID   8596785.
  29. 1 2 3 Hodges HC, Stanton BZ, Cermakova K, Chang CY, Miller EL, Kirkland JG, Ku WL, Veverka V, Zhao K, Crabtree GR (January 2018). "Dominant-negative SMARCA4 mutants alter the accessibility landscape of tissue-unrestricted enhancers". Nature Structural & Molecular Biology. 25 (1): 61–72. doi:10.1038/s41594-017-0007-3. PMC   5909405 . PMID   29323272.
  30. 1 2 Stanton BZ, Hodges C, Calarco JP, Braun SM, Ku WL, Kadoch C, Zhao K, Crabtree GR (February 2017). "Smarca4 ATPase mutations disrupt direct eviction of PRC1 from chromatin". Nature Genetics. 49 (2): 282–288. doi:10.1038/ng.3735. PMC   5373480 . PMID   27941795.
  31. Baliñas-Gavira, Carlos; Rodríguez, María I.; Andrades, Alvaro; Cuadros, Marta; Álvarez-Pérez, Juan Carlos; Álvarez-Prado, Ángel F.; de Yébenes, Virginia G.; Sánchez-Hernández, Sabina; Fernández-Vigo, Elvira; Muñoz, Javier; Martín, Francisco; Ramiro, Almudena R.; Martínez-Climent, José A.; Medina, Pedro P. (October 2020). "Frequent mutations in the amino-terminal domain of BCL7A impair its tumor suppressor role in DLBCL". Leukemia. 34 (10): 2722–2735. doi:10.1038/s41375-020-0919-5. ISSN   1476-5551. PMID   32576963.
  32. Andrades, Alvaro; Peinado, Paola; Alvarez-Perez, Juan Carlos; Sanjuan-Hidalgo, Juan; García, Daniel J.; Arenas, Alberto M.; Matia-González, Ana M.; Medina, Pedro P. (2023-02-21). "SWI/SNF complexes in hematological malignancies: biological implications and therapeutic opportunities". Molecular Cancer. 22 (1): 39. doi:10.1186/s12943-023-01736-8. ISSN   1476-4598. PMC   9942420 . PMID   36810086.
  33. Liquori, Alessandro; Ibañez, Mariam; Sargas, Claudia; Sanz, Miguel Ángel; Barragán, Eva; Cervera, José (2020-03-08). "Acute Promyelocytic Leukemia: A Constellation of Molecular Events around a Single PML-RARA Fusion Gene". Cancers. 12 (3): 624. doi: 10.3390/cancers12030624 . ISSN   2072-6694. PMC   7139833 . PMID   32182684.
  34. Wolffe AP (May 2001). "Chromatin remodeling: why it is important in cancer". Oncogene. 20 (24): 2988–90. doi:10.1038/sj.onc.1204322. PMID   11420713.
  35. Tu, William B.; Helander, Sara; Pilstål, Robert; Hickman, K. Ashley; Lourenco, Corey; Jurisica, Igor; Raught, Brian; Wallner, Björn; Sunnerhagen, Maria; Penn, Linda Z. (May 2015). "Myc and its interactors take shape". Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms. 1849 (5): 469–483. doi:10.1016/j.bbagrm.2014.06.002. ISSN   0006-3002. PMID   24933113.
  36. Marks PA, Dokmanovic M (December 2005). "Histone deacetylase inhibitors: discovery and development as anticancer agents". Expert Opinion on Investigational Drugs. 14 (12): 1497–511. doi:10.1517/13543784.14.12.1497. PMID   16307490. S2CID   1235026.
  37. Richon VM, O'Brien JP (2002). "Histone deacetylase inhibitors: a new class of potential therapeutic agents for cancer treatment" (PDF). Clinical Cancer Research. 8 (3): 662–4. PMID   11895892.
  38. Richon VM, Sandhoff TW, Rifkind RA, Marks PA (August 2000). "Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation". Proceedings of the National Academy of Sciences of the United States of America. 97 (18): 10014–9. Bibcode:2000PNAS...9710014R. doi: 10.1073/pnas.180316197 . PMC   27656 . PMID   10954755.
  39. Zhang Z, Yamashita H, Toyama T, Sugiura H, Ando Y, Mita K, Hamaguchi M, Hara Y, Kobayashi S, Iwase H (November 2005). "Quantitation of HDAC1 mRNA expression in invasive carcinoma of the breast*". Breast Cancer Research and Treatment. 94 (1): 11–6. doi:10.1007/s10549-005-6001-1. PMID   16172792. S2CID   27550683.
  40. Munshi, Anupama; Tanaka, Toshimitsu; Hobbs, Marvette L.; Tucker, Susan L.; Richon, Victoria M.; Meyn, Raymond E. (August 2006). "Vorinostat, a histone deacetylase inhibitor, enhances the response of human tumor cells to ionizing radiation through prolongation of gamma-H2AX foci". Molecular Cancer Therapeutics. 5 (8): 1967–1974. doi:10.1158/1535-7163.MCT-06-0022. ISSN   1535-7163. PMID   16928817. S2CID   26874948.
  41. "Zolinza® (vorinostat) Capsules. Full Prescribing Information" (PDF). Food and Drug Administration (FDA). Archived from the original (PDF) on 20 January 2022. Retrieved 9 February 2023.
  42. VanderMolen, Karen M.; McCulloch, William; Pearce, Cedric J.; Oberlies, Nicholas H. (August 2011). "Romidepsin (Istodax, NSC 630176, FR901228, FK228, depsipeptide): a natural product recently approved for cutaneous T-cell lymphoma". The Journal of Antibiotics. 64 (8): 525–531. doi:10.1038/ja.2011.35. ISSN   1881-1469. PMC   3163831 . PMID   21587264.
  43. "ISTODAX® (romidepsin) for injection, for intravenous use. Full prescribing information" (PDF). Food and Drug Administration (FDA). Archived from the original (PDF) on 25 January 2022. Retrieved 9 February 2023.
  44. Dowden J, Hong W, Parry RV, Pike RA, Ward SG (April 2010). "Toward the development of potent and selective bisubstrate inhibitors of protein arginine methyltransferases". Bioorganic & Medicinal Chemistry Letters. 20 (7): 2103–5. doi:10.1016/j.bmcl.2010.02.069. PMID   20219369.
  45. Wong, Lee H.; McGhie, James D.; Sim, Marcus; Anderson, Melissa A.; Ahn, Soyeon; Hannan, Ross D.; George, Amee J.; Morgan, Kylie A.; Mann, Jeffrey R.; Choo, K. H. Andy (March 2010). "ATRX interacts with H3.3 in maintaining telomere structural integrity in pluripotent embryonic stem cells". Genome Research. 20 (3): 351–360. doi:10.1101/gr.101477.109. ISSN   1549-5469. PMC   2840985 . PMID   20110566.
  46. Clapier CR, Cairns BR (2009). "The biology of chromatin remodeling complexes". Annual Review of Biochemistry. 78: 273–304. doi:10.1146/annurev.biochem.77.062706.153223. PMID   19355820.
  47. 1 2 Parry, Aled John; Narita, Masashi (2016). "Old cells, new tricks: chromatin structure in senescence". Mammalian Genome. 27 (7–8): 320–331. doi:10.1007/s00335-016-9628-9. ISSN   0938-8990. PMC   4935760 . PMID   27021489.
  48. Hayflick, L.; Moorhead, P. S. (1961-12-01). "The serial cultivation of human diploid cell strains". Experimental Cell Research. 25 (3): 585–621. doi:10.1016/0014-4827(61)90192-6. ISSN   0014-4827. PMID   13905658.
  49. 1 2 3 Chandra, Tamir; Ewels, Philip Andrew; Schoenfelder, Stefan; Furlan-Magaril, Mayra; Wingett, Steven William; Kirschner, Kristina; Thuret, Jean-Yves; Andrews, Simon; Fraser, Peter; Reik, Wolf (2015-01-29). "Global Reorganization of the Nuclear Landscape in Senescent Cells". Cell Reports. 10 (4): 471–483. doi:10.1016/j.celrep.2014.12.055. ISSN   2211-1247. PMC   4542308 . PMID   25640177.
  50. 1 2 3 4 5 6 Sun, Luyang; Yu, Ruofan; Dang, Weiwei (2018-04-16). "Chromatin Architectural Changes during Cellular Senescence and Aging". Genes. 9 (4): 211. doi: 10.3390/genes9040211 . ISSN   2073-4425. PMC   5924553 . PMID   29659513.
  51. 1 2 3 4 Criscione, Steven W.; Teo, Yee Voan; Neretti, Nicola (2016). "The chromatin landscape of cellular senescence". Trends in Genetics. 32 (11): 751–761. doi:10.1016/j.tig.2016.09.005. ISSN   0168-9525. PMC   5235059 . PMID   27692431.
  52. 1 2 3 4 Yang, Na; Sen, Payel (2018-11-03). "The senescent cell epigenome". Aging (Albany NY). 10 (11): 3590–3609. doi:10.18632/aging.101617. ISSN   1945-4589. PMC   6286853 . PMID   30391936.
  53. 1 2 Basta, Jeannine; Rauchman, Michael (2015). "The Nucleosome Remodeling and Deacetylase (NuRD) Complex in Development and Disease". Translational Research. 165 (1): 36–47. doi:10.1016/j.trsl.2014.05.003. ISSN   1931-5244. PMC   4793962 . PMID   24880148.
  54. 1 2 Li, Xueping; Ding, Dong; Yao, Jun; Zhou, Bin; Shen, Ting; Qi, Yun; Ni, Ting; Wei, Gang (2019-07-15). "Chromatin remodeling factor BAZ1A regulates cellular senescence in both cancer and normal cells". Life Sciences. 229: 225–232. doi:10.1016/j.lfs.2019.05.023. ISSN   1879-0631. PMID   31085244. S2CID   155090903.
  55. 1 2 López-Otín, Carlos; Blasco, Maria A.; Partridge, Linda; Serrano, Manuel; Kroemer, Guido (2013-06-06). "The Hallmarks of Aging". Cell. 153 (6): 1194–1217. doi:10.1016/j.cell.2013.05.039. ISSN   0092-8674. PMC   3836174 . PMID   23746838.
  56. 1 2 Tasdemir, Nilgun; Banito, Ana; Roe, Jae-Seok; Alonso-Curbelo, Direna; Camiolo, Matthew; Tschaharganeh, Darjus F.; Huang, Chun-Hao; Aksoy, Ozlem; Bolden, Jessica E.; Chen, Chi-Chao; Fennell, Myles (2016). "BRD4 Connects Enhancer Remodeling to Senescence Immune Surveillance". Cancer Discovery. 6 (6): 612–629. doi:10.1158/2159-8290.CD-16-0217. ISSN   2159-8290. PMC   4893996 . PMID   27099234.
  57. Wilson, V. L.; Jones, P. A. (1983-06-03). "DNA methylation decreases in aging but not in immortal cells". Science. 220 (4601): 1055–1057. Bibcode:1983Sci...220.1055W. doi:10.1126/science.6844925. ISSN   0036-8075. PMID   6844925.
  58. Kaeberlein, Matt; McVey, Mitch; Guarente, Leonard (1999-10-01). "The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms". Genes & Development. 13 (19): 2570–2580. doi:10.1101/gad.13.19.2570. ISSN   0890-9369. PMC   317077 . PMID   10521401.

Further reading