Open access peer-reviewed chapter

Transcription Elongation Factors in Health and Disease

Written By

Preeti Dabas

Submitted: 26 December 2021 Reviewed: 02 February 2022 Published: 08 March 2022

DOI: 10.5772/intechopen.103013

From the Edited Volume

Gene Expression

Edited by Fumiaki Uchiumi

Chapter metrics overview

250 Chapter Downloads

View Full Metrics

Abstract

Gene expression is a complex process that establishes and maintains a specific cell state. Transcription, an early event during the gene expression, is fine-tuned by a concerted action of a plethora of transcription factors temporally and spatially in response to various stimuli. Most of the earlier research has focused on the initiation of transcription as a key regulatory step. However, work done over the last two decades has highlighted the importance of regulation of transcription elongation by RNA Pol II in the implementation of gene expression programs during development. Moreover, accumulating evidence has suggested that dysregulation of transcription elongation due to dysfunction of transcription factors can result in developmental abnormalities and a broad range of diseases, including cancers. In this chapter, we review recent advances in our understanding of the dynamics of transcription regulation during the elongation stage, the significance of transcriptional regulatory complexes, and their relevance in the development of potential accurate therapeutic targets for different human diseases.

Keywords

  • transcription
  • RNA Pol II
  • elongation factor

1. Introduction

There are ~20,000 protein-coding genes in the human genome [1]. Cells modulate the expression of these genes in spatial and temporal manner in response to various stimuli. Gene expression is a highly regulated and complex process, which begins with the opening of chromatin, the transcription of the primary RNA transcript (hnRNA) from DNA, followed by processing of the hnRNA into mRNA, which is then translated into a protein that dictates cell functions. There has been an extensive study on precise control of gene expression at different stages by a plethora of factors leading to the concept of “gene-class specific” or selective gene control [2, 3, 4, 5, 6]. Tight control of gene expression is indispensable for normal cellular functions, and any dysregulation may lead to a wide range of diseases. The recent surge in knowledge and understanding of diseases that are caused by a mutation in regulatory sequences, transcription factors, cofactors, chromatin regulators, and non-coding RNA, such as diabetes, autoimmune disorders, neurological disorders, obesity, cardiovascular disease, and cancer, has altered our view of the underlying cause and primary focus of therapeutic targets.

It is imperative to understand the process of regulated gene expression to get insights into the mechanisms involved in the dysregulation of gene expression in various human diseases and to develop potential therapeutic targets for these diseases. Transcription has been considered the most important rate-limiting step during the expression of a gene. For a long time, initiation of transcription was considered as the key regulatory event during transcription and thus was the focus of major research in this field. However, for over a decade, the focus of research has shifted toward other steps during transcription, such as elongation and termination. Moreover, the regulation of transcription elongation has emerged to be the center of most therapeutic studies targeting fine-tuning of gene expression in diseases such as cancers.

In this chapter, I begin with a brief review of steps involved in gene regulation and the fundamentals of transcriptional control of gene expression. I further focus on a detailed understanding of regulation of transcription elongation by different transcription elongation factors and complexes and how a mutation or dysfunction of any of these factors contributes to altered transcription leading to progression of various diseases and cancer. I also highlight the recent advances in the development of precision tailored therapeutics by mediating transcriptional control of a gene.

Advertisement

2. Stages of regulation of gene expression

In eukaryotes, modulation of gene expression occurs at seven different steps (Figure 1).

Figure 1.

Schematic representation of steps involved in gene regulation. Regulation of gene expression occurs at seven different steps.

2.1 Chromatin structure

DNA wraps around proteins called histones to form nucleosomes. Each nucleosome is further condensed into chromatin. The condensation of eukaryotic DNA in chromatin acts to suppress the expression of genes by acting as a physical barrier to the transcription machinery [2]. The opening of chromatin, which allows access to genomic DNA, is indispensable for gene expression and formation of RNA from DNA template. This unwrapping of DNA from histone proteins is called chromatin remodeling and is carried out by enzymes that interact with histones and covalently attach functional groups to the amino terminal tail of histones. These histone-modifying proteins form complexes known as chromatin remodeling complexes. The most common modifications include methylation or acetylation of lysine residues on histone tail [3]. The outcome of these two modifications is entirely different: acetylation results in an open conformation of chromatin, thereby causing activation of gene expression; methylation results in a more compact chromatin conformation, hindering DNA accessibility to transcription factors and thus repressing transcription. Apart from histone modifications, methylation of DNA may also lead to a transcriptionally inactive state. A balance between the active and inactive state of chromatin is of profound importance for the maintenance of a healthy cellular environment. Dysfunctional chromatin remodeling complexes have been implicated in several disease conditions, including Williams syndrome [4, 5], Rett syndrome [6] breast cancer [7], and several other primary tumors [8].

2.2 Transcription

Transcription is the key regulatory step for gene expression in eukaryotes. It involves a concerted action of different proteins, such as transcription factors, mediator complex components, and RNA Pol II to produce RNA using DNA as a template. In eukaryotes, RNA Pol II is responsible for the synthesis of protein-coding genes as well as some non-coding RNAs such as small nuclear RNA (snRNA), microRNA (miRNA), cryptic unstable transcripts (CUTs), small nucleolar RNA (SnoRNA), and stable unannotated transcripts (SUTs) [9]. RNA Pol II mediated transcription is composed of three main steps: initiation, elongation, and termination, all of which are subjected to regulatory controls. During the stage of transcription initiation, the RNA Pol II in association with different transcription factors is recruited to the promoter region of the gene and forms a complex called pre-initiation complex (PIC). This opens the DNA, and the template strand positions itself in the active site of RNA Pol II, which then initiates the synthesis of the first few nucleotides of RNA. When the RNA length reaches ~10 nucleotides, RNA Pol II escapes the promoter and enters the gene body leading to productive elongation. Once Pol II reaches the end of the gene, RNA Pol II ceases RNA synthesis, which signals the release of the nascent RNA transcript as well as Pol II from the DNA template, thus terminating the process of transcription. All three stages of transcription are subject to tight control. Perhaps, due to its foremost position in the transcription cycle, the initiation step is extensively studied for mechanism and regulation. More recently, the transition of initiation to elongation has emerged to be the hub of major research in the field of transcription regulation. However, the mechanism and regulation of transcription termination have been less investigated.

A detailed description of regulation during transcription is described in the subsequent sections of this chapter.

2.3 RNA processing

The primary or nascent transcript is further processed to form functional mRNA. During processing, the primary transcript undergoes three types of modifications: 5′ end modification (capping), removal of introns (splicing), and 3′ end modification (polyadenylation) [10].

The newly synthesized RNA is stabilized by the addition of a 7-methylguanosine cap, which protects nascent from attack by nuclear exonucleases and helps in promoting transcription, splicing, polyadenylation, and nuclear export [11, 12]. Factors responsible for the capping of 5′ end of RNA, for example, eIF4E-antisense oligonucleotides, are being extensively used therapeutically in clinical trials that aim to curb dysregulated gene expression in cancer [13]. Small ribonucleoprotein particles (snRNPs) along with auxiliary proteins form a spliceosome complex, which mainly carries out the process of splicing by recognizing the splice sites and catalyzing the splice reaction [14]. Any dysregulation of the splicing mechanism results in diseased conditions [15]. Polyadenylation of 3’end of nascent RNA includes cleavage at polyadenylation site (PAS) of RNA and addition of poly (A) tail [16, 17, 18]. Alternate polyadenylation (APA) is yet another mechanism adopted by the cell to produce diversity in the mRNA pool. APA results in different isoforms of the same gene with varying 3’UTR [19]. Poly(A) tails are responsible for stability, translation efficiency, and degradation of RNA. Alteration in polyadenylation is associated with a plethora of diseased conditions, such as neonatal diabetes, fragile X-associated premature ovarian insufficiency, IPEX (immune dysfunction, polyendocrinopathy, enteropathy, X-linked), ectopic Cushing syndrome, and several cancer conditions such as endocrine tumor [20].

2.4 RNA transport

Once the mature RNA is made post RNA processing events, it is rearranged in an export competent mRNP complex with RNA-binding factors and shuttling proteins and transported to the cytoplasm. This process is tightly regulated. After the transport, in the cytoplasmic side, the DEAD-box helicase remodels the mRNP to dissociate RNA binding and shuttling proteins, preventing the mRNA from returning to the nucleus [21]. Furthermore, cap-binding protein (CBP), which binds to the 7-methylguanylate cap on 5′ end of processed RNA, is recognized by nuclear pore complex and exported to the cytoplasm, where it is replaced by translation factors eIF4E and eIF4G [21]. Few lines of research have shown that the transport machinery co-transcriptionally associates with mRNA [21, 22, 23], while others have shown that 3′ end processing of transcript marks the event responsible for the loading of export complex [24, 25]. For example, shuttling proteins, Hrp1p and Nab2p, associate with poly (A) tail [26, 27].

Any defect in mRNA transport or nuclear retention results in human disease such as lethal congenital contracture syndrome 1.

2.5 RNA stability and degradation

The life span of mRNA in cytosol determines the protein turnover. This is an important step to control gene expression since modulation of mRNA abundance allows cell to adapt and respond to various situations adequately. Generally, proteins with housekeeping functions are encoded by mRNA with a long half-life, while the proteins required only at specific developmental stages are encoded by mRNA with a short half-life [28, 29]. mRNA decay is a highly regulated process, resulting from interactions between mRNA, RNA-binding proteins, non-coding RNA, and various decay factors. The stringency of mRNA degradation depends on the presence of regulatory RNA elements, consisting of specific sequences found anywhere in mRNA, including 5′ and 3’ UTR. These sequences are recognized by RNA-binding proteins, forming mRNP complexes [30]. These RNA-binding proteins determine the fate of bound mRNA—to be translated, decayed, or stored in untranslated form as cytoplasmic granules. mRNA degradation begins with deadenylation or shortening of poly(A) tail carried out by Pan2/Pan3 complex and Ccr4/Pop2/Not complexes. After deadenylation, the mRNA is degraded either by 3′ → 5’ mRNA decay pathway mediated by exosome or 5′ → 3’ mRNA decay pathway mediated by exonuclease Xrn1 after decapping by Dcp1/Dcp2 decapping complex [31]. More recent studies have focused on the role of non-coding RNA molecules called miRNA in regulation of mRNA degradation and subsequently gene expression. miRNA works by either repressing translation or promoting degradation of mRNA having sequence complementary to miRNA [32]. Dysregulation of mRNA stability has been implicated in several diseases, including tumors. For instance, in myeloma and human T-cell leukemia, the stability of c-Myc RNA (an oncogene) due to loss of 3’UTR, which is responsible for its decay, results in its stability up to 4–8-fold higher compared with the wild type [33, 34].

2.6 Translation

Regulation of translation is a crucial mechanism for spatial control of gene expression [35]. There has been an explosion of studies highlighting the importance of regulation of translation in various physiological processes such as in normal development [36], in apoptosis [37], in stress response [38], etc. Dysregulation of translation has been implicated in different cancers [39]. In cells lacking nucleus, such as neurons, regulation of gene expression by translation has been shown to be of utmost importance. Several studies have demonstrated the initiation of translation as one of the crucial regulatory steps of protein synthesis [38, 40, 41, 42]. The role of P-bodies and small RNAs in translation regulation has also been elucidated. Initially, P-bodies were discovered as small foci rich in mRNA decay enzymes [43, 44, 45, 46, 47, 48]. When encountered with stress, yeast cells block translation initiation, as is evident by reduced polysome number and increased size of P-bodies [49]. Brengues et al. [50] have demonstrated that after the removal of stress and absence of new transcription, there is a decrease in the size of P-bodies and the reformation of polysomes [50]. This shows that P-bodies also act as storage sites for mRNA without undergoing degradation. On the other hand, small RNAs also the regulate translation and stability of mRNA [51, 52].

2.7 Posttranslational modifications and protein degradation

After its synthesis, the polypeptide is folded and modified by the addition of various chemical groups or by removal of certain amino acids from polypeptide (proteolytic cleavage). These protein modification steps act as targets for regulation. For example, phosphorylation of eIF2 results in inactivity and thus blocking of translation [53]. However, in some instances, phosphorylation may enhance the activity of a protein. Moreover, when not required by the cell, these proteins undergo degradation via the process of polyubiquitination, which is again regulated.

Advertisement

3. Regulation of transcription

In eukaryotes, there are three types of RNA polymerases responsible for transcription: RNA polymerase I is responsible for the synthesis of rRNA; RNA Pol II is responsible for the production of protein-coding mRNA, long non-coding RNA, snRNA, and miRNA; and RNA Pol III (RNA Pol III) is responsible for the synthesis of tRNA, some small non-coding RNA, 5S, and 5.8S RNA. Although transcription by all these enzymes is amenable to regulation, we will focus on the transcription of protein-coding genes in this chapter. Transcription is one of the most critical steps during gene expression and is regulated by a plethora of transcription factors working in a concerted manner with RNA Pol II to ensure proper initiation, elongation, and termination of transcription. RNA Pol II transcription involved initiation, elongation, and termination of transcription. For a long time, regulation of transcription was majorly focused on the initiation step. However, for over a decade, there has been an increase in mechanistic insights of regulation of transcription elongation, marking it as another regulatory event during transcription.

3.1 Transcription initiation

Initiation of transcription begins with recognizing and binding of RNA Pol II to the promoter. However, since RNA Pol II cannot recognize the promoter, it needs assistance from other proteins called as “general transcription factors” (GTFs) [54]. The GTFs are evolutionarily conserved proteins, and their ordered recruitment to RNA Pol II is necessary to initiate RNA synthesis. There are six types of GTFs: TFIIA, TFIIB, TFIID, TFIIE, TFIIF, and TFIIH. Table 1 summarizes the function of these GTFs.

General Transcription FactorSubunitsFunction
TFIIA3 (α, β, γ)Interact with TBP subunit of TFIID and stabilize PIC
TFIIB1Help in transcription start site selection, recruitment of TFIIH/RNA polymerase IIcomplex and assist in promoter escape
TFIID15 [TBP (TATA box binding protein) and 14 TAFs (TBP associating factors)]Recognize promoter by binding to TATA box, mediate interaction between activators and basal transcription machinery
TFIIE2 (TFA1, TFA2 in Saccharomyces cerevisiae or α, β in human)Facilitate TFIIH recruitment
TFIIF3 in S. Cerevisiae (Tfg1, Tfg2, Tfg3); 2 in human (RAP74, RAP20)Promote binding of RNA polymerase IIto TFIID-TFIIB DNA ternary complex, help in transcription start site selection
TFIIH10 [7 core subunits, 3 kinase modules (CAK)]DNA dependent ATPase, ATP dependent DNA helicase and CTD kinase, involved in promoter escape, promoter proximal pausing, elongation, RNA processing and termination

Table 1.

General transcription factors in eukaryotes and their functions.

Recruitment of RNA Pol II and general transcription factors (known as PIC or “Pre-Initiation Complex”) to the promoter is highly regulated during the initiation of transcription. Figure 2 shows the process of transcription initiation in eukaryotes. Specific transcription factors bind to the regulatory regions of the promoter. They work by modulating the assembly and activity of transcription machinery, either through direct interaction with components of basal transcription machinery or through action on chromatin [55, 56]. The mediator complex facilitates the connection between the activators associated with UAS and the PIC bound to the promoter region. The role played by the mediator complex, activators, and repressors marks yet another event of regulation during the initiation of transcription. The mediator complex is a multiprotein complex primarily comprised of head, middle, tail, and kinase modules. The head and middle domain form the core of the mediator complex, while the tail and kinase domains serve as regulatory modules [57, 58, 59, 60]. Although the mediator complex is conserved across evolution, the number of subunits vary in different species, comprising 19, 25, or up to 30 subunits in S. pombe, Saccharomyces cerevisiae, or humans, respectively. Five subunits have been reported to be metazoan-specific: MED23, MED25, MED26, MED28, and MED30 [61]. Figure 3 represents the subunit composition of the mediator complex in yeast and human. The mediator complex functions as a bridge between basal transcription machinery and specific transcription factors, resulting in the assembly of the pre-initiation complex (PIC) at the core promoter [62]. Besides recruitment of basal transcription machinery during initiation, the role of the mediator complex has been established in transition from initiation to elongation [63], during elongation [64, 65], as well as during termination [66], mRNA export [67], DNA repair [68], and S. pombe cell separation [69].

Figure 2.

Transcription initiation. Diagram representing stepwise recruitment of factors leading to initiation of transcription. Transcription activators bind to the UAS, which recruits co-activators including chromatin remodeling complexes. This opens the chromatin and facilitates association of RNA Pol II along with GTFs at the gene promoter, forming PIC. Association between activators and PIC is mediated by mediator complex, which phosphorylates RNA Pol II at CTD and initiates the process of transcription.

Figure 3.

Subunit composition of mediator complex in yeast and human. Diagram representing different modules comprising various subunits of mediator complex in yeast and human.

3.2 Transcription elongation

Promoter clearance is the transit phase between transcription initiation and elongation. Several factors determine the ability of RNA Pol II to move out of the promoter and enter the elongation phase. Co-crystallization of RNA Pol II with TFIIB has demonstrated that TFIIB obstructs the exit channel for newly synthesized RNA, and therefore, removal of TFIIB is imperative to promoter escape [70]. Transcription elongation is divided into two phases: early elongation and productive elongation [71]. The phosphorylation status of CTD of RNA Pol II is also an important determinant of the stage of transcription: during PIC assembly and initiation of transcription, CTD remains unphosphorylated [72, 73], whereas phosphorylation at serine 5 marks promoter clearance [74]. Serine 5 phosphorylated RNA Pol II associates with promoter-proximal regions during transcription initiation and early elongation, while serine 2 phosphorylation is associated with distal promoter regions during transcription elongation [75]. The CDK8 subunit of kinase module of mediator complex and CDK7 subunit of TFIIH are responsible for phosphorylation of CTD during initiation and early elongation phase, signaling RNA Pol II to clear promoter [76]. Phosphorylation of serine 2 on RNA Pol II CTD by P-TEFb triggers the passage into productive elongation from the early elongation phase [77, 78].

During the early elongation phase, RNA Pol II encounters various hurdles including transcriptional pause, arrest, or termination. The phenomenon of “promoter-proximal pausing” is characterized by transient pausing of RNA Pol II after synthesis of 20–60 nucleotides long RNA before resuming transcription elongation [71, 79]. Promoter-proximal pausing is well established in metazoans but less frequently observed in yeast [80, 81].

Evidence from biochemical studies has shown that pausing/arrest occurs as a result of backtracking of RNA Pol II on DNA template, thereby displacing the 3′ end of nascent RNA from the active site in RNA Pol II. This process can be spontaneously reversed (“pausing”) or not (“arrest”) [82]. The release of RNA Pol II from pause/arrest has emerged as an important mechanism to ensure continued and effective transcription elongation.

3.2.1 Transcription elongation factors

Biochemical experiments have shown that purified RNA Pol II proceeds optimally at rates of only 100–300 nucleotides/min in vitro as compared with the rate of 1200–2000 nucleotides/min in vivo [83, 84, 85]. The in vitro slow rate of mRNA synthesis was reported to be due to frequent pausing or arrest of RNA Pol II along the DNA [86, 87], suggesting elongation to be an inherently discontinuous process. An array of proteins known as “transcription elongation factors” function to regulate the rate of elongation.

These transcription elongation factors have been classified into different classes:

  • Factors that assist RNA Pol II to traverse through transient pausing sites, e.g., P-TEFb [73], DRB sensitivity inducing factor [88, 89];

  • Factors that can assist RNA Pol II to transcribe through chromatin, e.g., FACT, Swi/Snf [90, 91].

  • Factors that can increase the overall rate of transcribing RNA Pol II, e.g., Elongin [92, 93], ELL [94, 95], ELL2 [96].

  • Factors that suppress the activity of RNA Pol II, e.g., NELF [97].

    Some transcription elongation factors increase transcription of all protein coding genes, while others expedite transcription of only a set/class of genes [86]. Moreover, there are several GTFs as well as elongation factors that regulate transcription during either early or productive elongation phase of transcription.

3.2.2 Transcription factors regulating early elongation phase and pause release

3.2.2.1 TFIIF

TFIIE, TFIIF, and TFIIH have been implicated in post-initiation functions regulating early elongation and promoter escape during transcription [98].

In mammals, it is composed of two subunits, RAP30 and RAP74 [99]. Documented evidence suggests that TFIIF functions in suppressing RNA Pol II-associated transient pausing during active elongation via direct interaction with the ternary complex [100, 101]. Several lines of research have established that distinct functions of TFIIF during initiation and elongation are carried out by its different functional domains [100, 101, 102]. For instance, the initiation activity of TFIIF is mediated by a DNA binding domain in the C-terminal region of RAP30 while the elongation activity is carried out by the RNA Pol II binding regions in the upstream sequence of the RAP30 C-terminus [98].

3.2.2.2 TFIIH

TFIIH is a conserved protein composed of 10 subunits, seven of which form the core, while three subunits comprise a catalytic module called CDK activating kinase (CAK) comprising CDK7, ATP-dependent helicase XBP, and XPD [103].

The regulated recruitment of TFIIH to the promoter is orchestrated by TFIIE [104, 105, 106]. TFIIE and TFIIH work together in suppressing premature arrest of the early RNA Pol II, thereby facilitating promoter escape [98, 107]. The CDK7 phosphorylates CTD and initiates elongation [108]. Furthermore, it has been shown that the CDK7 subunit of TFIIH recruits ELL at sites of DNA damage and helps in transcription restart after repair of damaged DNA [109].

3.2.2.3 DSIF

A nucleoside analog, 5,6-dichloro-1-β-D-ribofuranosylbenzimidazole (DRB), works by obstructing the transition from initiation to elongation by inhibiting the phosphorylation of CTD of RNA Pol II [110, 111, 112]. DRB sensitivity inducing factor (DSIF) was initially identified from HeLa cell nuclear extracts as a transcription factor that promotes pausing of RNA Pol II in response to DRB [88, 97]. The two subunits of DSIF, Spt4 and Spt5, regulate the activity of RNA Pol II by genetically and physically interacting with it [88, 89, 97, 113, 114]. Association of DSIF with RNA Pol II after promoter escape is controlled by CDK7 dependent phosphorylation of the Spt5 subunit [115, 116]. Reduction in levels of the nascent and processed snRNA transcripts upon knockdown of DSIF in HeLa cell lines has pointed toward the function of DSIF as a transcriptional elongation regulator [117].

3.2.2.4 NELF

Negative elongation factor (NELF) is composed of five subunits, namely NELF-A, NELF-B, NELF-C, NELF-D, and NELF-E [118]. Interestingly, NELF is essential in Drosophila melanogaster and mammalian cells but is absent in Saccharomyces cerevisiae, C. elegans, and Arabidopsis thaliana [119, 120]. NELF promotes pausing of RNA Pol II by binding to RNA Pol II-DSIF complex through its NELF-E subunit [97, 118]. Yamaguchi et al. [97] demonstrated that DSIF/NELF interacts with the hypophosphorylated CTD of RNA Pol II to suppress its elongation function [97]. DSIF/NELF-mediated pausing of RNA Pol II gives sufficient time for recruitment of capping enzyme and addition of 5’cap to the nascent RNA [119, 121].

3.2.2.5 P-TEFb

Positive transcription elongation factor (P-TEFb) was first identified from a partially purified transcription system as a kinase inhibited by DRB [97]. It is composed of two subunits: cyclin T and CDK9. P-TEFb alleviates the NELF/DSIF-induced RNA Pol II pausing by phosphorylation of Spt5 subunit of DSIF and Ser2 residue of CTD heptad resulting in dissociation of NELF and transition from paused state to productive elongation state [71, 97, 119, 122, 123, 124]. Moreover, the FACT complex cooperates with P-TEFb to mitigate NELF/DSIF-mediated inhibition of transcription elongation [125]. In the cell, P-TEFb is subjected to stringent regulation and exists in an active as well as the inactive state. In mammals, most of the inactive P-TEFb exists as a part of complex, which includes an inactive ribonucleoprotein complex called 7SK snRNP, which consists of 7SK snRNA, P-TEFb, HEXIM1/2 (hexamethylene bisacetamide inducible protein), LARP7 (La-related protein 7), and MePCE (methyl phosphate capping enzyme) [126, 127]. However, when rapid transcription induction is required, P-TEFb is released from the inactive complex and is recruited to the transcription site by a specific activator or chromatin remodeling protein, bromodomain-containing protein 4 (BRD4) [79, 128, 129]. P-TEFb is a component of yet another multiprotein complex called super elongation complex (SEC), which is involved in productive elongation by RNA Pol II [130]. P-TEFb has been implicated as a potential therapeutic target in multiple myeloma, autoimmune diseases, cardiac hypertrophy, and infectious diseases [131, 132, 133, 134, 135, 136, 137]. An emerging concept of indirect therapeutic targeting using synthetic transcription elongation factor has shown enhanced gene expression in diseased conditions where a particular gene is downregulated. A synthetic transcription elongation factor is composed of a programmable DNA-binding ligand attached to a small molecule that can bind to and recruit the transcription elongation machinery, thereby regulating the gene expression. Syn-TEF1 has been shown to engage P-TEFb via recruitment of BRD4 at GAA repeats and restored the expression of the FXN gene in Fredrich Ataxia cell line to the levels observed in healthy cells [138].

Figure 4 depicts the role of P-TEFb in regulating the transition from the early elongation phase to productive elongation.

Figure 4.

Role of P-TEFb in regulating transcription elongation. (a) P-TEFb is recruited to the paused RNA Pol II as a result of the negative effects of DSIF and NELF. (b) At pause sites, P-TEFb phosphorylates RNA Pol II CTD as well as NELF and DSIF. (c) Due to phosphorylation, NELF dissociates, DSIF functions as a positive factor in the absence of NELF, recruits other factors for mRNA elongation and processing, resulting in productive elongation.

3.2.3 Transcription factors regulating productive elongation phase

A separate class of transcription factors regulate the process of transcription during productive elongation, which is described below:

3.2.3.1 Elongin

Elongin increases the overall rate of transcription by suppressing the transient pausing of RNA Pol II through its interaction with RNA Pol II and stabilizing it in an active conformation for an extended period [139, 140]. Elongin is composed of three subunits: Elongin A, Elongin B, and Elongin C. Elongin A was found to be enriched at the transcriptionally active sites in association with an active form of RNA Pol II on polytene chromosomes of Drosophila melanogaster [141]. Elongin B and C were also found to be components of various Cullin2/Cullin5-based ubiquitin ligase complexes and interact with different proteins in the complex via the BC box motif. Elongin BC serves as an adaptor linking Elongin A to Cul2/Cul5 and RING finger protein Rbx1/2 containing modules [142]. Several studies have shown that the association of Elongin ABC with Cul5/Rbx2 contributes significantly to the degradation of stalled RNA Pol II at sites of DNA damage [143, 144].

3.2.3.2 CSB

A mutation in Cockayne syndrome A (CSA) or Cockayne syndrome B (CSB) genes results in an accelerated neurodegenerative disorder called cocaine syndrome [145, 146]. Clinical studies using cells from Cockayne syndrome patients exhibited a defect in transcription coupled repair (TCR) but not in global genome repair, indicating the role of CSA/CSB in TCR [147, 148]. CSB was shown to be transiently associated with DNA, RNA Pol II, and nascent RNA, and in vitro studies have shown that purified CSB stimulated the rate of elongation by RNA Pol II [139, 149]. Transcribing RNA Pol II stalls upon detecting lesions in DNA [150]. Blocked RNA Pol II is either retracted or dissociated by CSA and CSB proteins, thereby making the damage site on DNA accessible to repair proteins, accomplishing DNA damage repair and augmenting the resumption of transcription [151, 152, 153].

3.2.3.3 TFIIS

TFIIS, a zinc finger transcription factor, is known to stimulate the rate of RNA transcript synthesis. TFIIS is required for stimulating transcription elongation by reducing pause time, and it also increases the processivity of RNA Pol II on nucleosomes as well as stimulates translational elongation [124, 154, 155]. Interaction of TFIIS with ELL-Associated Factor 2 (EAF2) promotes transactivation by FESTA/EAF2 in murine embryonic stem cells [156].

3.2.3.4 ELL family

ELL (Eleven nineteen Lysine rich Leukemia) was first identified as a translocating partner to trithorax-like mixed-lineage leukemia (MLL) gene, located on 11q23 chromosomal locus observed in acute myeloid leukemia [157]. Functional characterization and mechanistic studies have shown that ELL plays a role during recruitment of PIC, promoter clearance, and release of RNA Pol II from pause sites, thereby stimulating the overall rate of transcription [94, 158]. The functions of ELL in connection with various disease conditions have also been reported. The expression of HIF-1α, as well as its downstream target genes, is elevated in the absence of ELL as observed in PC3 prostate cancer cell lines [159]. The Tax protein of Human T-cell Leukemia Virus Type 1 (HTLV-1) interacts with ELL and incorporates it into the p300 and P-TEFb containing complexes, which enhance the transcription of immediate early genes [160].

In humans, two other ELL homologs, namely ELL2 and ELL3, were identified based on sequence similarity with ELL [124]. ELL2 regulates pre-mRNA processing by assisting RNA Pol II to select weak promoter-proximal poly(A) sites in immunoglobulin heavy gene (IgH) [161, 162, 163]. ELL2 has been found to be upregulated in neuroendocrine prostate tumor, while its mRNA level was reduced in prostate adenocarcinoma and multiple myeloma plasma cells [164, 165]. The role of ELL3 has been implicated in breast cancer and B-cell lymphoma [166, 167].

3.2.3.5 EAF family

Yeast two-hybrid screens are carried out to identify proteins that associate with human ELL and identified two ELL interacting partners, namely EAF1 and EAF2 (ELL Associated Factor) [168, 169]. EAF2, also known as androgen-upregulated 19 (U19), was recognized as a novel testosterone-regulated protein that induces apoptosis in the prostrate [170]. EAF family of proteins act as cofactors to ELL, stimulating its transcription elongation activity in vitro [171]. The importance of EAF in diseases was first highlighted by a study that showed that the association of EAF with ELL is essential for the immortalization of hematopoietic progenitor cells and the development of acute myeloid leukemia [169, 172]. Since then, there has been an explosion in the studies describing the role of EAF1 and EAF2 in the development and progression of various tumors. A reduced expression of EAF2 in human prostate cancer specimen, lower survival rates upon complete loss of EAF2, and increased cell migration and proliferation in prostate tumor cell lines upon EAF2 knockdown underscore the role of EAF2/U19 as a tumor suppressor in the prostate [173, 174, 175, 176, 177]. A murine model for EAF2 knockout has further implicated the role of EAF2 in other diseased conditions such as B-cell lymphoma, hepatocellular carcinoma, prostate intraepithelial neoplasia (PIN), lung adenocarcinoma, and enlarged cardiac cells with an abnormal vascular system as well as abnormalities in spermatogenesis [178, 179]. EAF2 knockdown is also associated with heightened humoral immune response and excessive generation of autoantibody. The immune balance function of EAF2 is mediated by apoptosis of germinal center B cells [180]. A study has identified a frameshift mutation that resulted in the loss of EAF2 function in colorectal cancer as well as gastric cancer [181]. Recently, a study established that absence of EAF1 in mouse prostate triggers pre-neoplastic prostatic intraepithelial neoplasia lesions. However, a combined loss of both EAF1 and EAF2 significantly enhanced the proliferation and inflammation in the murine prostate and resulted in a more aggressive tumor when compared with the individual loss of either EAF1 or EAF2, indicating that coordination between the two homologs is required for maintaining homeostasis in the prostate [182].

3.2.4 Elongation complexes

Several studies have demonstrated the existence of different multiprotein complexes called “Super Elongation Complex” (SEC), which are recruited to RNA Pol II to enhance its catalytic activity and thereby the rate of transcription elongation. These complexes are composed of transcription elongation factors such as p-TEFB, ELL1/2/3, EAF1/2 along with other proteins such as AFF4, ENL, AFF1, AFF9 [79]. In different organisms, super elongation complexes vary in composition and display functional diversity. For instance, in mammals, different combinations of four members of AFF family proteins (AFF1-4) and three members of ELL family protein (ELL, ELL2, and ELL3) confer functional variation to the SEC and form SEC-like complexes, SEC-L2 and SEC-L3 as shown in Figure 4. In place of AFF1/4, AFF2 and AFF3 represent the AFF family component of SEC-L2 and SEC-L3, respectively. Interestingly, biochemical studies have shown the absence of the one or more ELL family members in SEC-L2 and SEC-L3. The AFF4 and ELL2 containing SEC have been implicated in transcription elongation checkpoint control (TECC) in mammals. A similar super elongation complex was detected in D. melanogaster, comprising ELL, EAF1, P-TEFb, Lilli (AFF family member), and EAR (ENL) proteins [183]. Additionally, there is another complex present in D. melanogaster, known as Little Elongation Complex (LEC), comprising ICE1, ICE2, ELL, and EAF [184]. The role of LEC is found both in the initiation and elongation of snRNA transcription. Yet another elongation complex identified in hematopoietic cells called as “AEP complex” contains P-TEFb, AFF4, Afq31, and ENL protein [185]. The role of SEC in releasing paused RNA Pol II is very well described by several studies. However, other studies have also identified the role of SEC in rapid transcription by non-paused RNA Pol II in Drosophila embryos and in mouse embryonic stem cells [186, 187]. Recent work by Gopalan et al. [188] has recognized a rudimentary SEC in S. pombe consisting of only three members, ELL, EAF, and a newly identified AFF4 homolog EBP1 (ELL binding protein) [188]. The SECs were first reported two decades ago by their association with the viral transactivator of transcription (Tat) protein of HIV-1 and MLL-fusion partners involved in leukemogenesis [183, 189, 190, 191, 192]. During HIV infection, the Tat protein interacts with the P-TEFb component of SEC and recruits it to the HIV-1 long terminal repeat (LTR) to stimulate the expression of HIV-1 in host cells. Different AFF family members dictate gene-class specific recruitment of SEC [193]. For instance, AFF1 containing SEC is important for interaction with Tat protein in HIV pathogenesis. On the other hand, AFF4 containing SEC is involved in Hsp70 gene expression upon heat shock. Since several MLL fusion partners are components of SEC, the fusion protein-mediated recruitment of SEC to hox genes points toward dysregulation of developmental genes in leukemia [183, 191, 192]. Furthermore, mutations in SEC components also result in dysregulation of the transcription elongation process leading to several diseases. A missense mutation in AFF4 results in CHOPS syndrome, a developmental disorder [194]. Given the importance of regulation of transcription elongation in gene expression in healthy and diseased states, several research groups have worked toward delineating the roles of transcription elongation complexes as therapeutic targets in diseases including cancers. Small-molecule inhibitors targeting SEC, such as KL-1 and KL-2, have been implicated in targeted therapy of myc-driven tumors [195].

3.3 Transcription termination

In mammals, the transcription termination of protein-coding genes is mainly dependent on termination complex or “CPSF-CF complex” comprising cleavage and polyadenylation specificity factor (CPSF or CPF in yeast), cleavage stimulating factor (CstF or CF1A in yeast), cleavage factor I (CFI), and cleavage factor II (CFII) [196, 197]. CPSF directly binds to the body of Pol II and recognizes polyadenylation signal (PAS) (AAUAAA). Association of CPSF with PAS and Pol II triggers transcription pausing. CstF, which associates with Ser2 phosphorylated residues on Pol II CTD, recognizes GU-rich processing signal downstream of PAS and facilitates cleavage of transcripts. This dislodges the CPSF, and RNA Pol II is released from pause. In eukaryotes, two models have been proposed to facilitate the termination of Pol II-mediated transcription after cleavage of nascent RNA [196, 197, 198]. The allosteric model postulates that the transcription terminates following destabilization of the elongation complex triggered by the loss of elongation factors/conformational change in Pol II after transcription of PAS. The second “Torpedo model” posits that the transcription termination occurs due to the dismantling of Pol II elongation complex following the degradation of nascent RNA by exonuclease. SETX (Sen1 in yeast) resolves the R-loop formed by short RNA left after cleavage and DNA. This allows the recruitment of 5′-3′ exoribonuclease XRN2 (Rat1 in yeast), which chews up the nascent RNA downstream of the cleavage site and releases Pol II from DNA [198, 199]. However, an emerging view in this field is that the combination of these two models likely explains the process of termination [200, 201]. The mechanism of transcription termination is depicted in Figure 5. A detailed mechanistic understanding of transcription termination is still missing, despite a surge in studies highlighting the role of transcription termination in controlling gene expression.

Figure 5.

Mechanism of transcription termination at protein-coding genes in metazoans. Termination of transcription is triggered by the recruitment cleavage and polyadenylation factor complex (CPSF-CF complex) on the transcript. When RNA Pol II transcribes polyadenylation signal (PAS), the CPSF subunit binds to the PAS sequence on RNA while remaining bound to Pol II body. This results in pause of RNA Pol II. CstF subunit recognizes the GU-rich sequence downstream of PAS and creates a conformational change in the CPSF-CF complex, dissociating CPSF from Pol II and cleaving the nascent transcript between PAS and GU-rich sequence. R-loops formed by leftover transcript are resolved by Senataxin, allowing subsequent recruitment of XRN2 exonuclease. The remaining transcript downstream of the cleavage site is chewed up releasing Pol II and elongation complex by torpedo mechanism, thereby terminating the process of transcription.

3.4 Recent advances in regulation of gene expression

Numerous studies have shown that transcription is carried out in membraneless phase-separated compartments when biomacromolecules such as proteins and transcription factors undergo self-assembly via liquid-liquid-like phase separation (LLPS). These phase-separated condensates act as a hub of transcription, where specific transcription factors are exchanged to ensure proper temporal and spatial gene expression patterns [202]. It has been reported that Pol II forms phase-separated condensate at active genes through its low-complexity CTD to concentrate transcription regulators and initiate the process of transcription [203, 204, 205]. SEC can also dynamically extract P-TEFb from inactive HEXIM1-P-TEFb complex through phase separation, resulting in activation of transcription elongation. This was confirmed by disruption of SEC phase-separated droplets, which reduces the transcription efficiency [206]. The roles of LLPS in disease states have also been explored recently. Oncogenic fusion of ENL, a component of SEC, and MLL has been shown to enhance the phase separation capabilities contributing to overactivation of leukemic genes [206]. Mutation in the YEATS domain of ENL has also been shown to promote self-assembly of ENL into LLPS, resulting in augmented recruitment of SEC to promote transcription of oncogenes in Wilms’ tumor [207].

Another emerging concept of transcription regulation is through non-coding RNA and enhancer RNA (eRNA). Gene expression is regulated at multiple levels by long non-coding RNA (lncRNA): modulating chromatin structure and function, regulating transcription of neighboring or distal genes, RNA stability and translation [208, 209, 210, 211, 212, 213, 214]. The role of lncRNA has also recently been described in the regulation of nuclear condensates. Owing to the functional significance of lncRNA in transcription regulation, their role in the progression of diseases such as cancers and neurological disorders has been extensively studied. The transcription repression of tumor suppressors such as INK4A/ARF/INK4B has been associated with lncRNA ANRIL. ANRIL works by recruiting PRC1 and PRC2 complexes to promoters of these genes. Any dysregulation in ANRIL function may lead to silencing of these tumor suppressor genes contributing to tumor progression [215, 216, 217, 218]. BACE1-AS, an antisense of gene encoding BACE1 protein, a precursor of amyloid plates in Alzheimer’s disease (AD), promotes stability of BACE1 mRNA leading to accumulation of amyloid plates in the brains of AD patients. BACE1-AS also serves as a biomarker for AD and could be a potential therapeutic target to treat AD [219, 220]. lncRNAs are also involved in the suppression of gene expression through altered recruitment of transcription factors or Pol II or through reduced chromatin accessibility. For instance, following nerve injury, the lncRNA Silc1 is necessary for activation of the SOX11 transcription program for nerve regeneration [221].

Advertisement

4. Conclusion

Regulation of gene expression is imperative to the normal physiological functioning of the cell. In recent years, we have observed remarkable progress in our understanding of the regulation of gene expression. Transcriptional regulation has emerged to be the most critical and well-studied stage during the expression of a gene. In line of its foremost position in the transcription process, the initiation step is most extensively researched and was considered a major rate-limiting step during transcription. However, now the focus has shifted from activation to elongation stage, which has been established as another key regulatory event during gene expression under normal physiological state. Regulation of transcription termination has only recently started to become the focus of several studies, and more mechanistic insights are required to fully understand regulatory events during this stage of transcription.

Accumulating evidence in the last few years has suggested the prevalence of “gene-class specific” transcription elongation factors, adding another layer to transcriptional regulation. These transcription elongation factors have been reported to be part of several multiprotein elongation complexes, which enhance the probability of cross talk between these factors and increase the regulatory potential of cells. Identification of phase-separated assemblies of transcription complexes has provided a biophysical basis of dynamic regulation of transcription in response to cellular cues. Another less-explored layer of transcription regulation is through non-coding RNA. Although there has been a tremendous increase in our understanding of the regulatory capabilities of lncRNA, we still lack a rigorous investigation to relate sequence and structural features of non-coding RNA to their regulatory functions.

A recent surge in targeted gene therapy has opened the doors for therapeutic targeting of “gene-class specific” transcription elongation factors in various diseases including cancers. An interesting concept in therapeutic targeting is of synthetic transcription elongation factors, which modulate the expression of a particular gene by selectively engaging the transcription elongation machinery at a specific gene locus. However, more efforts and research are required to dwell into the genome-wide perturbation of gene expression as a result of the binding of synthetic transcription factors to other less specific loci. In the context of personalized medicines, disease-related non-coding RNAs are gaining attention due to their specific expression patterns, which makes them a good candidate for disease biomarkers. Growing mechanistic insight into the regulation of transcription elongation and the interplay between different steps of regulation of gene expression would offer new aspects for intervention with aberrant modulation of gene expression and precisely tuned therapeutics.

Advertisement

Acknowledgments

The author would like to thank Mr. Anurag Saroha for his help with the figures.

References

  1. 1. Ezkurdia I, Juan D, Rodriguez JM, Frankish A, Diekhans M, Harrow J, et al. Multiple evidence strands suggest that there may be as few as 19 000 human protein-coding genes. Human Molecular Genetics. 2014;23(22):5866-5878
  2. 2. Cosgrove MS, Boeke JD, Wolberger C. Regulated nucleosome mobility and the histone code. Nature Structural & Molecular Biology. 2004;11(11):1037
  3. 3. Kouzarides T. Chromatin modifications and their function. Cell. 2007;128(4):693-705
  4. 4. Hendrich B, Bickmore W. Human diseases with underlying defects in chromatin structure and modification. Human Molecular Genetics. 2001;10(20):2233-2242
  5. 5. Huang C, Sloan EA, Boerkoel CF. Chromatin remodeling and human disease. Current Opinion in Genetics & Development. 2003;13(3):246-252
  6. 6. Zoghbi HY. Postnatal neurodevelopmental disorders: Meeting at the synapse? Science. 2003;302(5646):826-830
  7. 7. Bochar DA, Wang L, Beniya H, Kinev A, Xue Y, Lane WS, et al. BRCA1 is associated with a human SWI/SNF-related complex: Linking chromatin remodeling to breast cancer. Cell. 2000;102(2):257-265
  8. 8. Davis PK, Brachmann RK. Chromatin remodeling and cancer. Cancer Biology & Therapy. 2003;2(1):23-30
  9. 9. Jacquier A. The complex eukaryotic transcriptome: Unexpected pervasive transcription and novel small RNAs. Nature Reviews Genetics. 2009;10(12):833-844
  10. 10. Bassett CL. Control of gene expression by mRNA transport and turnover. In: Regulation of Gene Expression in Plants. Boston, MA: Springer; 2007. pp. 148-188
  11. 11. Lewis JD, Izaurflde E. The role of the cap structure in RNA processing and nuclear export. European Journal of Biochemistry. 1997;247(2):461-469
  12. 12. Gu M, Lima CD. Processing the message: Structural insights into capping and decapping mRNA. Current Opinion in Structural Biology. 2005;15(1):99-106
  13. 13. Graff JR, Konicek BW, Vincent TM, Lynch RL, Monteith D, Weir SN, et al. Therapeutic suppression of translation initiation factor eIF4E expression reduces tumor growth without toxicity. The Journal of Clinical Investigation. 2007;117(9):2638-2648
  14. 14. Black DL. Mechanisms of alternative pre-messenger RNA splicing. Annual Review of Biochemistry. 2003;72(1):291-336
  15. 15. Chen M, Manley JL. Mechanisms of alternative splicing regulation: Insights from molecular and genomics approaches. Nature Reviews Molecular Cell Biology. 2009;10(11):741
  16. 16. Shepard PJ, Choi EA, Lu J, Flanagan LA, Hertel KJ, Shi Y. Complex and dynamic landscape of RNA polyadenylation revealed by PAS-Seq. RNA. 2011;17(4):761-772
  17. 17. Derti A, Garrett-Engele P, MacIsaac KD, Stevens RC, Sriram S, Chen R, et al. A quantitative atlas of polyadenylation in five mammals. Genome Research. 2012;22(6):1173-1183
  18. 18. Lin Y, Li Z, Ozsolak F, Kim SW, Arango-Argoty G, Liu TT, et al. An in-depth map of polyadenylation sites in cancer. Nucleic Acids Research. 2012;40(17):8460-8471
  19. 19. Tian B, Manley JL. Alternative polyadenylation of mRNA precursors. Nature Reviews Molecular Cell Biology. 2017;18(1):18
  20. 20. Danckwardt S, Hentze MW, Kulozik AE. 3′ end mRNA processing: Molecular mechanisms and implications for health and disease. The EMBO Journal. 2008;27(3):482-498
  21. 21. Wickramasinghe VO, Laskey RA. Control of mammalian gene expression by selective mRNA export. Nature Reviews Molecular Cell Biology. 2015;16(7):431-442
  22. 22. Lei EP, Krebber H, Silver PA. Messenger RNAs are recruited for nuclear export during transcription. Genes & Development. 2001;15(14):1771-1782
  23. 23. Maniatis T, Reed R. An extensive network of coupling among gene expression machines. Nature. 2002;416(6880):499
  24. 24. Neugebauer KM. On the importance of being co-transcriptional. Journal of Cell Science. 2002;115(20):3865-3871
  25. 25. Hilleren P, McCarthy T, Rosbash M, Parker R, Jensen TH. Quality control of mRNA 3′-end processing is linked to the nuclear exosome. Nature. 2001;413(6855):538
  26. 26. Hammell CM, Gross S, Zenklusen D, Heath CV, Stutz F, Moore C, et al. Coupling of termination, 3′ processing, and mRNA export. Molecular and Cellular Biology. 2002;22(18):6441-6457
  27. 27. Kessler MM, Henry MF, Shen E, Zhao J, Gross S, Silver PA, et al. Hrp1, a sequence specific RNA-binding protein that shuttles between the nucleus and the cytoplasm, is required for mRNA 3′-end formation in yeast. Genes & Development. 1997;11(19):2545-2556
  28. 28. Hector RE, Nykamp KR, Dheur S, Anderson JT, Non PJ, Urbinati CR, et al. Dual requirement for yeast hnRNP Nab2p in mRNA poly (A) tail length control and nuclear export. The EMBO Journal. 2002;21(7):1800-1810
  29. 29. Hollams EM, Giles KM, Thomson AM, Leedman PJ. MRNA stability and the control of gene expression: Implications for human disease. Neurochemical Research. 2002;27(10):957-980
  30. 30. Balagopal V, Fluch L, Nissan T. Ways and means of eukaryotic mRNA decay. Biochimica et Biophysica Acta (BBA)-Gene Regulatory Mechanisms. 2012;1819(6):593-603
  31. 31. Adjibade P, Mazroui R. Control of mRNA turnover: Implication of cytoplasmic RNA granules. Seminars in Cell & Developmental Biology. 2014;34:15-23
  32. 32. Yang E, van Nimwegen E, Zavolan M, Rajewsky N, Schroeder M, Magnasco M, et al. Decay rates of human mRNAs: Correlation with functional characteristics and sequence attributes. Genome Research. 2003;13(8):1863-1872
  33. 33. Hollis GF, Gazdar AF, Bertness V, Kirsch IR. Complex translocation disrupts c-myc regulation in a human plasma cell myeloma. Molecular and Cellular Biology. 1988;8(1):124-129
  34. 34. Aghib DF, Bishop JM, Ottolenghi S, Guerrasio A, Serra A, Saglio G. A 3’truncation of MYC caused by chromosomal translocation in a human T-cell leukemia increases mRNA stability. Oncogene. 1990;5(5):707-711
  35. 35. Schuman EM, Dynes JL, Steward O. Synaptic regulation of translation of dendritic mRNAs. Journal of Neuroscience. 2006;26(27):7143-7146
  36. 36. Thompson B, Wickens M, Kimble J. 19 translational control in development. Cold Spring Harbor Monograph Archive. 2007;48:507-544
  37. 37. Morley SJ, Coldwell MJ. 16 matters of life and death: Translation initiation during apoptosis. Cold Spring Harbor Monograph Archive. 2007;48:433-458
  38. 38. Holcik M, Sonenberg N. Translational control in stress and apoptosis. Nature Reviews Molecular Cell Biology. 2005;6(4):318
  39. 39. Schneider RJ, Sonenberg N. 15 translational control in cancer development and progression. Cold Spring Harbor Monograph Archive. 2007;48:401-431
  40. 40. Gebauer F, Hentze MW. Molecular mechanisms of translational control. Nature Reviews Molecular Cell Biology. 2004;5(10):827
  41. 41. Mathews MB, Sonenberg N, Hershey JW. 1 origins and principles of translational control. Cold Spring Harbor Monograph Archive. 2007;48:1-40
  42. 42. Preiss T, W. Hentze M. Starting the protein synthesis machine: Eukaryotic translation initiation. BioEssays. 2003;25(12):1201-1211
  43. 43. Bashkirov VI, Scherthan H, Solinger JA, Buerstedde JM, Heyer WD. A mouse cytoplasmic exoribonuclease (mXRN1p) with preference for G4 tetraplex substrates. The Journal of Cell Biology. 1997;136(4):761-773
  44. 44. Cougot N, Babajko S, Seraphin B. Cytoplasmic foci are sites of mRNA decay in human cells. The Journal of Cell Biology. 2004;165:31-40
  45. 45. Ingelfinger D, Arndt-Jovin DJ, Lührmann R, Achsel T. The human LSm1-7 proteins colocalize with the mRNA-degrading enzymes Dcp1/2 and Xrn1 in distinct cytoplasmic foci. RNA. 2002;8(12):1489-1501
  46. 46. Lykke-Andersen J. Identification of a human decapping complex associated with hUpf proteins in nonsense-mediated decay. Molecular and Cellular Biology. 2002;22(23):8114-8121
  47. 47. Sheth U, Parker R. Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science. 2003;300(5620):805-808
  48. 48. Van Dijk E, Cougot N, Meyer S, Babajko S, Wahle E, Séraphin B. Human Dcp2: A catalytically active mRNA decapping enzyme located in specific cytoplasmic structures. The EMBO Journal. 2002;21(24):6915-6924
  49. 49. Coller J, Parker R. General translational repression by activators of mRNA decapping. Cell. 2005;122(6):875-886
  50. 50. Brengues M, Teixeira D, Parker R. Movement of eukaryotic mRNAs between polysomes and cytoplasmic processing bodies. Science. 2005;310(5747):486-489
  51. 51. Bartel DP. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281-297
  52. 52. Filipowicz W. RNAi: The nuts and bolts of the RISC machine. Cell. 2005;122(1):17-20
  53. 53. Sonenberg N, Hinnebusch AG. Regulation of translation initiation in eukaryotes: Mechanisms and biological targets. Cell. 2009;136(4):731-745
  54. 54. Orphanides G, Lagrange T, Reinberg D. The general transcription factors of RNA Pol II. Genes & Development. 1996;10(21):2657-2683
  55. 55. Sikorski TW, Buratowski S. The basal initiation machinery: Beyond the general transcription factors. Current Opinion in Cell Biology. 2009;21(3):344-351
  56. 56. van Bakel H. Interactions of transcription factors with chromatin. In: A Handbook of Transcription Factors. Dordrecht: Springer; 2011. pp. 223-259
  57. 57. Kornberg RD. Mediator and the mechanism of transcriptional activation. Trends in Biochemical Sciences. 2005;30(5):235-239
  58. 58. Malik S, Roeder RG. The metazoan mediator co-activator complex as an integrative hub for transcriptional regulation. Nature Reviews Genetics. 2010;11(11):761
  59. 59. Cevher MA, Shi Y, Li D, Chait BT, Malik S, Roeder RG. Reconstitution of active human core mediator complex reveals a critical role of the MED14 subunit. Nature Structural & Molecular Biology. 2014;21(12):1028
  60. 60. Plaschka C, Lariviere L, Wenzeck L, Seizl M, Hemann M, Tegunov D, et al. Architecture of the RNA Pol II—Mediator core initiation complex. Nature. 2015;518(7539):376
  61. 61. Harper TM, Taatjes DJ. The complex structure and function of mediator. Journal of Biological Chemistry. 2018;293(36):13778-13785
  62. 62. Soutourina J. Transcription regulation by the mediator complex. Nature Reviews Molecular Cell Biology. 2018;19(4):262
  63. 63. Kim YJ, Björklund S, Li Y, Sayre MH, Kornberg RD. A multiprotein mediator of transcriptional activation and its interaction with the C-terminal repeat domain of RNA Pol II. Cell. 1994;77(4):599-608
  64. 64. Gaillard H, Tous C, Botet J, González-Aguilera C, Quintero MJ, Viladevall L, et al. Genome-wide analysis of factors affecting transcription elongation and DNA repair: A new role for PAF and Ccr4-not in transcription-coupled repair. PLoS Genetics. 2009;5(2):e1000364
  65. 65. Kremer SB, Kim S, Jeon JO, Moustafa YW, Chen A, Zhao J, et al. Role of mediator in regulating Pol II elongation and nucleosome displacement in Saccharomyces cerevisiae. Genetics. 2012;191(1):95-106
  66. 66. Mukundan B, Ansari A. Novel role for mediator complex subunit Srb5/Med18 in termination of transcription. Journal of Biological Chemistry. 2011;286(43):37053-37057
  67. 67. Schneider M, Hellerschmied D, Schubert T, Amlacher S, Vinayachandran V, Reja R, et al. The nuclear pore-associated TREX-2 complex employs mediator to regulate gene expression. Cell. 2015;162(5):1016-1028
  68. 68. Eyboulet F, Cibot C, Eychenne T, Neil H, Alibert O, Werner M, et al. Mediator links transcription and DNA repair by facilitating Rad2/XPG recruitment. Genes & Development. 2013;27(23):2549-2562
  69. 69. Mehta S, Miklos I, Sipiczki M, Sengupta S, Sharma N. The Med8 mediator subunit interacts with the Rpb4 subunit of RNA Pol II and Ace2 transcriptional activator in Schizosaccharomyces pombe. FEBS Letters. 2009;583(19):3115-3120
  70. 70. Bushnell DA, Westover KD, Davis RE, Kornberg RD. Structural basis of transcription: An RNA Pol II-TFIIB cocrystal at 4.5 angstroms. Science. 2004;303(5660):983-988
  71. 71. Kwak H, Lis JT. Control of transcriptional elongation. Annual Review of Genetics. 2013;47:483-508
  72. 72. Laybourn PJ, Dahmus ME. Transcription-dependent structural changes in the C-terminal domain of mammalian RNA polymerase subunit IIa/o. Journal of Biological Chemistry. 1989;264(12):6693-6698
  73. 73. Lu H, Flores O, Weinmann R, Reinberg D. The nonphosphorylated form of RNA Pol II preferentially associates with the preinitiation complex. Proceedings of the National Academy of Sciences. 1991;88(22):10004-10008
  74. 74. Christmann JL, Dahmus ME. Monoclonal antibody specific for calf thymus RNA polymerases IIO and IIA. Journal of Biological Chemistry. 1981;256(22):11798-11803
  75. 75. Jones JC, Phatnani HP, Haystead TA, MacDonald JA, Alam SM, Greenleaf AL. C-terminal repeat domain kinase I phosphorylates Ser2 and Ser5 of RNA Pol II C-terminal domain repeats. Journal of Biological Chemistry. 2004;279(24):24957-24964
  76. 76. Allen BL, Taatjes DJ. The mediator complex: A central integrator of transcription. Nature Reviews Molecular Cell Biology. 2015;16(3):155
  77. 77. Marshall NF, Price DH. Purification of P-TEFb, a transcription factor required for the transition into productive elongation. Journal of Biological Chemistry. 1995;270(21):12335-12338
  78. 78. Marshall NF, Peng J, Xie Z, Price DH. Control of RNA Pol II elongation potential by a novel carboxyl-terminal domain kinase. Journal of Biological Chemistry. 1996;271(43):27176-27183
  79. 79. Zhou Q, Li T, Price DH. RNA Pol II elongation control. Annual Review of Biochemistry. 2012;81:119-143
  80. 80. Mayer A, Lidschreiber M, Siebert M, Leike K, Söding J, Cramer P. Uniform transitions of the general RNA Pol II transcription complex. Nature Structural & Molecular Biology. 2010;17(10):1272-1278
  81. 81. Booth GT, Wang IX, Cheung VG, Lis JT. Divergence of a conserved elongation factor and transcription regulation in budding and fission yeast. Genome Research. 2016;26(6):799-811
  82. 82. Conaway JW, Shilatifard A, Dvir A, Conaway RC. Control of elongation by RNA Pol II. Trends in Biochemical Sciences. 2000;25(8):375-380
  83. 83. Ucker DS, Yamamoto KR. Early events in the stimulation of mammary tumor virus RNA synthesis by glucocorticoids. Novel assays of transcription rates. Journal of Biological Chemistry. 1984;259(12):7416-7420
  84. 84. Thummel CS, Burtis KC, Hogness DS. Spatial and temporal patterns of E74 transcription during Drosophila development. Cell. 1990;61(1):101-111
  85. 85. Tennyson CN, Klamut HJ, Worton RG. The human dystrophin gene requires 16 hours to be transcribed and is cotranscriptionally spliced. Nature Genetics. 1995;9(2):184
  86. 86. Reines D, Conaway JW, Conaway RC. The RNA Pol II general elongation factors. Trends in Biochemical Sciences. 1996;21(9):351-355
  87. 87. Uptain SM, Kane CM, Chamberlin MJ. Basic mechanisms of transcript elongation and its regulation. Annual Review of Biochemistry. 1997;66(1):117-172
  88. 88. Wada T, Takagi T, Yamaguchi Y, Ferdous A, Imai T, Hirose S, et al. DSIF, a novel transcription elongation factor that regulates RNA Pol II processivity, is composed of human Spt4 and Spt5 homologs. Genes & Development. 1998;12(3):343-356
  89. 89. Hartzog GA, Wada T, Handa H, Winston F. Evidence that Spt4, Spt5, and Spt6 control transcription elongation by RNA Pol II inSaccharomyces cerevisiae. Genes & Development. 1998;12(3):357-369
  90. 90. Orphanides G, Reinberg D. RNA Pol II elongation through chromatin. Nature. 2000;407(6803):471
  91. 91. Arndt KM, Kane CM. Running with RNA polymerase: Eukaryotic transcript elongation. Trends in Genetics. 2003;19(10):543-550
  92. 92. Bradsher JN, Jackson KW, Conaway RC, Conaway JW. RNA Pol II transcription factor SIII. I. Identification, purification, and properties. Journal of Biological Chemistry. 1993;268(34):25587-25593
  93. 93. Garrett KP, Aso T, Bradsher JN, Foundling SI, Lane WS, Conaway RC, et al. Positive regulation of general transcription factor SIII by a tailed ubiquitin homolog. Proceedings of the National Academy of Sciences. 1995;92(16):7172-7176
  94. 94. Shilatifard A, Lane WS, Jackson KW, Conaway RC, Conaway JW. An RNA Pol II elongation factor encoded by the human ELL gene. Science. 1996;271(5257):1873-1876
  95. 95. Shilatifard A, Haque D, Conaway RC, Conaway JW. Structure and function of RNA Pol II elongation factor ELL identification of two overlapping ell functional domains that govern its interaction with polymerase and the ternary elongation complex. Journal of Biological Chemistry. 1997;272(35):22355-22363
  96. 96. Shilatifard A, Duan DR, Haque D, Florence C, Schubach WH, Conaway JW, et al. ELL2, a new member of an ELL family of RNA Pol II elongation factors. Proceedings of the National Academy of Sciences. 1997;94(8):3639-3643
  97. 97. Yamaguchi Y, Wada T, Watanabe D, Takagi T, Hasegawa J, Handa H. Structure and function of the human transcription elongation factor DSIF. Journal of Biological Chemistry. 1999;274(12):8085-8092
  98. 98. Yan Q, Moreland RJ, Conaway JW, Conaway RC. Dual roles for transcription factor IIF in promoter escape by RNA Pol II. Journal of Biological Chemistry. 1999;274(50):35668-35675
  99. 99. Conaway RC, Conaway JW. General initiation factors for RNA Pol II. Annual Review of Biochemistry. 1993;62(1):161-190
  100. 100. Tan S, Aso T, Conaway RC, Conaway JW. Roles for both the RAP30 and RAP74 subunits of transcription factor IIF in transcription initiation and elongation by RNA Pol II. Journal of Biological Chemistry. 1994;269(41):25684-25691
  101. 101. Kephart DD, Wang BQ, Burton ZF, Price DH. Functional analysis of Drosophila factor 5 (TFIIF), a general transcription factor. Journal of Biological Chemistry. 1994;269(18):13536-13543
  102. 102. Wang BQ, Burton ZF. Functional domains of human RAP74 including a masked polymerase binding domain. Journal of Biological Chemistry. 1995;270(45):27035-27044
  103. 103. Lee TI, Young RA. Transcription of eukaryotic protein-coding genes. Annual Review of Genetics. 2000;34(1):77-137
  104. 104. Goodrich JA, Tjian R. Transcription factors IIE and IIH and ATP hydrolysis direct promoter clearance by RNA Pol II. Cell. 1994;77(1):145-156
  105. 105. Ohkuma Y, Roeder RG. Regulation of TFIIH ATPase and kinase activities by TFIIE during active initiation complex formation. Nature. 1994;368(6467):160
  106. 106. Holstege FC, Van der Vliet PC, Timmers HT. Opening of an RNA Pol II promoter occurs in two distinct steps and requires the basal transcription factors IIE and IIH. The EMBO Journal. 1996;15(7):1666-1677
  107. 107. Schilbach S, Hantsche M, Tegunov D, Dienemann C, Wigge C, Urlaub H, et al. Structures of transcription pre-initiation complex with TFIIH and mediator. Nature. 2017;551(7679):204
  108. 108. Glover-Cutter K, Larochelle S, Erickson B, Zhang C, Shokat K, Fisher RP, et al. TFIIHassociated Cdk7 kinase functions in phosphorylation of C-terminal domain Ser7 residues, promoter-proximal pausing, and termination by RNA Pol II. Molecular and Cellular Biology. 2009;29(20):5455-5464
  109. 109. Mourgues S, Gautier V, Lagarou A, Bordier C, Mourcet A, Slingerland J, et al. ELL, a novel TFIIH partner, is involved in transcription restart after DNA repair. Proceedings of the National Academy of Sciences. 2013;110(44):17927-17932
  110. 110. Dubois MF, Nguyen VT, Bellier S, Bensaude O. Inhibitors of transcription such as 5, 6-dichloro- 1-beta-D-ribofuranosylbenzimidazole and isoquinoline sulfonamide derivatives (H-8 and H-7) promote dephosphorylation of the carboxyl-terminal domain of RNA Pol II largest subunit. Journal of Biological Chemistry. 1994;269(18):13331-13336
  111. 111. Mancebo HS, Lee G, Flygare J, Tomassini J, Luu P, Zhu Y, et al. P-TEFb kinase is required for HIV tat transcriptional activation in vivo and in vitro. Genes & Development. 1997;11(20):2633-2644
  112. 112. Zhu Y, Pe’ery T, Peng J, Ramanathan Y, Marshall N, Marshall T, et al. Transcription elongation factor P-TEFb is required for HIV-1 tat transactivation in vitro. Genes & Ddevelopment. 1997;11(20):2622-2632
  113. 113. Swanson MS, Malone EA, Winston FR. SPT5, an essential gene important for normal transcription in Saccharomyces cerevisiae, encodes an acidic nuclear protein with a carboxy-terminal repeat. Molecular and Cellular Biology. 1991;11(6):3009-3019
  114. 114. Hartzog GA, Basrai MA, Ricupero-Hovasse SL, Hieter P, Winston F. Identification and analysis of a functional human homolog of the SPT4 gene of Saccharomyces cerevisiae. Molecular and Cellular Biology. 1996;16(6):2848-2856
  115. 115. Larochelle S, Batliner J, Gamble MJ, Barboza NM, Kraybill BC, Blethrow JD, et al. Dichotomous but stringent substrate selection by the dual-function Cdk7 complex revealed by chemical genetics. Nature Structural & Molecular Biology. 2006;13(1):55
  116. 116. Patel SA, Simon MC. Functional analysis of the CDK7. Cyclin H. Mat1 complex in mouse embryonic stem cells and embryos. Journal of Biological Chemistry. 2010;285(20):15587-15598
  117. 117. Yamamoto J, Hagiwara Y, Chiba K, Isobe T, Narita T, Handa H, et al. DSIF and NELF interact with integrator to specify the correct post-transcriptional fate of snRNA genes. Nature Communications. 2014;5:4263
  118. 118. Yamaguchi Y, Filipovska J, Yano K, Furuya A, Inukai N, Narita T, et al. Stimulation of RNA Pol II elongation by hepatitis delta antigen. Science. 2001;293(5527):124-127
  119. 119. Sims RJ, Belotserkovskaya R, Reinberg D. Elongation by RNA Pol II: The short and long of it. Genes & Development. 2004;18(20):2437-2468
  120. 120. Yamaguchi Y, Shibata H, Handa H. Transcription elongation factors DSIF and NELF: Promoter proximal pausing and beyond. Biochimica Et Biophysica Acta (BBA)-Gene Regulatory Mechanisms. 2013;1829(1):98-104
  121. 121. Pei Y, Shuman S. Interactions between fission yeast mRNA capping enzymes and elongation factor Spt5. Journal of Biological Chemistry. 2002;277(22):19639-19648
  122. 122. Ivanov D, Kwak YT, Guo J, Gaynor RB. Domains in the SPT5 protein that modulate its transcriptional regulatory properties. Molecular and Cellular Biology. 2000;20(9):2970-2983
  123. 123. Kim JB, Sharp PA. Positive transcription elongation factor B phosphorylates hSPT5 and RNA Pol II carboxyl-terminal domain independently of cyclin-dependent kinaseactivating kinase. Journal of Biological Chemistry. 2001;276(15):12317-12323
  124. 124. Shilatifard A, Conaway RC, Conaway JW. The RNA Pol II elongation complex. Annual Review of Biochemistry. 2003;72(1):693-715
  125. 125. Wada T, Orphanides G, Hasegawa J, Kim DK, Shima D, Yamaguchi Y, et al. FACT relieves DSIF/NELF-mediated inhibition of transcriptional elongation and reveals functional differences between P-TEFb and TFIIH. Molecular Cell. 2000;5(6):1067-1072
  126. 126. Rice AP. Dysregulation of positive transcription elongation factor B and myocardial hypertrophy. Circulation Research. 2009;104(12):1327-1329
  127. 127. Peterlin BM, Price DH. Controlling the elongation phase of transcription with P-TEFb. Molecular Cell. 2006;23(3):297-305
  128. 128. Jang MK, Mochizuki K, Zhou M, Jeong HS, Brady JN, Ozato K. The bromodomain protein Brd4 is a positive regulatory component of P-TEFb and stimulates RNA Pol II dependent transcription. Molecular Cell. 2005;19(4):523-534
  129. 129. Yang Z, Yik JH, Chen R, He N, Jang MK, Ozato K, et al. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Molecular Cell. 2005;19(4):535-545
  130. 130. Smith E, Lin C, Shilatifard A. The super elongation complex (SEC) and MLL in development and disease. Genes & Development. 2011;25(7):661-672
  131. 131. Yamamoto M, Onogi H, Kii I, Yoshida S, Iida K, Sakai H, et al. CDK9 inhibitor FIT-039 prevents replication of multiple DNA viruses. The Journal of Clinical Investigation. 2014;124(8):3479-3488
  132. 132. Hu Z, Yik JH, Cissell DD, Michelier PV, Athanasiou KA, Haudenschild DR. Inhibition of CDK9 prevents mechanical injury-induced inflammation, apoptosis and matrix degradation in cartilage explants. European Cells & Materials. 2016;30:200
  133. 133. Tanaka T, Okuyama-Dobashi K, Murakami S, Chen W, Okamoto T, Ueda K, et al. Inhibitory effect of CDK9 inhibitor FIT-039 on hepatitis B virus propagation. Antiviral Research. 2016;133:156-164
  134. 134. Zaborowska J, Isa NF, Murphy S. P-TEFb goes viral. BioEssays. 2016;38:S75-S85
  135. 135. Zhang J, Li G, Ye X. Cyclin T1/CDK9 interacts with influenza A virus polymerase and facilitates its association with cellular RNA Pol II. Journal of Virology. 2010;84(24):12619-12627
  136. 136. Schrecengost RS, Green CL, Zhuang Y, Keller SN, Smith RA, Maines LW, et al. In vitro and in vivo antitumor and anti-inflammatory capabilities of the novel GSK3 and CDK9 inhibitor ABC1183. Journal of Pharmacology and Experimental Therapeutics. 2018;365(1):107-116
  137. 137. Qing Y, Wang X, Wang H, Hu P, Li H, Yu X, et al. Pharmacologic targeting of the P-TEFb complex as a therapeutic strategy for chronic myeloid leukemia. Cell Communication and Signaling. 2021;19(1):1-6
  138. 138. Erwin GS, Grieshop MP, Ali A, Qi J, Lawlor M, Kumar D, et al. Synthetic transcription elongation factors license transcription across repressive chromatin. Science. 2017;358(6370):1617-1622
  139. 139. Conaway JW, Conaway RC. Transcription elongation and human disease. Annual Review of Biochemistry. 1999;68(1):301-319
  140. 140. Sharma N. Regulation of RNA Pol II-mediated transcriptional elongation: Implications in human disease. IUBMB Life. 2016;68(9):709-716
  141. 141. Gerber M, Eissenberg JC, Kong S, Tenney K, Conaway JW, Conaway RC, et al. In vivo requirement of the RNA Pol II elongation factor elongin A for proper gene expression and development. Molecular and Cellular Biology. 2004;24(22):9911-9919
  142. 142. Schoenfeld AR, Davidowitz EJ, Burk RD. Elongin BC complex prevents degradation of von Hippel-Lindau tumor suppressor gene products. Proceedings of the National Academy of Sciences. 2000;97(15):8507-8512
  143. 143. Yasukawa T, Kamura T, Kitajima S, Conaway RC, Conaway JW, Aso T. Mammalian Elongin A complex mediates DNA-damage-induced ubiquitylation and degradation of Rpb1. The EMBO Journal. 2008;27(24):3256-3266
  144. 144. Harreman M, Taschner M, Sigurdsson S, Anindya R, Reid J, Somesh B, et al. Distinct ubiquitin ligases act sequentially for RNA Pol II polyubiquitylation. Proceedings of the National Academy of Sciences. 2009;106(49):20705-20710
  145. 145. Nance MA, Berry SA. Cockayne syndrome: Review of 140 cases. American Journal of Medical Genetics. 1992;42(1):68-84
  146. 146. Wilson BT, Stark Z, Sutton RE, Danda S, Ekbote AV, Elsayed SM, et al. The Cockayne syndrome natural history (CoSyNH) study: Clinical findings in 102 individuals and recommendations for care. Genetics in Medicine. 2016;18(5):483
  147. 147. Venema JL, Mullenders LH, Natarajan AT, Van Zeeland AV, Mayne LV. The genetic defect in Cockayne syndrome is associated with a defect in repair of UV-induced DNA damage in transcriptionally active DNA. Proceedings of the National Academy of Sciences of the United States of America. 1990;87(12):4707
  148. 148. van Hoffen A, Natarajan AT, Mayne LV, Zeeland AA, Mullenders LH, Venema J. Deficient repair of the transcribed strand of active genes in Cockayne’s syndrome cells. Nucleic Acids Research. 1993;21(25):5890-5895
  149. 149. Vélez-Cruz R, Egly JM. Cockayne syndrome group B (CSB) protein: At the crossroads of transcriptional networks. Mechanisms of Ageing and Development. 2013;134(5-6):234-242
  150. 150. Mellon I, Spivak G, Hanawalt PC. Selective removal of transcription-blocking DNA damage from the transcribed strand of the mammalian DHFR gene. Cell. 1987;51(2):241-249
  151. 151. Troelstra C, van Gool A, de Wit J, Vermeulen W, Bootsma D, Hoeijmakers JH. ERCC6, a member of a subfamily of putative helicases, is involved in Cockayne’s syndrome and preferential repair of active genes. Cell. 1992;71(6):939-953
  152. 152. Hanawalt PC. Transcription-coupled repair and human disease. Science. 1994;266(5193):1957
  153. 153. Hanawalt PC, Spivak G. Transcription-coupled DNA repair: Two decades of progress and surprises. Nature Reviews Molecular Cell Biology. 2008;9(12):958-970
  154. 154. Mason PB, Struhl K. Distinction and relationship between elongation rate and processivity of RNA Pol II in vivo. Molecular Cell. 2005;17(6):831-840
  155. 155. Ishibashi T, Dangkulwanich M, Coello Y, Lionberger TA, Lubkowska L, Ponticelli AS, et al. Transcription factors IIS and IIF enhance transcription efficiency by differentially modifying RNA polymerase pausing dynamics. Proceedings of the National Academy of Sciences. 2014;111(9):3419-3424
  156. 156. Ito T, Saso K, Arimitsu N, Sekimizu K. Defective FESTA/EAF2-mediated transcriptional activation in S-II-deficient embryonic stem cells. Biochemical and Biophysical Research Communications. 2007;363(3):603-609
  157. 157. Thirman MJ, Levitan DA, Kobayashi H, Simon MC, Rowley JD. Cloning of ELL, a gene that fuses to MLL in at (11; 19)(q23; p13. 1) in acute myeloid leukemia. Proceedings of the National Academy of Sciences. 1994;91(25):12110-12114
  158. 158. Byun JS, Fufa TD, Wakano C, Fernandez A, Haggerty CM, Sung MH, et al. ELL facilitates RNA Pol II pause site entry and release. Nature Communications. 2012;3:633
  159. 159. Liu L, Ai J, Xiao W, Liu J, Wang Y, Xin D, et al. ELL is an HIF-1 partner that regulates and responds to hypoxia response in PC3 cells. The Prostate. 2010;70(7):797-805
  160. 160. Fufa TD, Byun JS, Wakano C, Fernandez AG, Pise-Masison CA, Gardner K. The tax oncogene enhances ELL incorporation into p300 and P-TEFb containing protein complexes to activate transcription. Biochemical and Biophysical Research Communications. 2015;465(1):5-11
  161. 161. Martincic K, Alkan SA, Cheatle A, Borghesi L, Milcarek C. Transcription elongation factor ELL2 directs immunoglobulin secretion in plasma cells by stimulating altered RNA processing. Nature Immunology. 2009;10(10):1102
  162. 162. Swaminathan B, Thorleifsson G, Jöud M, Ali M, Johnsson E, Ajore R, et al. Variants in ELL2 influencing immunoglobulin levels associate with multiple myeloma. Nature Communications. 2015;6:7213
  163. 163. Shell SA, Martincic K, Tran J, Milcarek C. Increased phosphorylation of the carboxyl-terminal domain of RNA Pol II and loading of polyadenylation and cotranscriptional factors contribute to regulation of the ig heavy chain mRNA in plasma cells. The Journal of Immunology. 2007;179(11):7663-7673
  164. 164. Pascal LE, Masoodi KZ, Liu J, Qiu X, Song Q, Wang Y, et al. Conditional deletion of ELL2 induces murine prostate intraepithelial neoplasia. Journal of Endocrinology. 2017;235(2):123-136
  165. 165. Ali M, Ajore R, Wihlborg AK, Niroula A, Swaminathan B, Johnsson E, et al. The multiple myeloma risk allele at 5q15 lowers ELL2 expression and increases ribosomal gene expression. Nature Communications. 2018;9(1):1649
  166. 166. Ahn HJ, Kim G, Park KS. Ell3 stimulates proliferation, drug resistance, and cancer stem cell properties of breast cancer cells via a MEK/ERK-dependent signaling pathway. Biochemical and Biophysical Research Communications. 2013;437(4):557-564
  167. 167. Alexander LE, Watters J, Reusch JA, Maurin M, Nepon-Sixt BS, Vrzalikova K, et al. Selective expression of the transcription elongation factor ELL3 in B cells prior to ELL2 drives proliferation and survival. Molecular Immunology. 2017;91:8-16
  168. 168. Simone F, Polak PE, Kaberlein JJ, Luo RT, Levitan DA, Thirman MJ. EAF1, a novel ELLassociated factor that is delocalized by expression of the MLL-ELL fusion protein. Blood. 2001;98(1):201-209
  169. 169. Simone F, Luo RT, Polak PE, Kaberlein JJ, Thirman MJ. ELL-associated factor 2 (EAF2), a functional homolog of EAF1 with alternative ELL binding properties. Blood. 2003;101(6):2355-2362
  170. 170. Hahn J, Xiao W, Jiang F, Simone F, Thirman MJ, Wang Z. Apoptosis induction and growth suppression by U19/Eaf2 is mediated through its ELL-binding domain. The Prostate. 2007;67(2):146-153
  171. 171. Kong SE, Banks CA, Shilatifard A, Conaway JW, Conaway RC. ELL-associated factors 1 and 2 are positive regulators of RNA Pol II elongation factor ELL. Proceedings of the National Academy of Sciences. 2005;102(29):10094-10098
  172. 172. Luo RT, Lavau C, Du C, Simone F, Polak PE, Kawamata S, et al. The elongation domain of ELL is dispensable but its ELL-associated factor 1 interaction domain is essential for MLL-ELL-induced leukemogenesis. Molecular and Cellular Biology. 2001;21(16):5678-5687
  173. 173. Qiao Z, Wang D, Hahn J, Ai J, Wang Z. Pirin down-regulates the EAF 2/U 19 protein and alleviates its growth inhibition in prostate cancer cells. The Prostate. 2014;74(2):113-120
  174. 174. Zhang Q, Zeng L, Zhao C, Ju Y, Konuma T, Zhou MM. Structural insights into histone crotonyllysine recognition by the AF9 YEATS domain. Structure. 2016;24(9):1606-1612
  175. 175. Wang Y, Pascal LE, Zhong M, Ai J, Wang D, Jing Y, et al. Combined loss of EAF2 and p53 induces prostate carcinogenesis in male mice. Endocrinology. 2017;158(12):4189-4205
  176. 176. Yang T, Jing Y, Dong J, Yu X, Zhong M, Pascal LE, et al. Regulation of ELL2 stability and polyubiquitination by EAF2 in prostate cancer cells. The Prostate. 2018;78(15):1201-1212
  177. 177. Zhong M, Pascal LE, Cheng E, Masoodi KZ, Chen W, Green A, et al. Concurrent EAF2 and ELL2 loss phenocopies individual EAF2 or ELL2 loss in prostate cancer cells and murine prostate. American Journal of Clinical and Experimental Urology. 2018;6(6):234
  178. 178. Xiao W, Zhang Q, Habermacher G, Yang X, Zhang AY, Cai X, et al. U19/Eaf2 knockout causes lung adenocarcinoma, B-cell lymphoma, hepatocellular carcinoma and prostatic intraepithelial neoplasia. Oncogene. 2008;27(11):1536
  179. 179. Xiao W, Ai J, Habermacher G, Volpert O, Yang X, Zhang AY, et al. U19/Eaf2 binds to and stabilizes von hippel-Lindau protein. Cancer Research. 2009;69(6):2599-2606
  180. 180. Li Y, Sabari BR, Panchenko T, Wen H, Zhao D, Guan H, et al. Molecular coupling of histone crotonylation and active transcription by AF9 YEATS domain. Molecular Cell. 2016;62(2):181-193
  181. 181. Jo YS, Kim SS, Kim MS, Yoo NJ, Lee SH. Candidate tumor suppressor gene EAF2 is mutated in colorectal and gastric cancers. Pathology Oncology Research. 2019;25(2):823-824
  182. 182. Pascal LE, Su F, Wang D, Ai J, Song Q, Wang Y, et al. Conditional deletion of eaf1 induces murine prostatic intraepithelial neoplasia in mice. Neoplasia. 2019;21(8):752-764
  183. 183. Lin C, Smith ER, Takahashi H, Lai KC, Martin-Brown S, Florens L, et al. AFF4, a component of the ELL/P-TEFb elongation complex and a shared subunit of MLL chimeras, can link transcription elongation to leukemia. Molecular Cell. 2010;37(3):429-437
  184. 184. Hu D, Smith ER, Garruss AS, Mohaghegh N, Varberg JM, Lin C, et al. The little elongation complex functions at initiation and elongation phases of snRNA gene transcription. Molecular Cell. 2013;51(4):493-505
  185. 185. Lu H, Li Z, Xue Y, Schulze-Gahmen U, Johnson JR, Krogan NJ, et al. AFF1 is a ubiquitous P-TEFb partner to enable tat extraction of P-TEFb from 7SK snRNP and formation of SECs for HIV transactivation. Proceedings of the National Academy of Sciences. 2014;111(1):E15-E24
  186. 186. Dahlberg O, Shilkova O, Tang M, Holmqvist PH, Mannervik M. P-TEFb, the super elongation complex and mediator regulate a subset of non-paused genes during early Drosophila embryo development. PLoS Genetics. 2015;11(2):e1004971
  187. 187. Lin C, Garrett AS, De Kumar B, Smith ER, Gogol M, Seidel C, et al. Dynamic transcriptional events in embryonic stem cells mediated by the super elongation complex (SEC). Genes & Development. 2011;25(14):1486-1498
  188. 188. Gopalan S, Gibbon DM, Banks CA, Zhang Y, Florens LA, Washburn MP, et al. Schizosaccharomyces pombe Pol II transcription elongation factor ELL functions as part of a rudimentary super elongation complex. Nucleic Acids Research. 2018;46(19):10095-10105
  189. 189. He N, Liu M, Hsu J, Xue Y, Chou S, Burlingame A, et al. HIV-1 tat and host AFF4 recruit two transcription elongation factors into a bifunctional complex for coordinated activation of HIV-1 transcription. Molecular Cell. 2010;38(3):428-438
  190. 190. Sobhian B, Laguette N, Yatim A, Nakamura M, Levy Y, Kiernan R, et al. HIV-1 tat assembles a multifunctional transcription elongation complex and stably associates with the 7SK snRNP. Molecular Cell. 2010;38(3):439-451
  191. 191. Yokoyama A, Lin M, Naresh A, Kitabayashi I, Cleary ML. A higher-order complex containing AF4 and ENL family proteins with P-TEFb facilitates oncogenic and physiologic MLL-dependent transcription. Cancer Cell. 2010;17(2):198-212
  192. 192. Luo Z, Lin C, Guest E, Garrett AS, Mohaghegh N, Swanson S, et al. The super elongation complex family of RNA pol II elongation factors: Gene target specificity and transcriptional output. Molecular and Cellular Biology. 2012;32(13):2608-2617
  193. 193. Lu H, Li Z, Zhang W, Schulze-Gahmen U, Xue Y, Zhou Q. Gene target specificity of the super elongation complex (SEC) family: How HIV-1 tat employs selected SEC members to activate viral transcription. Nucleic Acids Research. 2015;43(12):5868-5879
  194. 194. Izumi K, Nakato R, Zhang Z, Edmondson AC, Noon S, Dulik MC, et al. Germline gain-of-function mutations in AFF4 cause a developmental syndrome functionally linking the super elongation complex and cohesin. Nature Genetics. 2015;47(4):338-344
  195. 195. Liang K, Smith ER, Aoi Y, Stoltz KL, Katagi H, Woodfin AR, et al. Targeting processive transcription elongation via SEC disruption for MYC-induced cancer therapy. Cell. 2018;175(3):766-779
  196. 196. Kuehner JN, Pearson EL, Moore C. Unravelling the means to an end: RNA Pol II transcription termination. Nature Reviews Molecular Cell Biology. 2011;12(5):283-294
  197. 197. Jensen TH, Jacquier A, Libri D. Dealing with pervasive transcription. Molecular Cell. 2013;52(4):473-484
  198. 198. Porrua O, Libri D. Transcription termination and the control of the transcriptome: Why, where and how to stop. Nature Reviews Molecular Cell Biology. 2015;16(3):190-202
  199. 199. Rosonina E, Kaneko S, Manley JL. Terminating the transcript: Breaking up is hard to do. Genes & Development. 2006;20(9):1050-1056
  200. 200. Mischo HE, Gómez-González B, Grzechnik P, Rondón AG, Wei W, Steinmetz L, et al. Yeast Sen1 helicase protects the genome from transcription-associated instability. Molecular Cell. 2011;41(1):21-32
  201. 201. Richard P, Manley JL. Transcription termination by nuclear RNA polymerases. Genes & Development. 2009;23(11):1247-1269
  202. 202. Shin Y, Brangwynne CP. Liquid phase condensation in cell physiology and disease. Science. 2017;357(6357)
  203. 203. Kwon I, Kato M, Xiang S, Wu L, Theodoropoulos P, Mirzaei H, et al. Phosphorylation-regulated binding of RNA Pol II to fibrous polymers of low-complexity domains. Cell. 2013;155(5):1049-1060
  204. 204. Boehning M, Dugast-Darzacq C, Rankovic M, Hansen AS, Yu T, Marie-Nelly H, et al. RNA Pol II clustering through carboxy-terminal domain phase separation. Nature Structural & Molecular Biology. 2018;25(9):833-840
  205. 205. Lu H, Yu D, Hansen AS, Ganguly S, Liu R, Heckert A, et al. Phase-separation mechanism for C-terminal hyperphosphorylation of RNA Pol II. Nature. 2018;558(7709):318-323
  206. 206. Guo C, Che Z, Yue J, Xie P, Hao S, Xie W, et al. ENL initiates multivalent phase separation of the super elongation complex (SEC) in controlling rapid transcriptional activation. Science Advances. 2020;6(14):eaay4858
  207. 207. Wan L, Chong S, Xuan F, Liang A, Cui X, Gates L, et al. Impaired cell fate through gain-of-function mutations in a chromatin reader. Nature. 2020;577(7788):121-126
  208. 208. Isoda T, Moore AJ, He Z, Chandra V, Aida M, Denholtz M, et al. Non-coding transcription instructs chromatin folding and compartmentalization to dictate enhancer-promoter communication and T cell fate. Cell. 2017;171(1):103-119
  209. 209. Mumbach MR, Granja JM, Flynn RA, Roake CM, Satpathy AT, Rubin AJ, et al. HiChIRP reveals RNA-associated chromosome conformation. Nature Methods. 2019;16(6):489-492
  210. 210. Luo S, Lu JY, Liu L, Yin Y, Chen C, Han X, et al. Divergent lncRNAs regulate gene expression and lineage differentiation in pluripotent cells. Cell Stem Cell. 2016;18(5):637-652
  211. 211. Kim YJ, Xie P, Cao L, Zhang MQ, Kim TH. Global transcriptional activity dynamics reveal functional enhancer RNAs. Genome Research. 2018;28(12):1799-1811
  212. 212. Lee S, Kopp F, Chang TC, Sataluri A, Chen B, Sivakumar S, et al. Noncoding RNA NORAD regulates genomic stability by sequestering PUMILIO proteins. Cell. 2016;164(1-2):69-80
  213. 213. Tichon A, Perry RB, Stojic L, Ulitsky I. SAM68 is required for regulation of Pumilio by the NORAD long noncoding RNA. Genes & Development. 2018;32(1):70-78
  214. 214. Carrieri C, Cimatti L, Biagioli M, Beugnet A, Zucchelli S, Fedele S, et al. Long non-coding antisense RNA controls Uchl1 translation through an embedded SINEB2 repeat. Nature. 2012;491(7424):454-457
  215. 215. Aguilo F, Zhou MM, Walsh MJ. Long noncoding RNA, polycomb, and the ghosts haunting INK4b-ARF-INK4a expression. Cancer Research. 2011;71(16):5365-5369
  216. 216. Popov N, Gil J. Epigenetic regulation of the INK4b-ARF-INK4a locus: In sickness and in health. Epigenetics. 2010;5(8):685-690
  217. 217. Kotake Y, Nakagawa T, Kitagawa K, Suzuki S, Liu N, Kitagawa M, et al. Long non-coding RNA ANRIL is required for the PRC2 recruitment to and silencing of p15 INK4B tumor suppressor gene. Oncogene. 2011;30(16):1956-1962
  218. 218. Yap KL, Li S, Muñoz-Cabello AM, Raguz S, Zeng L, Mujtaba S, et al. Molecular interplay of the noncoding RNA ANRIL and methylated histone H3 lysine 27 by polycomb CBX7 in transcriptional silencing of INK4a. Molecular Cell. 2010;38(5):662-674
  219. 219. Faghihi MA, Zhang M, Huang J, Modarresi F, Van der Brug MP, Nalls MA, et al. Evidence for natural antisense transcript-mediated inhibition of microRNA function. Genome Biology. 2010;11(5):1-3
  220. 220. Feng L, Liao YT, He JC, Xie CL, Chen SY, Fan HH, et al. Plasma long non-coding RNA BACE1 as a novel biomarker for diagnosis of Alzheimer disease. BMC Neurology. 2018;18(1):1-8
  221. 221. Perry RB, Hezroni H, Goldrich MJ, Ulitsky I. Regulation of neuroregeneration by long noncoding RNAs. Molecular Cell. 2018;72(3):553-567

Written By

Preeti Dabas

Submitted: 26 December 2021 Reviewed: 02 February 2022 Published: 08 March 2022