ISSN 0006-2979, Biochemistry (Moscow), 2024, Vol. 89, No. 4, pp. 688-700 © Pleiades Publishing, Ltd., 2024.
688
REVIEW
Nonspecific Interactions in Transcription Regulation
and Organization of Transcriptional Condensates
Anna A. Valyaeva
1,2,3,a
* and Eugene V. Sheval
2,3
1
Faculty of Bioengineering and Bioinformatics, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, 119991 Moscow, Russia
3
Department of Cell Biology and Histology, Faculty of Biology, Lomonosov Moscow State University,
119991Moscow, Russia
a
e-mail: valyaeva.ann@gmail.com
Received October 10, 2023
Revised November 19, 2023
Accepted November 20, 2023
AbstractEukaryotic cells are characterized by a high degree of compartmentalization of their internal contents,
which ensures precise and controlled regulation of intracellular processes. During many processes, including
different stages of transcription, dynamic membraneless compartments termed biomolecular condensates are
formed. Transcription condensates contain various transcription factors and RNA polymerase and are formed by
high- and low-specificity interactions between the proteins, DNA, and nearby RNA. This review discusses recent
data demonstrating important role of nonspecific multivalent protein–protein and RNA–protein interactions in
organization and regulation of transcription.
DOI: 10.1134/S0006297924040084
Keywords: transcription, RNA, RNA polymerase, transcriptional condensates, intrinsically disordered regions
Abbreviations: ARM, arginine-rich motif; CTD,C-terminal do-
main; DFC,dense fibrillar component; GC,granular compo-
nent; IDR, intrinsically disordered region; mRNA, messen-
ger RNA; PRC2, Polycomb repressive complex 2; RNAPI,RNA
polymerase I; RNAPII, RNA polymerase II; RNAPIII, RNA
polymerase III; rRNA,ribosomal RNA; tRNA, transfer RNA.
* To whom correspondence should be addressed.
INTRODUCTION
Gene expression in the nucleus is orchestrated
by the coordinated activity of three RNA polymerases:
RNA polymerase I (RNAPI), which is responsible for
transcribing the precursor of 18S, 5.8S, and 25S ribo-
somal RNA(rRNA); RNA polymerase II(RNAPII), which
is predominantly involved in transcribing messenger
RNAs(mRNAs); and RNA polymerase III(RNAPIII), which
is responsible for transcribing ribosomal 5S rRNA, trans-
fer RNA (tRNA), 7SL RNA, and various other small RNAs.
Regulation of transcription is a complex process that
relies on the intricate interplay of numerous proteins
and RNAs that accumulate at specific loci in the nu-
cleus. An illustrative example of such accumulation
is nucleolus, the largest organelle within the nucleus,
where the processes of transcription, pre-mRNA pro-
cessing, and pre-ribosomal assembly take place[1, 2].
The key feature of the cell nucleus, which en-
ables transcription and its flexible regulation, is high
mobility of nuclear proteins. In 2000, data began to
accumulate indicating that many proteins in the nu-
cleolus, interchromatin granules, and chromatin are
not stably bound to the structures, but rather are in
continuous and relatively rapid exchange with the sur-
rounding nucleoplasm [3, 4]. Focus of this review is on
the nature of binding and exchange of the transcrip-
tion factors responsible for carrying out the relatively
long process of transcription. While constant binding
of transcription factors to promoters could theoreti-
cally activate transcription, early experiments showed
that the glucocorticoid receptor, a hormone-dependent
modulator of gene expression, undergoes rapid ex-
change between the chromatin target site and the sur-
rounding nucleoplasm [3]. It was later shown that both
the glucocorticoid receptor and its interacting partner,
glucocorticoid receptor interacting protein 1 (GRIP-1),
exist in a dynamic equilibrium with the promoter and
must repeatedly return to the DNA template during
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 689
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
transcription activation [5]. Other transcription factors
also exhibit dynamic interactions with promoters, oc-
cupying the promoter region for relatively short peri-
ods of time ranging from fractions of a second to tens
of seconds [6]. RNAPII subunits have also been shown
to dynamically assemble at the promoter site [7].
High exchange rates have also been demonstrat-
ed for some rRNA transcription factors – UBF [8-10],
TAFI48, PAF53, and TIF-IA/Rrn3 [9]. RNAPI subunits
also exhibited dynamic behavior at the promoter, sug-
gesting stochasticity and low efficiency in the assem-
bly of complexes required for transcription [9].
Behavior of transcription factors is described by
the 3D genome-scanning model [6], also referred to as
the diffusion and affinity model. This model assumes
that proteins diffuse freely and rapidly throughout
the nucleus, intermittently interacting with its compo-
nents. Most of these interactions are nonspecific and
non-functional. However, when proteins engage with
the target sites through high-affinity interactions, they
could become temporarily immobilized. This, with a
seemingly low probability, may result in formation of
the functional macromolecular complexes.
However, this model fails to describe all quantita-
tive regularities associated with regulation of the ge-
nome activity. For example, it has been calculated that
the rate at which proteins locate their sites on DNA in
bacterial cells is approximately 100 times higher than
predicted based on diffusion alone [11, 12]. Several
mechanisms have now been proposed that explain
enhanced efficiency and accuracy of protein accumu-
lation. At least some of these mechanisms rely pre-
dominantly on low-specificity interactions between
molecules, leading, in particular, to formation of the
membraneless structures now commonly referred to
as biomolecular condensates. These condensates serve
as repositories for the molecules essential for efficient
transcription. Given that formation of condensates in
many cases depends on the interaction between pro-
teins and RNAs, this review will systematically explore
the roles of proteins and RNAs involved in organiza-
tion and regulation of transcription through formation
of transcriptional condensates.
COMPARTMENTALIZATION
OF TRANSCRIPTION
Regulation of gene transcription is primarily
based on numerous high-affinity (specific) interactions
of various proteins with enhancers and promoters. For
example, transcription activation depends on the in-
teraction of the Mediator complex with the enhancer
and promoter, which facilitates landing of RNAPII
during transcription initiation [13, 14]. However, it has
been suggested in a number of studies that transcrip-
tion initiation can lead to formation of rather large
complexes – transcription factories – in which many
transcribed genes are concentrated. This hypothesis is
compelling as it holds the promise of coordinated reg-
ulation for numerous genes within a single transcrip-
tion factory. Existence of such complexes is primarily
supported by the microscopic data showcasing clus-
tering of RNAPII molecules in animal nuclei [15-20].
These observations were initially made using conven-
tional fluorescence microscopy rather than super-res-
olution localization microscopy, which allows precise
determination of the number of molecules in a given
area. The results obtained using the super-resolution
microscopy indicate that transcription factories are
relatively small and are predominantly composed of
only a few RNAPII molecules [21].
Subunits of the Mediator complex are also found
among the constituents of the transcription factories
[22]. It has been suggested that a fraction as small as
<10% of all clusters consisting of large assemblies of
RNAPII and Mediator may correspond to the previous-
ly described clusters of super-enhancers. Regulation
of certain crucial genes involves multiple enhancers,
known as super-enhancers [23-27]. Existing data indi-
cate that the large clusters of RNAPII and the Mediator
complex often are co-localized with the loci containing
super-enhancers [22, 28-30].
Live cell observations suggest that such transcrip-
tion factories are formed as dynamic, short-lived aggre-
gates that include RNAPII molecules [31-33]. A similar
situation has been demonstrated for the transcription
factors Zelda and Bicoid in the developing Drosophila
embryos [34]. Both of these factors generate dynamic,
short-lived condensates that transiently interact with
active sites of Bicoid-dependent transcription thus ac-
tivating it.
Finally, microscopic visualization of the transcrip-
tion of long genes should be considered. Long genes
form rather extended loops along which RNAPII mol-
ecules appear to move [35, 36]. These loops share sim-
ilarities with the structures observed in lampbrush
chromosomes [37-39]. Consequently, recent data do
not align with the concept of highly ordered transcrip-
tion factories; instead, they suggest the existence of
highly dynamic complexes. These dynamic complexes
are now commonly referred to as transcriptional con-
densates [40].
ROLE OF NONSPECIFIC INTERACTIONS
IN THE FORMATION
OF TRANSCRIPTIONAL CONDENSATES
The concept of biomolecular condensates originat-
ed in the study of membraneless organelles [41, 42].
Subsequently, this concept was extended to any com-
VALYAEVA, SHEVAL690
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
Fig. 1. Role of intrinsically disordered domains (IDRs) in formation of biomolecular condensates. a)Electrostatic interactions
between the protein IDRs in the presence of domains with different charges. b)Interactions between the protein IDRs and RNA.
c)Formation of a biomolecular condensate due to multivalent interactions provided by the dimers of proteins with IDRs binding
RNA molecules.
plex that does not exhibit stoichiometry in its composi-
tion [43]. Formation of such complexes, irrespective of
their size, relies on numerous low-affinity interactions
among the molecules. Crucially, interactions involving
intrinsically disordered protein domains (IDRs) play
a pivotal (though not exclusive) role [44-47] (Fig. 1,
a and b). Additionally, multivalent interactions involv-
ing structured domains are of significance [48, 49].
Interms of their role in the condensate formation, pro-
tein molecules are categorized into scaffold molecules,
participating in multivalent interactions and contrib-
uting to condensate formation, and client proteins,
which accumulate within already formed condensates
[49]. Although IDRs do not exhibit high affinity, they,
as well as the high-affinity RNA-binding domains, can
serve as foundation for multivalent interactions
either due to extensive IDR length or through oligom-
erization of the protein molecules (Fig.1c).
Ability of some proteins involved in transcription
to form condensates in vitro suggests that transcription
may be accompanied by the formation of biomolecular
condensates. The most detailed data are available for
some nuclear proteins, in particular for two pivotal
nucleolar proteins, FBL and NPM1 [50]. Numerous other
proteins engaged in the diverse aspects of genome ac-
tivity regulation demonstrate capability to form con-
densates in vitro. In some instances, the analysis was
limited to the IDR of these proteins, but even such data
appear to be sufficient to indicate potential involve-
ment of the proteins in the biomolecular condensate
formation. This group of proteins include components
of the Mediator complex and RNAPII [22, 28, 29], TAZ
[51], BRD4 [28], and transcription factors OCT4 [52]
and TAF15 [53].
One of the most striking pieces of evidence sup-
porting the formation of the IDR-dependent biomolec-
ular condensates during transcription activation comes
from the study of the Wnt/β-catenin signaling pathway
[54]. Upon pathway activation, β-catenin translocates
to the nucleus, where it forms a complex with the LEF/
TCF transcription factors. This complex recruits addi-
tional cofactors and enhances expression of the target
genes. LEF1 demonstrates the ability to form biomo-
lecular condensates with β-catenin both in vitro and
in vivo, and the formation of these condensates is es-
sential for transcription activation. Notably, LEF1 with
disrupted IDR loses its activity, which can be restored
by substituting with the intrinsically disordered do-
main of another protein (FUS). This emphasized the
significance of nonspecific interactions in the tran-
scription regulation.
Moreover, recent data suggest that IDRs can not
only enhance transcription efficiency by attracting
diverse proteins into the transcriptional condensates,
but also improve accuracy of the promoter recogni-
tion by transcription factors. Transcription factors
possess DNA-binding domains that bind to the specific
short DNA sequence motifs. However, these DNA-bind-
ing domains constitute only a fraction of transcrip-
tion factor sequences, with the majority being of low
complexity and lacking a stable 3D structure [55-57].
As mentioned earlier, IDRs are not capable of high-af-
finity and high-specificity interactions. Nevertheless,
these regions can “target” transcription factors to the
specific genomic loci, thereby increasing specificity of
the promoter recognition, when acting in concert with
high-affinity DNA-binding domains [58-60].
ROLE OF RNAPII C-TERMINAL DOMAIN
IN THE ORGANIZATION
OF TRANSCRIPTIONAL CONDENSATES
C-terminal domain of RNAPII (CTD) plays a dis-
tinctive role in organization of transcription processes.
The CTD consists of numerous heptad repeats with the
consensus sequence Tyr 1-Ser 2-Pro 3-Thr 4-Ser 5-Pro 6-
Ser 7 [61-63], and is characterized as an intrinsically
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 691
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
disordered region. Its length exhibits considerable
variation across different organisms, ranging from 26
repeats in Saccharomyces cerevisiae to 52 repeats in
humans.
When expressed in cells, the CTD accumulates at
active transcription sites, indicating its potential role
in the recruitment of RNAPII at target sites [64]. How-
ever, the data regarding its direct involvement in the
transcription process are conflicting. Catalytic activi-
ty of RNAPII relies on the conserved enzyme subunits
RPB1 and RPB2, rather than the CTD [65], so it seems
logical that the RNAPII with deleted CTD can perform
transcription in vitro but not in vivo [66]. Multiple in-
vestigations have explored this issue, with the most re-
cent study utilizing Raji cells suggesting that the CTD
is not required for transcription in live cells [67].
The findings presented in this study suggest that the
CTD could be required not for the transcription itself,
but rather for the precise localization of RNAPII at the
target sites, as well as for regulation of the transcrip-
tion-related processes and transcript maturation.
Both human and yeast CTD sequences are able to
form condensates in vitro, resulting in spherical aggre-
gates that contain intact RNAPII [68]. Importantly, the
ability of RNAPII to cluster in the nucleus correlates
with the number of heptad repeats: shortening the
human CTD to the length of the yeast CTD reduces
concentration of the accumulated RNAPII and its as-
sociation with chromatin in human cells. Conversely,
increasing the CTD length enhances clustering of the
RNAPII molecules[69]. Modeling confirms that the CTD
length promotes binding of RNAPII molecules to the
promoter and delays their release from it, while the
CTD–CTD interactions facilitate concentration of nu-
merous RNAPII molecules at the single point. Several
studies also suggest that CTD is involved in the recruit-
ment of the Mediator complex [22,28,70]. Disruption
of the Mediator complex results in disassembly of the
large complexes of hypophosphorylated RNAPII [71].
The studies cited above suggest that formation of
biomolecular condensates accumulating RNAPII, vari-
ous transcription factors, and other proteins could rep-
resent a crucial mechanism for enhancing transcrip-
tion processes. However, this mechanism might be less
effective in S. cerevisiae. The CTD in this organism is
much shorter than, for example, in humans, where
the longer CTD leads to the stronger CTD–CTD interac-
tions and formation of the less dynamic condensates
[68]. This distinction could potentially explain why the
RNAPII molecules in S. cerevisiae nuclei do not form
clusters, a characteristic feature of the human cell nu-
clei [72]. Observations of individual RNAPII molecules
in the living yeast cells demonstrated that CTD plays a
role in restricting diffusion of RNAPII within the nu-
clear region enriched with active genes, although no
visible condensate formation was evident. Shortening
of the CTD resulted in the reduced diffusion restric-
tion, increased target site search time, and impaired
formation of the preinitiation complex. According to
the authors, these differences could be attributed to
the unique organization of transcription within a small
yeast cell nucleus, where all components are concen-
trated in a limited space, making formation of the
transcriptional condensates redundant [72].
Significant insights have been generated using
an optogenetic system designed to induce biomolecu-
lar condensates based on the IDRs of FET family tran-
scriptional regulators [such as Fused in sarcoma (FUS),
Ewings sarcoma (EWS), and TAF15] [73]. These IDRs
can initiate phase separation process in the living cells
with TAF15 demonstrating capacity to interact with
the RNAPII CTDs and attract RNAPII molecules to form
condensates. Notably, nascent CTD clusters at primed
genomic loci lower the energy barrier for growth of
the TAF15 condensates, which, in turn, further recruit
RNAPII to initiate transcription. Thus, the authors have
identified a positive feedback mechanism that enhanc-
es RNAPII accumulation at transcription sites, leading
to improved transcription efficiency. At the same time,
the data suggest that elongation occurs outside of the
induced condensates, as the RNAPII molecules with
unmodified CTD colocalize with the condensates con-
sisting of IDRs of TAF15, and phosphorylated CTD is re-
leased into the surrounding region.
Therefore, the existing data demonstrate that non-
specific interactions involving the disordered CTD of
RNAPII can enhance transcription efficiency. It is cru-
cial to note, however, that many of the data providing
information on the mechanisms underlying formation
of transcriptional condensates and role of IDRs are
derived from the in vitro experiments or experiments
with isolated IDRs, rather than the whole proteins.
Within the intact proteins, IDRs may exhibit different
behavior. For example, recent findings suggest that the
IDRs alone could be insufficient for inducing transcrip-
tion factor clustering [74], so this issue needs further
investigation.
TWO TYPES OF TRANSCRIPTIONAL
CONDENSATES PROVIDE SEPARATION
OF INITIATION AND ELONGATION PROCESSES
Recent advancements in super-resolution micros-
copy techniques enable estimation of the distances
between transcription factors and nascent transcripts.
For example, in the case of the Pou5f1 and Nanog genes,
nascent transcripts were found to be in close proximi-
ty to the elongating RNAPII molecules but distant from
the loci enriched with the SOX2 and BRD4 transcrip-
tion factors [75]. These findings imply spatial separa-
tion between the elongation and initiation complexes.
VALYAEVA, SHEVAL692
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
Fig. 2. Models of initiation and elongation condensates that provide regulation of transcription by RNA polymeraseII (RNAPII).
Intriguingly, results obtained with the HIV-1 reporter
gene indicate that the initiation complex (initiation
condensate) exists for a relatively short time (~1min),
during which RNAPII molecules are recruited to the
promoter region [76]. Apparently, during transcription
and subsequent processes such as splicing and nuclear
transport, the molecules travel through several con-
densates, ensuring high efficiency at each stage of the
RNA molecule life cycle[77].
To date, sufficiently compelling data confirming ex-
istence of condensates where the components involved
in transcription initiation are concentrated have been
obtained (Fig. 2). The situation regarding existence of
the elongation condensates is somewhat more contro-
versial. As mentioned earlier, long genes form extend-
ed loops along which RNAPII molecules move during
transcription [35, 36]. It is likely that the described
patterns could be extended to all other genes, but this
requires experimental confirmation, focusing on the
key aspects. Nevertheless, these morphological obser-
vations do not align well with the idea of isolated elon-
gation condensates. Further experiments will be able
to provide clarity on these findings and reconcile the
existing contradictions.
ROLE OF NONSPECIFIC INTERACTIONS
IN THE ORGANIZATION
OF THE PRE-rRNA TRANSCRIPTION
Nucleolus is the product of transcription from the
relatively small regions of the genome containing nu-
merous rRNA genes, referred to as nucleolus organizer
regions [78]. Around 80 ribosomal proteins and over
200 pre-ribosome assembly factors accumulate in the
nucleoli [79]. Additionally, nucleolus harbors a sub-
stantial number of other proteins unrelated or only in-
directly related to its primary functionproduction of
pre-ribosomes [80].
Unlike most of the nuclear bodies, nucleoli exhib-
it relatively complex internal organization. Active nu-
cleoli in mammals, birds, and some reptiles consist of
three distinct subcompartments: fibrillar center (FC),
dense fibrillar component (DFC), and granular com-
ponent(GC), clearly visible in electron microscopy im-
ages [81, 82]. The DFC and GC form during ribosome
biogenesis.
Transcription of rDNA occurs at the boundary of
FC and DFC; consequently, DFC represents a cluster of
proteins involved in the early pre-rRNA processing.
Central role in maintenance of structural integrity of
this nucleolar subcompartment is played by fibrillarin
(FBL), which binds to the pre-rRNA molecules during
transcription [83]. Thus, the transcribed pre-rRNA mol-
ecules nucleate assembly of DFC, where both tran-
scription and initial steps of pre-rRNA processing take
place simultaneously. It must be noted that this pro-
cess relies on the disordered glycine and arginine-rich
N-terminal region of FBL, known as GAR domain [83].
FBL, a highly conserved methyltransferase, possesses an
amino acid sequence and 3D structure that have under-
gone minimal changes from archaea to higher eukary-
otes [84]. Emergence of the GAR domain has enabled
FBL to acquire additional functions essential for its role
in the significantly more complex eukaryotic cells. Spe-
cifically, this domain serves as a nuclear localization
signal [85]. Another potential function is involvement
of the GAR domain in organization of the DFC [83].
GC is traditionally viewed as a compartment dedi-
cated to pre-rRNA processing and pre-ribosome assem-
bly. Structural integrity of the GC relies on dynamic
interactions between the pre-ribosomes with NPM1
playing a pivotal role in this process [86]. During mat-
uration, properties of the GC change to facilitate di-
rected transport of pre-ribosomes from the DFC to the
nucleolus periphery [87].
Similar to the RNAPII transcriptional condensates,
multiple condensates are formed inside the nucleolus,
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 693
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
determining directional movement of the synthesized
pre-rRNA molecules.
ROLE OF THE SYNTHESIZED TRANSCRIPTS
IN THE REGULATION OF GENE EXPRESSION
Initiation of phase separation processes during
the biomolecular condensate formation often relies on
the RNA molecules serving as successful substrates for
multivalent interactions. The subsequent discussion
will highlight key studies suggesting involvement of
RNA molecules in the low-specificity interactions cru-
cial for transcription regulation.
Variety of RNAs and transcription process itself
influence three-dimensional structure of chromatin
and gene expression mediated by the chromatin state
[88-92]. Active transcription and accumulation of the
chromatin-associated transcripts at the gene locus, in-
cluding pre-mRNAs and other nascent RNAs, maintain
open state of the chromatin through a positive feed-
back mechanism. Transcripts regulate chromatin state
and associated gene expression through electrostatic
interactions with the similarly charged DNA and oppo-
sitely charged histone proteins [93]. Additionally, they
bind to various protein factors comprising chroma-
tin[94].
Transcription and chromatin remodeling factors
YY1, CTCF, DNMT1, and DNMT3A exemplify DNA-bind-
ing factors whose association with chromatin can be
regulated by nascent transcripts [95]. For the ubiqui-
tously expressed transcription factor YY1, interaction
with the nascent RNAs stabilizes its binding with chro-
matin, predominantly at promoter sites [96]. In con-
trast, nascent RNAs have an inhibitory effect on asso-
ciation of the DNMT1 and DNMT3A methyltransferases
with chromatin [97, 98]. For the DNMT1, which per-
forms maintenance methylation of CpG islands, inter-
action with various nascent transcripts [99], including
of its own mRNA [100], has been demonstrated using
the whole-genome RNA–protein interactome analysis
methods (fRIP-seq). Interestingly, DNMT1 prefers bind-
ing to the non-canonical G-quadruplex structure called
pUG-fold, with binding affinity appearing proportional
to the length of the GU repeat [100]. Interaction with
the RNA pUG-fold prevents DNMT1 from binding to
DNA, thereby inhibiting its enzyme activity. Converse-
ly, certain long non-coding RNAs attract methyltrans-
ferase to the specific chromatin regions without inhib-
iting its activity [101-105], and could even determine
its cellular localization [106].
Antagonistic relationships with the nascent tran-
scripts are also observed for the Polycomb repressive
complex 2 (PRC2) [107]. PRC2, an epigenetic transcrip-
tion regulation complex, methylates lysine 27 of the
histone H3, participating in the repression of gene
expression. Due to its function, PRC2 is primarily as-
sociated with CpG islands in the repressed genes.
High-throughput sequencing of the RNAs interacting
with PRC2 components (CLIP/iCLIP-seq) revealed PRC2
interactions with the pre-mRNAs of the majority of
active genes [107, 108]. Inhibition of RNAPII or total
RNA degradation in the cell led to the recruitment of
PRC2 to active genes [109-111], with this effect being
reversible [112, 113]. In vitro experiments demonstrat-
ed RNA competition with nucleosomes for PRC2 bind-
ing and inhibition of PRC2 catalytic activity [107,
114-116]. Interestingly, higher affinity was shown for
the RNAs forming G-quadruplex structures [112, 117].
Bytitrating PRC2 complexes with themselves, nascent
transcripts reduce likelihood of these repressive com-
plexes binding to chromatin, thereby preventing gene
silencing. It has been hypothesized that all PRC2–chro-
matin interactions are mediated by the RNA “bridge”
[118]. However, evidence from the rChIP-seq exper-
iments indicating necessity of RNA presence in the
complex with PRC2 for chromatin localization seems
to be an artifact of the experiment and normalization.
Later studies demonstrated that treating the immuno-
precipitated chromatin with RNase A and sonication
in a low ionic strength solution reduced specificity of
immunoprecipitation and increased background sig-
nal of nonspecific chromatin precipitation [119, 120].
Therefore, active transcription leading to formation of
new transcripts that bind and immobilize repressive
epigenetic regulators maintains itself.
Involvement of the nascent transcripts in regu-
lation of the binding of chromatin-associated protein
factors to chromatin seems to be a more widespread
mechanism for controlling the structure and state of
chromatin than previously thought. For example, ex-
periments involving inhibition of transcription and
RNA degradation in the cells have demonstrated that
disappearance of RNA from chromatin coincides with
the changes in the chromatin-associated proteome
[121, 122]. Transcriptional inhibition and concurrent
decrease in the number of nascent transcripts result
in the recruitment of chromatin modifier proteins and
chromatin remodeling factors (DNA methyltransferas-
es of the DNMT family, EHMT1/2, MLL2/SET1A, HUSH,
NuRD, NURF, NoRC, CHRAC, NuA4, INO80, BAF, ATRX/
DAXX, cohesin, CTCF, SMCHD1, SAFB) and transcrip-
tion factors (POU5F1, ZFP57, UBTF, TP53, MYBL2, and
UTF1) to chromatin. Meanwhile the number of chro-
matin-associated RNA processing factors were lost
from the chromatin during RNAPII inhibition. A sim-
ilar qualitative change in the chromatin-associated
proteome occurs in the case of RNA degradation [121].
Furthermore, significant proportion of the chro-
matin- associated proteins are known as RNA-bind-
ing proteins [122]. These proteins, often have IDRs or
low complexity sequences, and can, through binding
VALYAEVA, SHEVAL694
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
to nascent transcripts in close proximity to RNAPII,
facilitate formation of biomolecular condensates via
the phase separation mechanism, promoting concen-
tration of the factors necessary for gene expression at
the transcription site.
ROLE OF THE SYNTHESIZED TRANSCRIPTS
IN THE FORMATION
OF TRANSCRIPTIONAL CONDENSATES
Numerous transcription factors interacting with
chromatin at regulatory sites (promoters, enhancers)
possess, along with DNA-binding and effector domains,
a conserved RNA-binding domain resembling the argi-
nine-rich motif (ARM) of the HIV-1 Tat protein [123].
Integrated analysis of the DNA–protein interactions
(ChIP-seq) and RNA–protein interactomes (CLIP-seq) of
several transcription factors (GATA1, YY1, and CTCF)
revealed interactions of these factors with the RNA
originating from the loci near the factor binding sites
on chromatin (enhancer, promoter, and nascent tran-
scripts). Mutations in the ARM-like domain led to the
decreased expression of target genes, while its deletion
increased proportion of the freely diffusing transcrip-
tion factors in the nucleus. These findings suggest that
interactions of transcription factors with the nascent
RNAs via the ARM-like domain may contribute to for-
mation of the transcriptional condensates and provide
precise regulation of gene expression.
In line with the aforementioned observations, a
model for transcriptional condensate formation me-
diated by interaction of the nascent RNAs and protein
factors possessing RNA-binding domains and unstruc-
tured regions that promote formation of biomolecu-
lar condensates (RNA-mediated feedback model) ap-
pears attractive [122, 124]. In the experiments utilizing
in vitro and in silico systems, it was demonstrated that
the RNA molecules, both from promoters or enhancers,
irrespective of their sequence, promoted formation
of transcriptional condensates involving the Media-
tor complex [125]. However, this effect was observed
only within a limited range of RNA concentrations,
with the fixed concentration of the Mediator complex.
At such RNA concentrations, the system reached an
equilibrium state between the negative charge of nu-
cleic acid, proportional to its length, and the positive
charge of protein factors. Increase in the amount of
RNA carrying a negative charge disrupted the equilib-
rium and led to the condensate dissolution. Validity
of the RNA-mediated feedback model proposed by the
authors was also demonstrated in the in vivo experi-
ments. Transcription initiation and production of the
short nascent transcripts led to formation of the tran-
scriptional condensate and simultaneous increase in
the reporter gene expression. However, when an ex-
cessive level of transcription and a threshold concen-
tration of the nascent RNAs were reached, expression
of the reporter gene decreased, and the temporarily
formed transcriptional condensates resulting from
nonspecific electrostatic interactions between the syn-
thesized transcripts and protein factors were dissolved
[124-126]. This mechanism of gene expression regula-
tion is realized in enhancers through the short-lived
enhancer RNAs [125, 126]. This mechanism can explain
transcriptional bursts observed for many genes in the
cases when the enhancer- and promoter-associated
condensates interact or fuse [126, 127].
CONCLUSION
The data accumulated in recent years indicate
that transcription is accompanied by formation of the
molecular complexes with non-stoichiometric composi-
tion – transcriptional condensates. Apparently, forma-
tion of such condensates is determined by combination
of the high-affinity specific interactions and nonspecif-
ic interactions between the proteins, RNA, and DNA
sites that are part of these condensates. Aswith other
described biomolecular condensates, transcriptional
condensates that contain RNAPII are extremely dy-
namic. In contrast, transcriptional condensates formed
as a result of RNAPI activity form stable membraneless
structures – nucleoli. Nevertheless, in many respects,
formation of the transcriptional condensates is similar
in the cases of RNAPI and RNAPII.
The IDRs of proteins engaged in transcription
play a pivotal role in formation of the transcription-
al condensates. Although these domains participate
in nonspecific interactions, they significantly enhance
efficiency and specificity of the condensate formation
during transcription. This applies to various process-
es, including both search for the target sites in the
nucleus and interaction between the individual com-
ponents. Notably, positive feedback mechanisms may
be at play, wherein molecular interactions lead to for-
mation of the transcriptional condensate facilitating
attraction of additional molecules of the same proteins
and potentially other molecules involved in the pro-
cess. This phenomenon is particularly evident in the
nucleolus, which attracts a diverse array of molecules
[80] contributing to its multifunctionality [128].
Nascent transcripts play a crucial role in the tran-
scription regulation. Transcripts still bound to poly-
merase or RNAs in the transcriptional condensate
due to the large negative charge of their sugar-phos-
phate backbone, can attract RNA-binding proteins to
the transcription site through nonspecific interactions
with the positively charged IDRs, along with specific
interactions facilitated by the RNA secondary struc-
tures. Varying length of RNA, ranging from a few tens
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 695
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
to several thousand nucleotides for different RNA
types, enables multivalent interactions with proteins.
Ability of the nascent transcripts to attract protein
factors to the transcription site can be used to de-
velop tools for gene expression regulation involving
CRISPR-Cas9 systems with activator domains of the
transcription factors [129].
However, it should be noted that many concepts
lack complete experimental validation. For example,
there is only limited data available on the structure of
elongation transcriptional condensates. Some morpho-
logical observations suggest that elongation of some
genes occurs on the extended DNA strands resembling
loops of the lampbrush chromosomes [35, 36]. While
these observations do not definitively contradict the
idea of the existence of elongation condensates, they
introduce some uncertainty. Moreover, consideration
should be given to the data on the transcriptional con-
densate dissolution promoted by high levels of the tran-
scribed RNA molecules [124-126]. Disassembly of such
condensate may facilitate movement of the RNAPII
molecules along the DNA strand. These questions re-
quire further investigation.
Thus, high efficiency and flexible regulation of
transcription largely stem from the weak, non-specif-
ic, yet multiple interactions that give rise to formation
of the highly dynamic at the molecular level transcrip-
tional condensates.
Contributions. A.A.V. and E.V.Sh. collected and an-
alyzed information, wrote and edited the manuscript.
Funding. The work was financially supported
by the Russian Science Foundation (project no. 21-74-
20134).
Ethics declarations. This work does not contain
any studies involving human and animal subjects.
Theauthors of this work declare that they have nocon-
flicts of interest.
REFERENCES
1. Scheer,U., and Hock,R. (1999) Structure and function
of the nucleolus, Curr. Opin. Cell Biol., 11, 385-390,
doi:10.1016/S0955-0674(99)80054-4.
2. Hernandez-Verdun, D., Roussel,P., Thiry, M., Sirri,V.,
and Lafontaine, D. L. J. (2010) The nucleolus: struc-
ture/function relationship in RNA metabolism, Wiley
Interdiscip. Rev. RNA, 1, 415-431, doi:10.1002/wrna.39.
3. McNally, J.G., Müller, W.G., Walker,D., Wolford,R., and
Hager, G.L. (2000) The glucocorticoid receptor: rapid
exchange with regulatory sites in living cells, Science,
287, 1262-1265, doi:10.1126/science.287.5456.1262.
4. Phair, R.D., and Misteli,T. (2000) High mobility of pro-
teins in the mammalian cell nucleus, Nature, 404, 604-
609, doi:10.1038/35007077.
5. Becker,M., Baumann,C., John,S., Walker, D.A., Vign-
eron,M., etal. (2002) Dynamic behavior of transcrip-
tion factors on a natural promoter in living cells,
EMBO Rep., 3, 1188-1194, doi: 10.1093/embo-reports/
kvf244.
6. Phair, R. D., Scaffidi, P., Elbi, C., Vecerová, J., Dey, A.,
et al. (2004) Global nature of dynamic protein-chro-
matin interactions in vivo: three-dimensional ge-
nome scanning and dynamic interaction networks
of chromatin proteins, Mol. Cell. Biol., 24, 6393-6402,
doi:10.1128/MCB.24.14.6393-6402.2004.
7. Kimura,H., Sugaya,K., and Cook, P.R. (2002) The tran-
scription cycle of RNA polymerase II in living cells,
J.Cell Biol., 159, 777-782, doi:10.1083/jcb.200206019.
8. Chen,D., and Huang,S. (2001) Nucleolar components
involved in ribosome biogenesis cycle between the
nucleolus and nucleoplasm in interphase cells, J.Cell
Biol., 153, 169-176, doi:10.1083/jcb.153.1.169.
9. Dundr, M., Hoffmann-Rohrer, U., Hu, Q., Grummt, I.,
Rothblum, L.I., et al. (2002) A kinetic framework for
a mammalian RNA polymerase in vivo, Science, 298,
1623-1626, doi:10.1126/science.1076164.
10. Stixová, L., Bártová, E., Matula, P., Daněk, O., Legar-
tová, S., and Kozubek, S. (2011) Heterogeneity in the
kinetics of nuclear proteins and trajectories of sub-
structures associated with heterochromatin, Epi-
genetics Chromatin, 4, 5, doi:10.1186/1756-8935-4-5.
11. Mirny, L., Slutsky, M., Wunderlich, Z., Tafvizi, A.,
Leith, J., and Kosmrlj, A. (2009) How a protein
searches for its site on DNA: the mechanism of facil-
itated diffusion, J. Phys. A Math. Theor., 42, 434013,
doi:10.1088/1751-8113/42/43/434013.
12. Jana,T., Brodsky,S., and Barkai,N. (2021) Speed-speci-
ficity trade-offs in the transcription factors search for
their genomic binding sites, Trends Genet., 37, 421-432,
doi:10.1016/j.tig.2020.12.001.
13. Richter, W. F., Nayak, S., Iwasa, J., and Taatjes, D. J.
(2022) The Mediator complex as a master regulator of
transcription by RNA polymeraseII, Nat. Rev. Mol. Cell
Biol., 23, 732-749, doi:10.1038/s41580-022-00498-3.
14. Rengachari, S., Schilbach, S., and Cramer, P. (2023)
Mediator structure and function in transcription
initiation, Biol. Chem., 404, 829-837, doi: 10.1515/
hsz-2023-0158.
15. Bregman, D. B., Du, L., van der Zee, S., and Warren,
S.L. (1995) Transcription-dependent redistribution of
the large subunit of RNA polymeraseII to discrete nu-
clear domains, J.Cell Biol., 129, 287-298, doi:10.1083/
jcb.129.2.287.
16. Van Steensel, B., Brink, M., van der Meulen, K., van
Binnendijk, E. P., Wansink, D. G., et al. (1995) Local-
ization of the glucocorticoid receptor in discrete clus-
ters in the cell nucleus, J. Cell Sci., 108, 3003-3011,
doi:10.1242/jcs.108.9.3003.
17. Iborra, F.J., Pombo,A., Jackson, D.A., and Cook, P.R.
(1996) Active RNA polymerases are localized within
VALYAEVA, SHEVAL696
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
discrete transcription “factories” in human nuclei,
J.Cell Sci., 109, 1427-1436, doi:10.1242/jcs.109.6.1427.
18. Grande, M.A., van der Kraan,I., de Jong,L., and van
Driel, R. (1997) Nuclear distribution of transcription
factors in relation to sites of transcription and RNA
polymeraseII, J.Cell Sci., 110, 1781-1791, doi:10.1242/
jcs.110.15.1781.
19. Pombo, A., Jackson, D. A., Hollinshead, M., Wang, Z.,
Roeder, R. G., and Cook, P. R. (1999) Regional special-
ization in human nuclei: visualization of discrete
sites of transcription by RNA polymeraseIII, EMBOJ.,
18, 2241-2253, doi:10.1093/emboj/18.8.2241.
20. Osborne, C.S., Chakalova,L., Brown, K.E., Carter, D.,
Horton,A., et al. (2004) Active genes dynamically co-
localize to shared sites of ongoing transcription, Nat.
Genet., 36, 1065-1071, doi:10.1038/ng1423.
21. Castells-Garcia, A., Ed-Daoui, I., González-Almela, E.,
Vicario, C., Ottestrom, J., et al. (2022) Super resolu-
tion microscopy reveals how elongating RNA poly-
meraseII and nascent RNA interact with nucleosome
clutches, Nucleic Acids Res., 50, 175-190, doi:10.1093/
nar/gkab1215.
22. Cho, W.-K., Spille, J.-H., Hecht,M., Lee,C., Li,C., etal.
(2018) Mediator and RNA polymerase II clusters asso-
ciate in transcription-dependent condensates, Science,
361, 412-415, doi:10.1126/science.aar4199.
23. Hnisz,D., Abraham, B.J., Lee, T. I., Lau, A., Saint-An-
dré, V., et al. (2013) Super-enhancers in the con-
trol of cell identity and disease, Cell, 155, 934-947,
doi:10.1016/j.cell.2013.09.053.
24. Whyte, W. A., Orlando, D. A., Hnisz, D., Abraham,
B. J., Lin, C. Y., et al. (2013) Master transcription fac-
tors and mediator establish super-enhancers at key
cell identity genes, Cell, 153, 307-319, doi: 10.1016/
j.cell.2013.03.035.
25. Lovén,J., Hoke, H.A., Lin, C.Y., Lau,A., Orlando, D.A.,
et al. (2013) Selective inhibition of tumor oncogenes
by disruption of super-enhancers, Cell, 153, 320-334,
doi:10.1016/j.cell.2013.03.036.
26. Mansour, M.R., Abraham, B.J., Anders,L., Berezovska-
ya, A., Gutierrez, A., et al. (2014) Oncogene regula-
tion. An oncogenic super-enhancer formed through
somatic mutation of a noncoding intergenic ele-
ment, Science, 346, 1373-1377, doi: 10.1126/science.
1259037.
27. Wang, X., Cairns, M. J., and Yan, J. (2019) Super-en-
hancers in transcriptional regulation and genome
organization, Nucleic Acids Res., 47, 11481-11496,
doi:10.1093/nar/gkz1038.
28. Sabari, B. R., Dall’Agnese, A., Boija, A., Klein, I. A.,
Coffey, E. L., et al. (2018) Coactivator condensation
at super-enhancers links phase separation and gene
control, Science, 361, eaar3958, doi: 10.1126/science.
aar3958.
29. Guo, Y.E., Manteiga, J.C., Henninger, J.E., Sabari, B.R.,
Dall’Agnese, A., et al. (2019) Pol II phosphorylation
regulates a switch between transcriptional and splic-
ing condensates, Nature, 572, 543-548, doi: 10.1038/
s41586-019-1464-0.
30. Shrinivas, K., Sabari, B. R., Coffey, E. L., Klein, I. A.,
Boija,A., etal. (2019) Enhancer features that drive for-
mation of transcriptional condensates, Mol. Cell, 75,
549-561.e7, doi:10.1016/j.molcel.2019.07.009.
31. Chen,X., Wei,M., Zheng, M.M., Zhao,J., Hao,H., etal.
(2016) Study of RNA polymerase II clustering inside
live-cell nuclei using bayesian nanoscopy, ACS Nano,
10, 2447-2454, doi:10.1021/acsnano.5b07257.
32. Cisse, I. I., Izeddin, I., Causse, S. Z., Boudarene, L.,
Senecal, A., et al. (2013) Real-time dynamics of RNA
polymeraseII clustering in live human cells, Science,
341, 664-667, doi:10.1126/science.1239053.
33. Cho, W.-K., Jayanth, N., English, B. P., Inoue, T., An-
drews, J. O., et al. (2016) RNA Polymerase II cluster
dynamics predict mRNA output in living cells, Elife,
5, e13617, doi:10.7554/eLife.13617.
34. Mir,M., Stadler, M.R., Ortiz, S.A., Hannon, C.E., Har-
rison, M. M., et al. (2018) Dynamic multifactor hubs
interact transiently with sites of active transcription
in Drosophila embryos, Elife, 7, e40497, doi: 10.7554/
eLife.40497.
35. Leidescher,S., Ribisel,J., Ullrich,S., Feodorova,Y., Hil-
debrand,E., etal. (2022) Spatial organization of tran-
scribed eukaryotic genes, Nat. Cell Biol., 24, 327-339,
doi:10.1038/s41556-022-00847-6.
36. Ball, M.L., Koestler, S.A., Muresan,L., Rehman, S.A.,
O’Holleran, K., and White, R. (2023) The anatomy of
transcriptionally active chromatin loops in Drosoph-
ila primary spermatocytes using super-resolution
microscopy, PLoS Genet., 19, e1010654, doi: 10.1371/
journal.pgen.1010654.
37. Morgan, G. T. (2007) Localized co-transcriptional re-
cruitment of the multifunctional RNA-binding protein
CELF1 by lampbrush chromosome transcription units,
Chromosome Res., 15, 985-1000, doi: 10.1007/s10577-
007-1179-1.
38. Morgan, G. T. (2018) Imaging the dynamics of tran-
scription loops in living chromosomes, Chromosoma,
127, 361-374, doi:10.1007/s00412-018-0667-8.
39. Krasikova, A., Fishman, V., and Kulikova, T. (2023)
Lampbrush chromosome studies in the post-genom-
ic era, Bioessays, 45, e2200250, doi: 10.1002/bies.
202200250.
40. Rippe,K., and Papantonis,A. (2021) RNA polymeraseII
transcription compartments: from multivalent chro-
matin binding to liquid droplet formation? Nat.
Rev. Mol. Cell Biol., 22, 645-646, doi: 10.1038/s41580-
021-00401-6.
41. Brangwynne, C.P., Eckmann, C.R., Courson, D.S., Ry-
barska, A., Hoege, C., et al. (2009) Germline P gran-
ules are liquid droplets that localize by controlled
dissolution/condensation, Science, 324, 1729-1732,
doi:10.1126/science.1172046.
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 697
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
42. Shin, Y., and Brangwynne, C. P. (2017) Liquid phase
condensation in cell physiology and disease, Science,
357, eaaf4382, doi:10.1126/science.aaf4382.
43. Bhat,P., Honson, D., and Guttman, M. (2021) Nuclear
compartmentalization as a mechanism of quantitative
control of gene expression, Nat. Rev. Mol. Cell Biol.,
22, 653-670, doi:10.1038/s41580-021-00387-1.
44. Elbaum-Garfinkle, S., Kim, Y., Szczepaniak, K.,
Chen,C.C.-H., Eckmann, C.R., etal. (2015) The disor-
dered P granule protein LAF-1 drives phase separation
into droplets with tunable viscosity and dynamics,
Proc. Natl. Acad. Sci. USA, 112, 7189-7194, doi:10.1073/
pnas.1504822112.
45. Molliex,A., Temirov,J., Lee, J., Coughlin,M., Kanaga-
raj, A. P., et al. (2015) Phase separation by low com-
plexity domains promotes stress granule assembly
and drives pathological fibrillization, Cell, 163, 123-
133, doi:10.1016/j.cell.2015.09.015.
46. Nott, T. J., Petsalaki, E., Farber, P., Jervis, D., Fuss-
ner, E., et al. (2015) Phase transition of a disordered
nuage protein generates environmentally respon-
sive membraneless organelles, Mol. Cell, 57, 936-947,
doi:10.1016/j.molcel.2015.01.013.
47. Pak, C.W., Kosno,M., Holehouse, A.S., Padrick, S.B.,
Mittal,A., etal. (2016) Sequence determinants of intra-
cellular phase separation by complex coacervation of
a disordered protein, Mol. Cell, 63, 72-85, doi:10.1016/
j.molcel.2016.05.042.
48. Li,P., Banjade,S., Cheng, H.-C., Kim,S., Chen,B., etal.
(2012) Phase transitions in the assembly of multivalent
signalling proteins, Nature, 483, 336-340, doi:10.1038/
nature10879.
49. Banani, S.F., Rice, A.M., Peeples, W.B., Lin,Y., Jain,S.,
et al. (2016) Compositional control of phase-separat-
ed cellular bodies, Cell, 166, 651-663, doi: 10.1016/
j.cell.2016.06.010.
50. Feric, M., Vaidya, N., Harmon, T. S., Mitrea, D. M.,
Zhu, L., et al. (2016) Coexisting liquid phases under-
lie nucleolar subcompartments, Cell, 165, 1686-1697,
doi:10.1016/j.cell.2016.04.047.
51. Lu,Y., Wu,T., Gutman,O., Lu,H., Zhou,Q., etal. (2020)
Phase separation of TAZ compartmentalizes the tran-
scription machinery to promote gene expression, Nat.
Cell Biol., 22, 453-464, doi:10.1038/s41556-020-0485-0.
52. Boija, A., Klein, I. A., Sabari, B. R., Dall’Agnese, A.,
Coffey, E. L., et al. (2018) Transcription factors acti-
vate genes through the phase-separation capacity of
their activation domains, Cell, 175, 1842-1855.e16,
doi:10.1016/j.cell.2018.10.042.
53. Altmeyer, M., Neelsen, K. J., Teloni, F., Pozdnyako-
va, I., Pellegrino, S., et al. (2015) Liquid demixing
of intrinsically disordered proteins is seeded by
poly(ADP-ribose), Nat. Commun., 6, 8088, doi:10.1038/
ncomms9088.
54. Zhao, B., Li, Z., Yu, S., Li, T., Wang, W., et al. (2023)
LEF1 enhances β-catenin transactivation through IDR-
dependent liquid-liquid phase separation, Life Sci Alli-
ance,
6, e202302118, doi:10.26508/lsa.202302118.
55. Liu,J., Perumal, N. B., Oldfield, C. J., Su, E.W., Uver-
sky, V. N., and Dunker, A.K. (2006) Intrinsic disorder
in transcription factors, Biochemistry, 45, 6873-6888,
doi:10.1021/bi0602718.
56. Minezaki, Y., Homma, K., Kinjo, A. R., and Nishika-
wa, K. (2006) Human transcription factors contain a
high fraction of intrinsically disordered regions es-
sential for transcriptional regulation, J.Mol. Biol., 359,
1137-1149, doi:10.1016/j.jmb.2006.04.016.
57. Staby,L., O’Shea,C., Willemoës,M., Theisen,F., Krage-
lund, B.B., and Skriver,K. (2017) Eukaryotic transcrip-
tion factors: paradigms of protein intrinsic disorder,
Biochem.J., 474, 2509-2532, doi:10.1042/BCJ20160631.
58. Brodsky, S., Jana, T., Mittelman, K., Chapal, M., Ku-
mar, D. K., et al. (2020) Intrinsically disordered re-
gions direct transcription factor in vivo binding
specificity, Mol. Cell, 79, 459-471.e4, doi: 10.1016/
j.molcel.2020.05.032.
59. Sandra, U. S., Shukla, A., and Kolthur-Seetharam, U.
(2020) Search and capture: disorder rules gene
promoter selection, Trends Genet., 36, 721-722,
doi:10.1016/j.tig.2020.07.005.
60. Jonas, F., Carmi, M., Krupkin, B., Steinberger, J.,
Brodsky, S., et al. (2023) The molecular grammar of
protein disorder guiding genome-binding locations,
Nucleic Acids Res., 51, 4831-4844, doi: 10.1093/nar/
gkad184.
61. Meinhart, A., Kamenski, T., Hoeppner, S., Baumli, S.,
and Cramer, P. (2005) A structural perspective of
CTD function, Genes Dev., 19, 1401-1415, doi:10.1101/
gad.1318105.
62. Eick,D., and Geyer,M. (2013) The RNA polymeraseII
carboxy-terminal domain (CTD) code, Chem. Rev., 113,
8456-8490, doi:10.1021/cr400071f.
63. Zaborowska, J., Egloff, S., and Murphy, S. (2016) The
polII CTD: new twists in the tail, Nat. Struct. Mol. Biol.,
23, 771-777, doi:10.1038/nsmb.3285.
64. Lu, F., Portz, B., and Gilmour, D. S. (2019) The C-ter-
minal domain of RNA polymeraseII is a multivalent
targeting sequence that supports Drosophila devel-
opment with only consensus heptads, Mol. Cell, 73,
1232-1242.e4, doi:10.1016/j.molcel.2019.01.008.
65. Bernecky,C., Herzog,F., Baumeister,W., Plitzko, J.M.,
and Cramer,P. (2016) Structure of transcribing mam-
malian RNA polymerase II, Nature, 529, 551-554,
doi:10.1038/nature16482.
66. Zehring, W. A., Lee, J. M., Weeks, J. R., Jokerst, R. S.,
and Greenleaf, A.L. (1988) The C-terminal repeat do-
main of RNA polymerase II largest subunit is essen-
tial in vivo but is not required for accurate transcrip-
tion initiation in vitro, Proc. Natl. Acad. Sci. USA, 85,
3698-3702, doi:10.1073/pnas.85.11.3698.
67. Yahia,Y., Pigeot,A., El Aabidine, A.Z., Shah,N., Kara-
su,N., etal. (2023) RNA polymerase II CTD is dispens-
VALYAEVA, SHEVAL698
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
able for transcription and required for termination
in human cells, EMBO Rep., 24, e56150, doi:10.15252/
embr.202256150.
68. Boehning,M., Dugast-Darzacq, C., Rankovic,M., Han-
sen, A.S., Yu,T., etal. (2018) RNA polymerase II clus-
tering through carboxy-terminal domain phase sepa-
ration, Nat. Struct. Mol. Biol., 25, 833-840, doi:10.1038/
s41594-018-0112-y.
69. Quintero-Cadena,P., Lenstra, T.L., and Sternberg, P.W.
(2020) RNA Pol II length and disorder enable cooper-
ative scaling of transcriptional bursting, Mol. Cell, 79,
207-220.e8, doi:10.1016/j.molcel.2020.05.030.
70. Flores-Solis, D., Lushpinskaia, I. P., Polyansky, A. A.,
Changiarath, A., Boehning, M., et al. (2023) Driving
forces behind phase separation of the carboxy-termi-
nal domain of RNA polymeraseII, Nat. Commun., 14,
5979, doi:10.1038/s41467-023-41633-8.
71. Jaeger, M. G., Schwalb, B., Mackowiak, S. D., Velych-
ko,T., Hanzl,A., etal. (2020) Selective Mediator depen-
dence of cell-type-specifying transcription, Nat. Genet.,
52, 719-727, doi:10.1038/s41588-020-0635-0.
72. Ling, Y. H., Ye, Z., Liang, C., Yu, C., Park, G., et al. (2024)
Disordered C-terminal domain drives spatiotemporal
confinement of RNAPII to enhance search for chro-
matin targets, Nat. Cell Biol., 26, 581-592, doi: 10.1038/
s41556-024-01382-2.
73. Wei, M.-T., Chang, Y.-C., Shimobayashi, S. F., Shin, Y.,
Strom, A.R., and Brangwynne, C.P. (2020) Nucleated
transcriptional condensates amplify gene expression,
Nat. Cell Biol., 22, 1187-1196, doi: 10.1038/s41556-
020-00578-6.
74. Hannon, C. E., and Eisen, M. B. (2024) Intrinsic protein
disorder is insufficient to drive subnuclear clustering
in embryonic transcription factors, Elife, 12, RP88221,
doi: 10.7554/eLife.88221.
75. Li,J., Dong,A., Saydaminova,K., Chang,H., Wang,G.,
etal. (2019) Single-molecule nanoscopy elucidates RNA
polymeraseII transcription at single genes in live cells,
Cell, 178, 491-506.e28, doi:10.1016/j.cell.2019.05.029.
76. Forero-Quintero, L. S., Raymond, W., Handa, T., Sax-
ton, M.N., Morisaki,T., etal. (2021) Live-cell imaging
reveals the spatiotemporal organization of endoge-
nous RNA polymerase II phosphorylation at a single
gene, Nat. Commun., 12, 3158, doi: 10.1038/s41467-
021-23417-0.
77. Ishov, A.M., Gurumurthy,A., and Bungert,J. (2020) Co-
ordination of transcription, processing, and export of
highly expressed RNAs by distinct biomolecular con-
densates, Emerg. Top Life Sci., 4, 281-291, doi:10.1042/
ETLS20190160.
78. McClintock,B. (1934) The relation of a particular chro-
mosomal element to the development of the nucleoli
in Zea mays, Zeitschr. Zellforschung Mikroskopische
Anatomie, 21, 294-326, doi:10.1007/BF00374060.
79. Correll, C.C., Bartek,J., and Dundr,M. (2019) The nu-
cleolus: a multiphase condensate balancing ribosome
synthesis and translational capacity in health, ag-
ing and ribosomopathies, Cells,
8, 869, doi: 10.3390/
cells8080869.
80. Ahmad, Y., Boisvert, F.-M., Gregor, P., Cobley, A., and
Lamond, A.I. (2009) NOPdb: Nucleolar Proteome Da-
tabase-2008 update, Nucleic Acids Res., 37, D181-D184,
doi:10.1093/nar/gkn804.
81. Scheer, U., and Weisenberger, D. (1994) The nucle-
olus, Curr. Opin. Cell Biol., 6, 354-359, doi: 10.1016/
0955-0674(94)90026-4.
82. Lamaye,F., Galliot,S., Alibardi,L., Lafontaine, D.L.J.,
and Thiry,M. (2011) Nucleolar structure across evolu-
tion: the transition between bi- and tri-compartmen-
talized nucleoli lies within the class Reptilia, J.Struct.
Biol., 174, 352-359, doi:10.1016/j.jsb.2011.02.003.
83. Yao, R.-W., Xu,G., Wang,Y., Shan,L., Luan, P.-F., etal.
(2019) Nascent pre-rRNA sorting via phase separation
drives the assembly of dense fibrillar components
in the human nucleolus, Mol. Cell, 76, 767-783.e11,
doi:10.1016/j.molcel.2019.08.014.
84. Shubina, M.Y., Musinova, Y.R., and Sheval, E.V. (2016)
Nucleolar methyltransferase fibrillarin: evolution of
structure and function, Biochemistry (Moscow), 81,
941-950, doi:10.1134/S0006297916090030.
85. Shubina, M. Y., Arifulin, E. A., Sorokin, D. V., Sosina,
M. A., Tikhomirova, M. A., et al. (2020) The GAR do-
main integrates functions that are necessary for the
proper localization of fibrillarin (FBL) inside eukary-
otic cells, PeerJ, 8, e9029, doi:10.7717/peerj.9029.
86. Mitrea, D.M., Cika, J.A., Guy, C.S., Ban,D., Banerjee,
P. R., et al. (2016) Nucleophosmin integrates within
the nucleolus via multi-modal interactions with pro-
teins displaying R-rich linear motifs and rRNA, Elife,
5, e13571, doi:10.7554/eLife.13571.
87. Riback, J. A., Eeftens, J. M., Lee, D. S. W., Quinodoz,
S. A., Donlic, A., et al. (2023) Viscoelasticity and ad-
vective flow of RNA underlies nucleolar form and
function, Mol. Cell, 83, 3095-3107.e9, doi: 10.1016/
j.molcel.2023.08.006.
88. Heinz,S., Texari,L., Hayes, M.G.B., Urbanowski, M.,
Chang, M. W., et al. (2018) Transcription elonga-
tion can affect genome 3D structure, Cell, 174, 1522-
1536.e22, doi:10.1016/j.cell.2018.07.047.
89. Barutcu, A. R., Blencowe, B. J., and Rinn, J. L. (2019)
Differential contribution of steady-state RNA and ac-
tive transcription in chromatin organization, EMBO
Rep., 20, e48068, doi:10.15252/embr.201948068.
90. Saldaña-Meyer, R., Rodriguez-Hernaez, J., Escobar, T.,
Nishana, M., Jácome-López, K., et al. (2019) RNA in-
teractions are essential for CTCF-mediated genome
organization, Mol. Cell, 76, 412-422.e5, doi: 10.1016/
j.molcel.2019.08.015.
91. Quinodoz, S. A., Jachowicz, J. W., Bhat, P., Ollikain-
en,N., Banerjee, A.K., etal. (2021) RNA promotes the
formation of spatial compartments in the nucleus,
Cell, 184, 5775-5790.e30, doi:10.1016/j.cell.2021.10.014.
TRANSCRIPTION REGULATION AND TRANSCRIPTIONAL CONDENSATES 699
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
92. Boeren,J., and Gribnau, J. (2021) Xist-mediated chro-
matin changes that establish silencing of an entire
X chromosome in mammals, Curr. Opin. Cell Biol.,
70, 44-50, doi:10.1016/j.ceb.2020.11.004.
93. Dueva, R., Akopyan, K., Pederiva, C., Trevisan, D.,
Dhanjal,S., etal. (2019) Neutralization of the positive
charges on histone tails by RNA promotes an open
chromatin structure, Cell Chem. Biol., 26, 1436-1449.e5,
doi:10.1016/j.chembiol.2019.08.002.
94. Nozawa, R.-S., Boteva, L., Soares, D. C., Naughton, C.,
Dun, A. R., et al. (2017) SAF-A regulates interphase
chromosome structure through oligomerization with
chromatin-associated RNAs, Cell, 169, 1214-1227.e18,
doi:10.1016/j.cell.2017.05.029.
95. Skalska,L., Beltran-Nebot,M., Ule,J., and Jenner, R.G.
(2017) Regulatory feedback from nascent RNA to
chromatin and transcription, Nat. Rev. Mol. Cell Biol.,
18, 331-337, doi:10.1038/nrm.2017.12.
96. Sigova, A. A., Abraham, B. J., Ji, X., Molinie, B., Han-
nett, N. M., et al. (2015) Transcription factor trap-
ping by RNA in gene regulatory elements, Science,
350, 978-981, doi:10.1126/science.aad3346.
97. Holz-Schietinger,C., and Reich, N.O. (2012) RNA mod-
ulation of the human DNA methyltransferase 3A,
Nucleic Acids Res., 40, 8550-8557, doi: 10.1093/nar/
gks537.
98. Di Ruscio, A., Ebralidze, A. K., Benoukraf, T., Ama-
bile, G., Goff, L. A., et al. (2013) DNMT1-interacting
RNAs block gene-specific DNA methylation, Nature,
503, 371-376, doi:10.1038/nature12598.
99. Hendrickson, D. G., Kelley, D. R., Tenen, D., Bern-
stein,B., and Rinn, J.L. (2016) Widespread RNA bind-
ing by chromatin-associated proteins, Genome Biol.,
17, 28, doi:10.1186/s13059-016-0878-3.
100. Jansson-Fritzberg, L. I., Sousa, C. I., Smallegan, M. J.,
Song, J.J., Gooding, A.R., etal. (2023) DNMT1 inhibi-
tion by pUG-fold quadruplex RNA, RNA, 29, 346-360,
doi:10.1261/rna.079479.122.
101. Mohammad,F., Mondal,T., Guseva,N., Pandey, G. K.,
and Kanduri,C. (2010) Kcnq1ot1 noncoding RNA medi-
ates transcriptional gene silencing by interacting with
Dnmt1, Development, 137, 2493-2499, doi: 10.1242/
dev.048181.
102. Chalei,V., Sansom, S.N., Kong,L., Lee,S., Montiel, J.F.,
etal. (2014) The long non-coding RNA Dali is an epi-
genetic regulator of neural differentiation, Elife, 3,
e04530, doi:10.7554/eLife.04530.
103. Bao,X., Wu,H., Zhu,X., Guo,X., Hutchins, A.P., etal.
(2015) The p53-induced lincRNA-p21 derails somatic
cell reprogramming by sustaining H3K9me3 and CpG
methylation at pluripotency gene promoters, Cell Res.,
25, 80-92, doi:10.1038/cr.2014.165.
104. O’Leary, V. B., Hain, S., Maugg, D., Smida, J., Azimza-
deh,O., etal. (2017) Long non-coding RNA PARTICLE
bridges histone and DNA methylation, Sci. Rep., 7,
1790, doi:10.1038/s41598-017-01875-1.
105. Somasundaram, S., Forrest, M. E., Moinova, H., Co-
hen, A., Varadan, V., et al. (2018) The DNMT1-associ-
ated lincRNA DACOR1 reprograms genome-wide DNA
methylation in colon cancer, Clin. Epigenetics, 10, 127,
doi:10.1186/s13148-018-0555-3.
106. Jones, R., Wijesinghe, S., Wilson, C., Halsall, J., Lilo-
glou, T., and Kanhere, A. (2021) A long intergen-
ic non-coding RNA regulates nuclear localization
of DNA methyl transferase-1, iScience, 24, 102273,
doi:10.1016/j.isci.2021.102273.
107. Beltran,M., Yates, C.M., Skalska,L., Dawson,M., Reis,
F. P., et al. (2016) The interaction of PRC2 with RNA
or chromatin is mutually antagonistic, Genome Res.,
26, 896-907, doi:10.1101/gr.197632.115.
108. Kaneko,S., Son,J., Shen, S.S., Reinberg,D., and Bona-
sio,R. (2013) PRC2 binds active promoters and contacts
nascent RNAs in embryonic stem cells, Nat. Struct.
Mol. Biol., 20, 1258-1264, doi:10.1038/nsmb.2700.
109. Kaneko, S., Son, J., Bonasio, R., Shen, S. S., and Rein-
berg, D. (2014) Nascent RNA interaction keeps PRC2
activity poised and in check, Genes Dev., 28, 1983-1988,
doi:10.1101/gad.247940.114.
110. Riising, E.M., Comet,I., Leblanc,B., Wu,X., Johansen,
J.V., and Helin,K. (2014) Gene silencing triggers poly-
comb repressive complex2 recruitment to CpG islands
genome wide, Mol. Cell, 55, 347-360, doi: 10.1016/
j.molcel.2014.06.005.
111. Hosogane,M., Funayama,R., Shirota,M., and Nakaya-
ma,K. (2016) Lack of transcription triggers H3K27me3
accumulation in the gene body, Cell Rep., 16, 696-706,
doi:10.1016/j.celrep.2016.06.034.
112. Beltran, M., Tavares, M., Justin, N., Khandelwal, G.,
Ambrose, J., et al. (2019) G-tract RNA removes Poly-
comb repressive complex 2 from genes, Nat. Struct.
Mol. Biol., 26, 899-909, doi:10.1038/s41594-019-0293-z.
113. Garland, W., Comet, I., Wu, M., Radzisheuskaya, A.,
Rib,L., et al. (2019) A functional link between nucle-
ar RNA decay and transcriptional control mediated
by the polycomb repressive complex 2, Cell Rep., 29,
1800-1811.e6, doi:10.1016/j.celrep.2019.10.011.
114. Cifuentes-Rojas, C., Hernandez, A. J., Sarma, K., and
Lee, J.T. (2014) Regulatory interactions between RNA
and polycomb repressive complex 2, Mol. Cell, 55,
171-185, doi:10.1016/j.molcel.2014.05.009.
115. Zhang, Q., McKenzie, N. J., Warneford-Thomson, R.,
Gail, E. H., Flanigan, S. F., et al. (2019) RNA exploits
an exposed regulatory site to inhibit the enzymatic
activity of PRC2, Nat. Struct. Mol. Biol., 26, 237-247,
doi:10.1038/s41594-019-0197-y.
116. Song, J., Gooding, A. R., Hemphill, W. O., Love, B. D.,
Robertson, A., et al. (2023) Structural basis for inac-
tivation of PRC2 by G-quadruplex RNA, Science, 381,
1331-1337, doi:10.1126/science.adh0059.
117. Wang, X., Goodrich, K. J., Gooding, A. R., Naeem, H.,
Archer,S., etal. (2017) Targeting of polycomb repres-
sive complex2 to RNA by short repeats of consecutive
VALYAEVA, SHEVAL700
BIOCHEMISTRY (Moscow) Vol. 89 No. 4 2024
guanines, Mol. Cell, 65, 1056-1067.e5, doi: 10.1016/
j.molcel.2017.02.003.
118. Long, Y., Hwang, T., Gooding, A. R., Goodrich, K. J.,
Rinn, J. L., and Cech, T. R. (2020) RNA is essential
for PRC2 chromatin occupancy and function in hu-
man pluripotent stem cells, Nat. Genet., 52, 931-938,
doi:10.1038/s41588-020-0662-x.
119. Healy, E., Zhang, Q., Gail, E. H., Agius, S. C., Sun, G.,
etal. (2024) The apparent loss of PRC2 chromatin oc-
cupancy as an artefact of RNA depletion, Cell Rep., 43,
113858, doi: 10.1016/j.celrep.2024.113858.
120. Hickman, A.H., and Jenner, R.G. (2024) Apparent RNA
bridging between PRC2 and chromatin is an artefact
of non-specific chromatin precipitation upon RNA
degradation, Cell Rep., 43, 113856, doi: 10.1016/j.celrep.
2024.113856.
121. Skalska,L., Begley,V., Beltran,M., Lukauskas,S., Khan-
delwal, G., Faull, P., Bhamra, A., Tavares, M., Well-
man, R., Tvardovskiy, A., et al. (2021) Nascent RNA
antagonizes the interaction of a set of regulatory
proteins with chromatin, Mol. Cell, 81, 2944-2959.e10,
doi:10.1016/j.molcel.2021.05.026.
122. Shao, W., Bi, X., Pan, Y., Gao, B., Wu, J., et al. (2022)
Phase separation of RNA-binding protein promotes
polymerase binding and transcription, Nat. Chem.
Biol., 18, 70-80, doi:10.1038/s41589-021-00904-5.
123. Oksuz, O., Henninger, J. E., Warneford-Thomson, R.,
Zheng, M.M., Erb,H., Vancura, A., Overholt, K. J., etal.
(2023) Transcription factors interact with RNA to reg-
ulate genes, Mol. Cell, 83, 2449-2463.e13, doi:10.1016/
j.molcel.2023.06.012.
124. Maharana, S., Wang, J., Papadopoulos, D. K., Rich-
ter,D., Pozniakovsky,A., etal. (2018) RNA buffers the
phase separation behavior of prion-like RNA binding
proteins, Science, 360, 918-921, doi: 10.1126/science.
aar7366.
125. Henninger, J. E., Oksuz, O., Shrinivas, K., Sagi, I., Le-
Roy, G., et al. (2021) RNA-mediated feedback control
of transcriptional condensates, Cell, 184, 207-225.e24,
doi:10.1016/j.cell.2020.11.030.
126. Sharp, P. A., Chakraborty, A. K., Henninger, J. E., and
Young, R. A. (2022) RNA in formation and regula-
tion of transcriptional condensates, RNA, 28, 52-57,
doi:10.1261/rna.078997.121.
127. Pownall, M.E., Miao,L., Vejnar, C.E., M’Saad,O., Sher-
rard,A., etal. (2023) Chromatin expansion microscopy
reveals nanoscale organization of transcription and
chromatin, Science, 381, 92-100, doi: 10.1126/science.
ade5308.
128. Iarovaia, O. V., Minina, E. P., Sheval, E. V., Onich-
tchouk,D., Dokudovskaya,S., Razin, S. V., and Vassetz-
ky, Y. S. (2019) Nucleolus: a central hub for nuclear
functions, Trends Cell Biol., 29, 647-659, doi: 10.1016/
j.tcb.2019.04.003.
129. Liang, Y., Xu, H., Cheng, T., Fu, Y., Huang, H., et al.
(2022) Gene activation guided by nascent RNA-
bound transcription factors, Nat. Commun.
, 13, 7329,
doi:10.1038/s41467-022-35041-7.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in pub-
lished maps and institutional affiliations.