An intimate relationship exists between nuclear architecture and gene activity. Unraveling the fine-scale three-dimensional structure of the genome and its impact on gene regulation is a major goal of current epigenetic research, one with direct implications for understanding the molecular mechanisms underlying human phenotypic variation and disease susceptibility. In this context, the novel revolutionary genome editing technologies and emerging new ways to manipulate genome folding offer new promises for the treatment of human disorders.

Fundamental, yet unanswered, questions in biology are how genome organization and chromosomal folding influence basic cellular processes such as transcription, how they relate to the development of disease, and whether they can be manipulated therapeutically. In mammals, gene regulatory elements are scattered throughout the genome, collectively occupying a significant fraction of the genomic noncoding DNA content. Initially, noncoding DNA (comprising ∼98% of the human genome) was considered to be largely “junk” DNA, lacking function. However, the fast-growing collection of genome-wide datasets describing chromatin features across increasing numbers of cell types has dramatically changed this view. These studies have started to reveal the organizational complexity of mammalian genomes, and it is at present speculated that ∼40% to 80% of the genome shows a biochemical signature that could imply functional relevance.1-3  Transcriptional enhancers represent a critical component of this noncoding regulatory genome as they bestow a unique identity on cells by establishing cell type-specific spatio-temporal gene expression patterns.4,5  In line with their essential roles in transcriptional regulation, numerous recent studies have causally linked aberrant enhancer function to human disorders and phenotypic variation, further demonstrating the important roles played by transcriptional enhancers in human biology.6-19  Through this review article, we aim to provide a concise update on new insights obtained in the last few years concerning the molecular mechanisms by which regulatory elements regulate gene expression, often over large genomic distances, and how disruption of these processes can contribute to the development of human disease. We will also discuss emerging therapeutic strategies aimed at manipulating the function of enhancers for the treatment of human genetic disorders.

Many thousands of potential enhancers have been identified in the human genome,1  of which thousands are active in a given cell type.20-22  Enhancers are often localized at large distances from the genes they regulate, with an estimated median enhancer-target gene distance of 120 kb,23  although extreme cases of >1 Mb have been documented.8,12,24  They can be positioned both intragenic and intergenic, or even in nonrelated genes, and do not necessarily regulate transcription of the nearest gene.25  Enhancers regulate genes over large genomic distances via chromatin looping, bringing distal enhancers and the regulatory protein complexes that bind them in close nuclear proximity to their target genes. Chromatin loop formation has therefore been shown to be a better predictor of enhancer target genes than enhancer-gene linear proximity25  (although it is important to note that chromatin looping does not functionally connect enhancers and promoters per se).

Well-studied examples of such long-range gene regulation in the hematopoietic system are the erythroid globin,26-28 BCL11A,13 Myb,29  and Kit30  gene loci. Gene regulatory chromatin-looping events are thought to be dynamic and actively modulated during differentiation to accommodate for the changes in target gene expression necessary during development and cellular differentiation, although a recent study in Drosophila indicates that enhancer-promoter loops may be remarkably stable during development.31 

Because of the high degree of complexity and specificity required for enhancer gene communication, chromosome conformation needs to be highly organized. A substantial body of evidence suggests that the determinants of promoter enhancer specificity can be very diverse,32-34  ranging from transcription factors (TFs),30,35-38  chromatin modifying proteins,39  or so-called “architectural” proteins (ie, CTCF, Cohesin, and Mediator)40-44  to noncoding RNAs (eg, enhancer RNAs and long noncoding RNAs).45,46  Enhancer-promoter interactions are promoted by the confinement of such interactions to chromosome structural domains called topological-associated domains (TADs), which partition chromosomes into discrete submegabase- to megabase-sized domains.47-49  This observation suggests a “loops within loops” model, where TADs provide a structural environment that prevents enhancer promiscuity.50  Other studies suggest that active genes regulated by similar TF complexes tend to cluster in the nuclear space or even colocalize in specific nuclear foci referred to as transcription factories.50-54  Although the occurrence of (active) gene movement toward relatively static transcription factories is presently still under debate,54,55  it is clear that different scales of genome folding are intimately linked to transcription regulation by allowing proper enhancer-gene contacts, placing enhancers and genome spatial organization at the heart of transcription.

The majority of identified disease-associated genomic mutations and polymorphisms are located in noncoding DNA regions, often colocalizing with potential regulatory sequences.1  For several examples, single nucleotide polymorphisms have been shown to significantly influence long-range chromatin folding (Figure 1).14,15  In addition, cancer cells typically show massive structural and spatial chromosomal rearrangements that potentially displace regulatory elements from their native into an ectopic environment.6-11  As a consequence, novel long-range interactions can be established between normally unrelated enhancer-promoter pairs, leading to improper gene regulation and cellular transformation (Figure 1).7,56-58  An elegant study by the Delwel laboratory recently reported a prime example of how aberrant enhancer “rewiring” can cause disease.7  They investigated acute myeloid leukemia (AML) cells bearing inv(3)/t(3;3) chromosomal rearrangements, which are characterized by the transpositioning of a GATA2 enhancer into the EVI1 stem cell regulator locus. This chromosomal abnormality causes leukemia through the inappropriate long-range activation of EVI1 expression by the ectopic GATA2 enhancer, possibly reinforced by the accompanying reduction of GATA2 expression. Studies by the Yamamoto group using a transgenic mouse model of inv(3) AML further confirmed these observations.58  Another recent study investigated the molecular basis of pediatric medulloblastoma, in which complex chromosomal rearrangements activate the GFI1 and GFI1B oncogenes by placing them under the transcriptional control of unrelated enhancer elements,56  an event the authors referred to as “enhancer hijacking.” These studies highlight the potential pathological impact of regulatory element displacement in human disease, underscoring the value of investigating spatial genomic organization when dissecting the molecular events associated with cancer.

Figure 1

Genomic alterations can affect gene regulation via chromosome conformation. (A) Chromatin folding plays an essential role in transcriptional control by distally located regulatory elements. (B) Disease-associated single nucleotide polymorphisms located in distal regulatory elements can influence long-range chromatin interactions through their detrimental effect on the recruitment of TF complexes (eg, by destroying a TF binding motif), resulting in reduced expression of the target gene. (C) Chromosomal aberrations (eg, translocations) can relocate distal enhancers near a disease-associated gene, leading to the formation of pathological long-range chromatin interactions that ectopically activate the expression of this gene.

Figure 1

Genomic alterations can affect gene regulation via chromosome conformation. (A) Chromatin folding plays an essential role in transcriptional control by distally located regulatory elements. (B) Disease-associated single nucleotide polymorphisms located in distal regulatory elements can influence long-range chromatin interactions through their detrimental effect on the recruitment of TF complexes (eg, by destroying a TF binding motif), resulting in reduced expression of the target gene. (C) Chromosomal aberrations (eg, translocations) can relocate distal enhancers near a disease-associated gene, leading to the formation of pathological long-range chromatin interactions that ectopically activate the expression of this gene.

Close modal

Because of their key role in human disease, targeting these newly established disease-associated long-range interactions has become an attractive therapeutic strategy. Recent advances in genome editing technologies, such as using Zinc-finger nucleases (ZnFs), transcription activator-like effector nucleases (TALENs), or the CRISPR/Cas9 system (see Gaj et al.59  and Gupta and Musunuru60  for comprehensive reviews), make targeted enhancer-modifying strategies feasible and open up exciting new avenues for the therapeutic manipulation of genome topology.

In the abovementioned study by Gröschel and colleagues,7  excision of the oncogenic GATA2 enhancer from the AML genome, using either TALEN- or CRISPR/Cas9-mediated genome editing, induced in vitro growth arrest and differentiation of the edited leukemic cells, demonstrating the promising potential of hijacked oncogenic enhancers as therapeutic targets.

Enhancer activity is often highly cell type specific, and even widely expressed genes seem to possess tissue-specific enhancers driving their expression.61  This cell type-specific nature of enhancers makes them suitable targets for tissue-specific modulation of gene expression. Targeting enhancers (rather than promoters or the gene products themselves) offers the advantage of allowing cell type-specific silencing of gene expression (Figure 2). Such a strategy would even allow reducing the levels of currently undruggable proteins, including TFs.62  This principle was first explored by Bauer et al, using genome-editing technology to delete an erythroid-specific enhancer of the BCL11A TF gene.13  BCL11A is widely recognized as an important therapeutic target for the treatment of β-thalassemias and sickle cell anemia (the β-hemoglobinopathies), 2 common erythroid genetic disorders caused by, respectively, a quantitative or qualitative defect in adult hemoglobin production.63,64  In erythroid cells, BCL11A plays an important role in β-globin gene regulation, and its depletion in adult erythroid cells leads to a strong reactivation of fetal (β-like) γ-globin gene expression, which can efficiently compensate for the low abundant or defective adult hemoglobin.65-67  However, as BCL11A is a widely expressed TF and has been implicated in lymphomagenesis,68,69  targeting the protein itself remains problematic. In erythroid cells, BCL11A expression is controlled by intronic enhancers located 55 to 62 kb downstream of the transcription start site.13  Targeted deletion of these enhancers dramatically reduces BCL11A expression specifically in erythroid cells, showing that targeting enhancers allows for the tissue-specific silencing of broadly expressed genes. Other genome editing-based strategies to manipulate enhancer biology can be envisioned, such as the specific targeting of repressor (domain) fusion proteins to disease-associated enhancers. Mendenhall and colleagues pioneered this approach by fusing transcription activator–like (TAL) effector repeat domains to the LSD1 histone demethylase. Targeting this fusion protein to the genome efficiently reduced enhancer activity, resulting in the downregulation of nearby genes.70 

Figure 2

Enhancer targeting strategies with proven therapeutical potential. (A) Deletion of enhancers can cell type specifically modulate disease-associated gene expression. (B) Artificial tethering of TFs to regulatory elements can restore disease-impaired chromatin looping, leading to reactivation of gene expression. Similar tethering strategies can also be used to artificially reactivate naturally silenced genes for therapeutic use74  or for enhancer silencing by tethering inhibitory proteins to the enhancer.70 

Figure 2

Enhancer targeting strategies with proven therapeutical potential. (A) Deletion of enhancers can cell type specifically modulate disease-associated gene expression. (B) Artificial tethering of TFs to regulatory elements can restore disease-impaired chromatin looping, leading to reactivation of gene expression. Similar tethering strategies can also be used to artificially reactivate naturally silenced genes for therapeutic use74  or for enhancer silencing by tethering inhibitory proteins to the enhancer.70 

Close modal

As stated before, TFs are involved in establishing and stabilizing long-range chromatin interactions. The non-DNA binding adaptor protein LDB1 is required for the development of multiple tissues, including the hematopoietic system.71,72  LDB1 assembles a multiprotein TFC complex in erythroid cells, containing 2 essential DNA-binding TFs GATA1 and TAL1, and LDB1 is required for enhancer-promoter looping at the β-globin and Myb loci.29,36,37,73  A direct demonstration that LDB1 is the critical complex component mediating chromatin looping came from an elegant study by the Blobel laboratory.73  During erythropoiesis, the LDB1 TF complex binds the β-globin gene promoter and upstream locus control region (LCR) to achieve LCR promoter looping and high-level globin gene expression.36,37  In GATA1-deficient erythroid progenitors, LDB1 is only targeted to the LCR (via its interaction with TAL1) but not to the β-globin promoter. In these cells, long-range LCR-promoter interactions are absent, and the β-globin genes are therefore not expressed. Artificial ZnF-mediated tethering of LDB1, or the LDB1 dimerization domain only, to the β-globin promoter in these GATA1-deficient cells was shown to be sufficient for establishing the LCR-promoter loop, resulting in a (partial) activation of β-globin gene expression.73 

This “forced-looping” strategy was then used in an attempt to reactivate fetal γ-globin gene expression in adult red blood cells as a potential new strategy for the treatment of β-hemoglobinopathies.74  Deng et al showed that forced chromatin looping of the β-globin LCR to the developmentally silenced fetal γ-globin gene reactivates its expression in cultured adult erythroid cells. This time, they manipulated local long-range chromatin interactions by targeting the ZnF-LDB1 dimerization domain fusion-protein to the human γ-globin gene promoter. Importantly, γ-globin expression reached levels that would be sufficient to significantly ameliorate the clinical disease course of β-hemoglobinopathy patients.75  This study provides the first proof of principle that long-range chromatin interactions can be artificially controlled—potentially even for therapeutic purposes—providing an invaluable addition to the ever-growing genomics toolbox at our disposal (Figure 2).

Despite the attractiveness of enhancer targeting as a potential broadly applicable therapeutic approach, several important challenges need to be faced before such strategies become clinically applicable. Whereas ZnF-LDB1-based reactivation of γ-globin did not appear to have a major impact on the majority of a selected set of erythroid genes controlled by the endogenous LDB1 complex, the influence of overexpressed engineered fusion proteins on the activity of endogenous regulatory protein complexes and the genes they regulate needs to be rigorously tested. Similarly, the off-target effects of genomic engineering technologies are still under intense investigation, and several studies have highlighted the need for more optimized targeting strategies.76,77  In addition, because of the important role of regulatory elements in organizing spatial genome topology, potential side effects of (even small) genomic alterations on local three-dimensional chromosomal organization and gene regulation will have to be thoroughly investigated. Finally, there is the considerable challenge of efficient and specific delivery of the genome editing constructs to the target cell. With a few exceptions,78,79  this is still only feasible using ex vivo cultured (stem) cells that can be transplanted back into the recipient.80  Major technological developments in this area are required if the current genome editing methods are to be used for systemic therapy in vivo.

Enhancers that regulate gene expression over large distances by chromatin looping processes are critical for proper development and tissue homeostasis. Recent progress has clearly shown that the genetic disruption of enhancer function plays a widespread and important role in human phenotypic variation, disease susceptibility, and even disease etiology. As new technologies that allow the targeted manipulation of regulatory elements develop at an astonishing rate, we expect therapeutic strategies aimed at intervening with disease-associated enhancer-gene communication or at establishing therapeutically beneficial enhancer-gene interactions to become feasible in the future.

The authors thank the members of the Grosveld and Soler laboratories for helpful discussions. The authors apologize to the colleagues whose work could not be cited due to space limitations.

Contribution: All authors contributed to the writing of the paper and approved the final version.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

The current affiliation for R.S. is Centre for Genomic Regulation, Barcelona, Spain.

Correspondence: E. Soler, INSERM UMR967, CEA/DSV/iRCM, Laboratory of Molecular Hematopoiesis, 92265, Fontenay-aux-Roses, France; e-mail: eric.soler@cea.fr.

1
ENCODE Project Consortium
An integrated encyclopedia of DNA elements in the human genome.
Nature
2012
, vol. 
489
 
7414
(pg. 
57
-
74
)
2
Kellis
 
M
Wold
 
B
Snyder
 
MP
, et al. 
Defining functional DNA elements in the human genome.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
17
(pg. 
6131
-
6138
)
3
Stamatoyannopoulos
 
JA
What does our genome encode?
Genome Res
2012
, vol. 
22
 
9
(pg. 
1602
-
1611
)
4
Ong
 
C-T
Corces
 
VG
Enhancer function: new insights into the regulation of tissue-specific gene expression.
Nat Rev Genet
2011
, vol. 
12
 
4
(pg. 
283
-
293
)
5
Bulger
 
M
Groudine
 
M
Functional and mechanistic diversity of distal transcription enhancers.
Cell
2011
, vol. 
144
 
3
(pg. 
327
-
339
)
6
Farh
 
KK-H
Marson
 
A
Zhu
 
J
, et al. 
Genetic and epigenetic fine mapping of causal autoimmune disease variants.
Nature
 
2014 Oct 29. doi: 10.1038/nature13835
7
Gröschel
 
S
Sanders
 
MA
Hoogenboezem
 
R
, et al. 
A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia.
Cell
2014
, vol. 
157
 
2
(pg. 
369
-
381
)
8
Herranz
 
D
Ambesi-Impiombato
 
A
Palomero
 
T
, et al. 
A NOTCH1-driven MYC enhancer promotes T cell development, transformation and acute lymphoblastic leukemia.
Nat Med
2014
, vol. 
20
 
10
(pg. 
1130
-
1137
)
9
Corcoran
 
LM
Cory
 
S
Adams
 
JM
Transposition of the immunoglobulin heavy chain enhancer to the myc oncogene in a murine plasmacytoma.
Cell
1985
, vol. 
40
 
1
(pg. 
71
-
79
)
10
Busslinger
 
M
Klix
 
N
Pfeffer
 
P
Graninger
 
PG
Kozmik
 
Z
Deregulation of PAX-5 by translocation of the Emu enhancer of the IgH locus adjacent to two alternative PAX-5 promoters in a diffuse large-cell lymphoma.
Proc Natl Acad Sci USA
1996
, vol. 
93
 
12
(pg. 
6129
-
6134
)
11
Fahrlander
 
PD
Sümegi
 
J
Yang
 
JQ
Wiener
 
F
Marcu
 
KB
Klein
 
G
Activation of the c-myc oncogene by the immunoglobulin heavy-chain gene enhancer after multiple switch region-mediated chromosome rearrangements in a murine plasmacytoma.
Proc Natl Acad Sci USA
1985
, vol. 
82
 
11
(pg. 
3746
-
3750
)
12
Lettice
 
LA
Heaney
 
SJ
Purdie
 
LA
, et al. 
A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly.
Hum Mol Genet
2003
, vol. 
12
 
14
(pg. 
1725
-
1735
)
13
Bauer
 
DE
Kamran
 
SC
Lessard
 
S
, et al. 
An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level.
Science
2013
, vol. 
342
 
6155
(pg. 
253
-
257
)
14
Stadhouders
 
R
Aktuna
 
S
Thongjuea
 
S
, et al. 
HBS1L-MYB intergenic variants modulate fetal hemoglobin via long-range MYB enhancers.
J Clin Invest
2014
, vol. 
124
 
4
(pg. 
1699
-
1710
)
15
Visser
 
M
Kayser
 
M
Palstra
 
R-J
HERC2 rs12913832 modulates human pigmentation by attenuating chromatin-loop formation between a long-range enhancer and the OCA2 promoter.
Genome Res
2012
, vol. 
22
 
3
(pg. 
446
-
455
)
16
Hindorff
 
LA
Sethupathy
 
P
Junkins
 
HA
, et al. 
Potential etiologic and functional implications of genome-wide association loci for human diseases and traits.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
23
(pg. 
9362
-
9367
)
17
Weinhold
 
N
Jacobsen
 
A
Schultz
 
N
Sander
 
C
Lee
 
W
Genome-wide analysis of noncoding regulatory mutations in cancer.
Nat Genet
2014
, vol. 
46
 
11
(pg. 
1160
-
1165
)
18
Kioussis
 
D
Vanin
 
E
deLange
 
T
Flavell
 
RA
Grosveld
 
FG
Beta-globin gene inactivation by DNA translocation in gamma beta-thalassaemia.
Nature
1983
, vol. 
306
 
5944
(pg. 
662
-
666
)
19
Soudon
 
J
Bernard
 
O
Mathieu-Mahul
 
D
Larsen
 
CJ
c-myc gene expression in a leukemic T-cell line bearing a t(8;14) (q24;q11) translocation.
Leukemia
1991
, vol. 
5
 
1
(pg. 
60
-
65
)
20
Andersson
 
R
Gebhard
 
C
Miguel-Escalada
 
I
, et al. 
FANTOM Consortium
An atlas of active enhancers across human cell types and tissues.
Nature
2014
, vol. 
507
 
7493
(pg. 
455
-
461
)
21
Heintzman
 
ND
Hon
 
GC
Hawkins
 
RD
, et al. 
Histone modifications at human enhancers reflect global cell-type-specific gene expression.
Nature
2009
, vol. 
459
 
7243
(pg. 
108
-
112
)
22
Ho
 
JWK
Jung
 
YL
Liu
 
T
, et al. 
Comparative analysis of metazoan chromatin organization.
Nature
2014
, vol. 
512
 
7515
(pg. 
449
-
452
)
23
Sanyal
 
A
Lajoie
 
BR
Jain
 
G
Dekker
 
J
The long-range interaction landscape of gene promoters.
Nature
2012
, vol. 
489
 
7414
(pg. 
109
-
113
)
24
Velagaleti
 
GVN
Bien-Willner
 
GA
Northup
 
JK
, et al. 
Position effects due to chromosome breakpoints that map approximately 900 Kb upstream and approximately 1.3 Mb downstream of SOX9 in two patients with campomelic dysplasia.
Am J Hum Genet
2005
, vol. 
76
 
4
(pg. 
652
-
662
)
25
Smemo
 
S
Tena
 
JJ
Kim
 
K-H
, et al. 
Obesity-associated variants within FTO form long-range functional connections with IRX3.
Nature
2014
, vol. 
507
 
7492
(pg. 
371
-
375
)
26
Tolhuis
 
B
Palstra
 
RJ
Splinter
 
E
Grosveld
 
F
de Laat
 
W
Looping and interaction between hypersensitive sites in the active beta-globin locus.
Mol Cell
2002
, vol. 
10
 
6
(pg. 
1453
-
1465
)
27
Hughes
 
JR
Lower
 
KM
Dunham
 
I
, et al. 
High-resolution analysis of cis-acting regulatory networks at the α-globin locus.
Philos Trans R Soc Lond B Biol Sci
2013
, vol. 
368
 
1620
pg. 
20120361
 
28
Vernimmen
 
D
Uncovering enhancer functions using the α-globin locus.
PLoS Genet
2014
, vol. 
10
 
10
pg. 
e1004668
 
29
Stadhouders
 
R
Thongjuea
 
S
Andrieu-Soler
 
C
, et al. 
Dynamic long-range chromatin interactions control Myb proto-oncogene transcription during erythroid development.
EMBO J
2012
, vol. 
31
 
4
(pg. 
986
-
999
)
30
Jing
 
H
Vakoc
 
CR
Ying
 
L
, et al. 
Exchange of GATA factors mediates transitions in looped chromatin organization at a developmentally regulated gene locus.
Mol Cell
2008
, vol. 
29
 
2
(pg. 
232
-
242
)
31
Ghavi-Helm
 
Y
Klein
 
FA
Pakozdi
 
T
, et al. 
Enhancer loops appear stable during development and are associated with paused polymerase.
Nature
2014
, vol. 
512
 
7512
(pg. 
96
-
100
)
32
Maksimenko
 
O
Georgiev
 
P
Mechanisms and proteins involved in long-distance interactions.
Front Genet
2014
, vol. 
5
 pg. 
28
 
33
Gorkin
 
DU
Leung
 
D
Ren
 
B
The 3D genome in transcriptional regulation and pluripotency.
Cell Stem Cell
2014
, vol. 
14
 
6
(pg. 
762
-
775
)
34
van Arensbergen
 
J
van Steensel
 
B
Bussemaker
 
HJ
In search of the determinants of enhancer-promoter interaction specificity.
Trends Cell Biol
2014
, vol. 
24
 
11
(pg. 
695
-
702
)
35
Drissen
 
R
Palstra
 
R-J
Gillemans
 
N
, et al. 
The active spatial organization of the beta-globin locus requires the transcription factor EKLF.
Genes Dev
2004
, vol. 
18
 
20
(pg. 
2485
-
2490
)
36
Song
 
S-H
Hou
 
C
Dean
 
A
A positive role for NLI/Ldb1 in long-range beta-globin locus control region function.
Mol Cell
2007
, vol. 
28
 
5
(pg. 
810
-
822
)
37
Krivega
 
I
Dale
 
RK
Dean
 
A
Role of LDB1 in the transition from chromatin looping to transcription activation.
Genes Dev
2014
, vol. 
28
 
12
(pg. 
1278
-
1290
)
38
Stadhouders
 
R
de Bruijn
 
MJW
Rother
 
MB
, et al. 
Pre-B cell receptor signaling induces immunoglobulin κ locus accessibility by functional redistribution of enhancer-mediated chromatin interactions.
PLoS Biol
2014
, vol. 
12
 
2
pg. 
e1001791
 
39
Kim
 
S-I
Bultman
 
SJ
Kiefer
 
CM
Dean
 
A
Bresnick
 
EH
BRG1 requirement for long-range interaction of a locus control region with a downstream promoter.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
7
(pg. 
2259
-
2264
)
40
Splinter
 
E
Heath
 
H
Kooren
 
J
, et al. 
CTCF mediates long-range chromatin looping and local histone modification in the beta-globin locus.
Genes Dev
2006
, vol. 
20
 
17
(pg. 
2349
-
2354
)
41
Ribeiro de Almeida
 
C
Stadhouders
 
R
Thongjuea
 
S
Soler
 
E
Hendriks
 
RW
DNA-binding factor CTCF and long-range gene interactions in V(D)J recombination and oncogene activation.
Blood
2012
, vol. 
119
 
26
(pg. 
6209
-
6218
)
42
Ong
 
C-T
Corces
 
VG
CTCF: an architectural protein bridging genome topology and function.
Nat Rev Genet
2014
, vol. 
15
 
4
(pg. 
234
-
246
)
43
Kagey
 
MH
Newman
 
JJ
Bilodeau
 
S
, et al. 
Mediator and cohesin connect gene expression and chromatin architecture.
Nature
2010
, vol. 
467
 
7314
(pg. 
430
-
435
)
44
Zuin
 
J
Dixon
 
JR
van der Reijden
 
MIJA
, et al. 
Cohesin and CTCF differentially affect chromatin architecture and gene expression in human cells.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
3
(pg. 
996
-
1001
)
45
Quinodoz
 
S
Guttman
 
M
Long noncoding RNAs: an emerging link between gene regulation and nuclear organization.
Trends Cell Biol
2014
, vol. 
24
 
11
(pg. 
651
-
663
)
46
Ørom
 
UA
Shiekhattar
 
R
Long noncoding RNAs usher in a new era in the biology of enhancers.
Cell
2013
, vol. 
154
 
6
(pg. 
1190
-
1193
)
47
Dixon
 
JR
Selvaraj
 
S
Yue
 
F
, et al. 
Topological domains in mammalian genomes identified by analysis of chromatin interactions.
Nature
2012
, vol. 
485
 
7398
(pg. 
376
-
380
)
48
Nora
 
EP
Lajoie
 
BR
Schulz
 
EG
, et al. 
Spatial partitioning of the regulatory landscape of the X-inactivation centre.
Nature
2012
, vol. 
485
 
7398
(pg. 
381
-
385
)
49
Jin
 
F
Li
 
Y
Dixon
 
JR
, et al. 
A high-resolution map of the three-dimensional chromatin interactome in human cells.
Nature
2013
, vol. 
503
 
7475
(pg. 
290
-
294
)
50
Soler
 
E
Grosveld
 
F
 
Transcription regulation in stem cells. In Durand C, Charbord P, eds. Stem Cell Biology and Regenerative Medicine. Aalborg, Denmark: River Publishers; 2014:29-58
51
de Wit
 
E
Bouwman
 
BA
Zhu
 
Y
, et al. 
The pluripotent genome in three dimensions is shaped around pluripotency factors.
Nature
2013
, vol. 
501
 
7466
(pg. 
227
-
231
)
52
Schoenfelder
 
S
Sexton
 
T
Chakalova
 
L
, et al. 
Preferential associations between co-regulated genes reveal a transcriptional interactome in erythroid cells.
Nat Genet
2010
, vol. 
42
 
1
(pg. 
53
-
61
)
53
Ghamari
 
A
van de Corput
 
MPC
Thongjuea
 
S
, et al. 
In vivo live imaging of RNA polymerase II transcription factories in primary cells.
Genes Dev
2013
, vol. 
27
 
7
(pg. 
767
-
777
)
54
Papantonis
 
A
Cook
 
PR
Transcription factories: genome organization and gene regulation.
Chem Rev
2013
, vol. 
113
 
11
(pg. 
8683
-
8705
)
55
Cisse
 
II
Izeddin
 
I
Causse
 
SZ
, et al. 
Real-time dynamics of RNA polymerase II clustering in live human cells.
Science
2013
, vol. 
341
 
6146
(pg. 
664
-
667
)
56
Northcott
 
PA
Lee
 
C
Zichner
 
T
, et al. 
Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma.
Nature
2014
, vol. 
511
 
7510
(pg. 
428
-
434
)
57
Kovalchuk
 
AL
Ansarah-Sobrinho
 
C
Hakim
 
O
, et al. 
Mouse model of endemic Burkitt translocations reveals the long-range boundaries of Ig-mediated oncogene deregulation.
Proc Natl Acad Sci USA
2012
, vol. 
109
 
27
(pg. 
10972
-
10977
)
58
Yamazaki
 
H
Suzuki
 
M
Otsuki
 
A
, et al. 
A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3)(q21;q26) by activating EVI1 expression.
Cancer Cell
2014
, vol. 
25
 
4
(pg. 
415
-
427
)
59
Gaj
 
T
Gersbach
 
CA
Barbas
 
CF
ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering.
Trends Biotechnol
2013
, vol. 
31
 
7
(pg. 
397
-
405
)
60
Gupta
 
RM
Musunuru
 
K
Expanding the genetic editing tool kit: ZFNs, TALENs, and CRISPR-Cas9.
J Clin Invest
2014
, vol. 
124
 
10
(pg. 
4154
-
4161
)
61
Kieffer-Kwon
 
K-R
Tang
 
Z
Mathe
 
E
, et al. 
Interactome maps of mouse gene regulatory domains reveal basic principles of transcriptional regulation.
Cell
2013
, vol. 
155
 
7
(pg. 
1507
-
1520
)
62
Koehler
 
AN
A complex task? Direct modulation of transcription factors with small molecules.
Curr Opin Chem Biol
2010
, vol. 
14
 
3
(pg. 
331
-
340
)
63
Weatherall
 
DJ
The inherited diseases of hemoglobin are an emerging global health burden.
Blood
2010
, vol. 
115
 
22
(pg. 
4331
-
4336
)
64
Bauer
 
DE
Kamran
 
SC
Orkin
 
SH
Reawakening fetal hemoglobin: prospects for new therapies for the β-globin disorders.
Blood
2012
, vol. 
120
 
15
(pg. 
2945
-
2953
)
65
Sankaran
 
VG
Menne
 
TF
Xu
 
J
, et al. 
Human fetal hemoglobin expression is regulated by the developmental stage-specific repressor BCL11A.
Science
2008
, vol. 
322
 
5909
(pg. 
1839
-
1842
)
66
Sankaran
 
VG
Xu
 
J
Ragoczy
 
T
, et al. 
Developmental and species-divergent globin switching are driven by BCL11A.
Nature
2009
, vol. 
460
 
7259
(pg. 
1093
-
1097
)
67
Chen
 
Z
Luo
 
HY
Steinberg
 
MH
Chui
 
DHK
BCL11A represses HBG transcription in K562 cells.
Blood Cells Mol Dis
2009
, vol. 
42
 
2
(pg. 
144
-
149
)
68
Kuo
 
T-Y
Hsueh
 
Y-P
Expression of zinc finger transcription factor Bcl11A/Evi9/CTIP1 in rat brain.
J Neurosci Res
2007
, vol. 
85
 
8
(pg. 
1628
-
1636
)
69
Satterwhite
 
E
Sonoki
 
T
Willis
 
TG
, et al. 
The BCL11 gene family: involvement of BCL11A in lymphoid malignancies.
Blood
2001
, vol. 
98
 
12
(pg. 
3413
-
3420
)
70
Mendenhall
 
EM
Williamson
 
KE
Reyon
 
D
, et al. 
Locus-specific editing of histone modifications at endogenous enhancers.
Nat Biotechnol
2013
, vol. 
31
 
12
(pg. 
1133
-
1136
)
71
Mukhopadhyay
 
M
Teufel
 
A
Yamashita
 
T
, et al. 
Functional ablation of the mouse Ldb1 gene results in severe patterning defects during gastrulation.
Development
2003
, vol. 
130
 
3
(pg. 
495
-
505
)
72
Love
 
PE
Warzecha
 
C
Li
 
L
Ldb1 complexes: the new master regulators of erythroid gene transcription.
Trends Genet
2014
, vol. 
30
 
1
(pg. 
1
-
9
)
73
Deng
 
W
Lee
 
J
Wang
 
H
, et al. 
Controlling long-range genomic interactions at a native locus by targeted tethering of a looping factor.
Cell
2012
, vol. 
149
 
6
(pg. 
1233
-
1244
)
74
Deng
 
W
Rupon
 
JW
Krivega
 
I
, et al. 
Reactivation of developmentally silenced globin genes by forced chromatin looping.
Cell
2014
, vol. 
158
 
4
(pg. 
849
-
860
)
75
Platt
 
OS
Hydroxyurea for the treatment of sickle cell anemia.
N Engl J Med
2008
, vol. 
358
 
13
(pg. 
1362
-
1369
)
76
Fu
 
Y
Foden
 
JA
Khayter
 
C
, et al. 
High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells.
Nat Biotechnol
2013
, vol. 
31
 
9
(pg. 
822
-
826
)
77
Cho
 
SW
Kim
 
S
Kim
 
Y
, et al. 
Analysis of off-target effects of CRISPR/Cas-derived RNA-guided endonucleases and nickases.
Genome Res
2014
, vol. 
24
 
1
(pg. 
132
-
141
)
78
Komáromy
 
AM
Alexander
 
JJ
Rowlan
 
JS
, et al. 
Gene therapy rescues cone function in congenital achromatopsia.
Hum Mol Genet
2010
, vol. 
19
 
13
(pg. 
2581
-
2593
)
79
Yin
 
H
Xue
 
W
Chen
 
S
, et al. 
Genome editing with Cas9 in adult mice corrects a disease mutation and phenotype.
Nat Biotechnol
2014
, vol. 
32
 
6
(pg. 
551
-
553
)
80
Mandal
 
PK
Ferreira
 
LMR
Collins
 
R
, et al. 
Efficient ablation of genes in human hematopoietic stem and effector cells using CRISPR/Cas9.
Cell Stem Cell
2014
, vol. 
15
 
5
(pg. 
643
-
652
)
Sign in via your Institution