Tải bản đầy đủ (.pdf) (8 trang)

Tài liệu Báo cáo khoa học: Transient DNA ⁄ RNA-protein interactions docx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (512.74 KB, 8 trang )

REVIEW ARTICLE
Transient DNA

RNA-protein interactions
Francisco J. Blanco
1,2
and Guillermo Montoya
3
1 Structural Biology Unit, CIC bioGUNE, Derio, Spain
2 IKERBASQUE, Basque Foundation for Science, Bilbao, Spain
3 Structural Biology and Biocomputing Programme, Spanish National Cancer Research Centre (CNIO), Madrid, Spain
Introduction
Protein–nucleic acids interactions
and structural genomics
In their celebrated reports on the double helix, Watson
and Crick showed the strong structure–function rela-
tionship in the DNA molecule. This relationship is
more intricate in the case of RNA, because of its lar-
ger structural and functional diversity, and even more
so when we consider proteinÆnucleic acid complexes,
given the much larger diversity found in proteins.
Rapid genome-sequencing methods, large-scale gene
expression analysis and high-throughput structural
genomics projects have greatly augmented the number
of known biomacromolecular structures. Currently,
about 72 000 structures are deposited in the Protein
Data Bank, but only 3% are nucleic acids and about
4% are proteinÆnucleic acid complexes. It is difficult to
know whether these figures mirror the prevalence of
proteins and their complexes in the cell, or whether
they arise from the greater difficulties in the identifica-


tion and experimental determination of proteinÆnucleic
acid complexes. Structural genomics initiatives target
the low-hanging fruits, small globular proteins that
can be easily expressed as soluble material in heterolo-
gous systems. An analogous endeavor for RNA mole-
cules has not yet been initiated, very probably because
of the experimental difficulties [1]. Indeed, preparing
large amounts of homogeneous RNA for crystalliza-
tion is not trivial. In addition, RNA samples need
careful manipulation, are notably difficult to crystal-
lize, give poor contrast in cryo-electron microscopy,
and suffer from severe signal overlap in NMR spectra.
Protein–protein interactions may also be transient and
difficult to capture experimentally, but they have been
intensively targeted on a large scale by means of com-
plementary methods, such as yeast two-hybrid and
Keywords
DNA; dynamics; endonuclease; interaction;
nucleosome; protein; ribosome; RNA;
structure; transient
Correspondence
F. J. Blanco, Structural Biology Unit, CIC
bioGUNE, Parque Tecnolo
´
gico de Vizcaya,
48160 Derio, Spain
Fax: +34 9465 72502
Tel: +34 9465 72521
E-mail:
(Received 23 November 2010, revised 17

February 2011, accepted 11 March 2011)
doi:10.1111/j.1742-4658.2011.08095.x
The great pace of biomolecular structure determination has provided a
plethora of protein structures, but not as many structures of nucleic acids
or of their complexes with proteins. The recognition of DNA and RNA
molecules by proteins may produce large and relatively stable assemblies
(such as the ribosome) or transient complexes (such as DNA clamps sliding
through the DNA). These transient interactions are most difficult to char-
acterize, but even in ‘stable’ complexes captured in crystal structures, the
dynamics of the whole or part of the assembly pose great technical difficul-
ties in understanding their function. The development and refinement of
powerful experimental and computational tools have made it possible to
learn a great deal about the relevance of these fleeting events for numerous
biological processes. We discuss here the most recent findings and the chal-
lenges that lie ahead in the quest for a better understanding of protein–
nucleic acid interactions.
Abbreviations
PCNA, proliferating cell nuclear antigen; RCC1, regulator of chromosome condensation 1.
FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS 1643
tandem affinity purification followed by MS. Still, in a
systematic exploration of protein complexes in the
yeast interactome by tandem affinity purification fol-
lowed by MS [2], the protein proliferating cell nuclear
antigen (PCNA) was not detected in any of the 589
purified complexes, despite being a very promiscuous
protein and an essential component of the replication
machinery [3,4].
Replication and transcription
regulation – proteins in search of their
sites on the nucleic acids

PCNA belongs to the group of DNA sliding clamps,
which are multimeric toroidal-shaped structures that
encircle the DNA duplex and act as platforms for rep-
licative polymerases and other proteins. These proces-
sivity factors enable the polymerases to add thousands
of bases per second without detaching from the geno-
mic template. The crystal structure of the homotrimer-
ic yeast PCNA bound to a DNA duplex was recently
solved [5], but only a few of the bases were seen, prob-
ably because of the transient nature of the interaction
in solution (Fig. 1). The recent assignment of the
NMR spectrum of the PCNA ring [6,7] provides the
basis for solution studies of its interactions with DNA
and other proteins [8,9]. The yeast helicase minichro-
mosome maintenance complex forms part of the prere-
plicative complex, and has been recently reconstituted
and loaded onto dsDNA [10]. Electron microscopy
shows a barrel of two head-to-head hexamers that
encircles a stretch of DNA of approximately 68 bp
and passively slides along the DNA duplex.
Sliding on the DNA may be common to proteins
other than DNA clamps before they bind to their spe-
cific target sequences. Both one-dimensional diffusion
of proteins on the DNA (sliding) and direct transfer
between distinct binding sites (translocation) would
accelerate the search process relative to the diffusion-
controlled association–dissociation mechanism (Fig. 2)
in the presence of a huge background of nonspecific
DNA, as occurs in the nucleus of the cell, with an esti-
mated DNA concentration of 100 gÆL

)1
[11]. Sliding
and translocation events involve transient interactions
that are difficult to observe and even more difficult to
quantify. Crystal structures sometimes provide hints
about these events, e.g. by the lack of electron density
of DNA or protein regions, or by the observation of
different conformations of amino acid side chains asso-
ciated with the nucleic acid. Analysis in solution by
NMR is a more powerful approach to characterize
these systems [12], allowing the study of the kinetics of
translocation [13,14], as well as the structures of tran-
sient, nonspecific complexes. For instance, the struc-
ture of the Lac repressor bound to a nonspecific
(low-affinity) DNA sequence suggests that binding is
primarily driven by electrostatics, as most of the pro-
tein–DNA interactions do not involve the bases, but
the phosphates and sugars of the DNA backbone [15].
Most of these interactions are preserved in the com-
plex with the specific sequence, but, in addition,
numerous interactions with the bases take place. When
the overall structures of the two complexes are com-
pared, neither the DNA nor the protein undergo large
A
B
C
k
off,A
k
on,A

k
on,B
k
off,B
k
on,C
k
off,C
k
dt,CB
k
dt,BC
Fig. 2. Proteins find their binding sites in an ocean of DNA
sequences. Scheme of a nucleic acid-binding protein (gray ellipses)
involved in four dynamic equilibria (double arrows) and one-dimen-
sional diffusion on the DNA double helix (single arrows). Binding to
sites A, B or C occurs with different affinities (different kinetic k
on
and k
off
rates). Exchange between sites A and B (located close
together in the DNA sequence) can occur through association–dis-
sociation–reassociation or via diffusion on the DNA. Exchange
between sites B and C (located far away in the DNA sequence but
close enough in space to collide) can occur through association–dis-
sociation–reassociation or via direct transfer (k
dt
). For the sake of
clarity, some of the species participating in the individual equilibria
are omitted.

Fig. 1. Structure of a PCNAÆDNA complex. Two views of the
homotrimeric yeast PCNA ring bound to a short DNA duplex, as
deposited in the Protein Data Bank (entry 3K4X). The three poly-
peptide chains are shown as ribbons of different colors, and the
DNA as an orange rod (backbone) and blue–green sticks (bases).
The figure was prepared with
PYMOL ().
Transient protein–DNA ⁄ RNA interactions F. J. Blanco and G. Montoya
1644 FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS
conformational changes, but a protein segment that is
disordered in solution becomes less flexible in the com-
plex with the low-affinity DNA, and structured into an
a-helix in the complex with the specific DNA. There-
fore, the large local conformational landscape that the
protein populates in solution is reduced upon DNA
binding, and much more so when it specifically binds
to its high-affinity sequence. The protein conforma-
tional landscape can be narrowed by small-molecule
allosteric effectors favoring efficient DNA binding, as
found for the transcriptional activator CAP [16], and
the landscape of free DNA includes transient confor-
mations whose relevance still needs to be investigated
[17]. Many transcription factors bind to thousands of
places in the genome, not necessarily located in proxi-
mal promoter regions, and dissociate very fast in vivo,
which may be relevant for long-range and combinato-
rial regulation of transcription [18].
Electrostatics plays a driving role in transient pro-
teinÆnucleic acid interactions, as well as in selecting
and stabilizing the specific ones. Indeed, it may be the

most important factor in the indirect readout as
opposed to the direct readout. These terms differenti-
ate between the recognition mechanisms based on
details of the DNA structure facilitating protein bind-
ing (indirect) and specific amino acid–base contacts
(direct). A recent examination of proteinÆnucleic acid
structures has shown how minor groove narrowing
enhances the negative electrostatic potential of DNA
and forms an arginine-binding site that is widely used
in protein–nucleic acid recognition [19]. Minor groove
width is primarily DNA sequence-dependent (A-tracts
tend to narrow the groove, whereas GC pairs tend to
widen it), although the geometry observed in a given
complex is probably the result of both intrinsic and
protein-induced conformation effects.
Interactions lost in translation
The 2009 Nobel prize in chemistry for studies of the
structure and function of the ribosome rewarded a
long-term effort by several laboratories. The ribosome
was probably the first biomacromolecular machine to
appear in the early stages of life, and performs its
function in essentially the same way in the three king-
doms. The ribosome translates the three-base codons
of mRNAs into the amino acid sequence of the pro-
teins encoded in the corresponding gene (Fig. 3).
Because of its size (about 2.5 MDa) and lack of sym-
metry, it took a long period of sample preparation
refinement and the use of modern diffraction instru-
mentation and methodology to obtain the high-resolu-
tion structure of the 70S ribosome [20]. In the process,

a wealth of information has been obtained about the
mechanism by which the ribosome attains its high level
of accuracy in translation, its catalytic triad (rRNA,
ribosomal protein, and the peptidyl-tRNA substrate),
and the mode of action of many antibiotics, enabling
the design of novel ones (for a brief review, see
the Nobel Foundation Scientific Background published
50S proteins
50S rRNA
tRNA, E-site
tRNA, P-site
tRNA, A-site
mRNA
30S proteins
30S rRNA
Fig. 3. Structure of the 70S ribosome. The structure of the ribosome of Thermus thermophilus (Protein Data Bank entries 1GIX and 1GIY) is
shown, with the rRNA molecules represented by thin coils, the tRNAs by spheres, the mRNA by a thick coil, and the proteins by ribbons. In
the two views shown, the 50S subunit is at the top and the 30S subunit is at the bottom. The figure was obtained from Proteopedia [55].
F. J. Blanco and G. Montoya Transient protein–DNA ⁄ RNA interactions
FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS 1645
by the Royal Swedish Academy of Sciences at http://
nobelprize.org/nobel_prizes/chemistry/laureates/2009/
cheadv09.pdf).
However, translation is a dynamic process, the ribo-
some is a highly dynamic machine, and the crystal
structures can provide only snapshots of intermediates
along the process. It will be very difficult to obtain
crystal structures of all representative states of the
ribosome in action, but a low-resolution picture has
emerged from time-resolved electron microscopy of the

Escherichia coli ribosome. By unbiased hierarchical
classification of 2 000 000 images, 50 structures of the
ribosomal substates during translocation were refined,
and the trajectories of the two tRNAs as they move
through the ribosome were visualized with a resolution
in the 10–20-A
˚
range [21]. Translocation is the final
step in polypeptide chain elongation, and involves the
concerted movement of the tRNAs, the mRNA, and
the 30S subunit relative to the 50S one. The authors
used a molecular system in which retrotranslocation
was the actual movement observed. After addition of
tRNA
fMet
to ribosomes loaded with fMetVal-tRNA
Val
,
retrotranslocation occurred on a scale of several min-
utes, and the samples at different time points were
extracted and frozen for cryo-electron microscopy. It
was found that, at physiological temperatures, no dis-
tinct 30S subunit could be outlined, as it existed in
dynamic equilibrium with a large number of conforma-
tional substates. The emerging picture of the ribosome
during translocation is that of a machine that couples
spontaneous, thermally driven conformational changes
to directed movement.
Transient interactions also occur in tRNA loading.
Whereas each Xxx-tRNA

Xxx
is loaded with the Xxx
amino acid corresponding to its anticodon by a specific
synthase, most bacteria and all archaeons lack glutami-
nyl-tRNA
Gln
synthase. They produce Gln-tRNA
Gln
in
a two-step pathway: glutamylation of the tRNA
Gln
(by
the same low-specificity enzyme that glutamylates the
tRNA
Glu
), and amidation by the corresponding amido-
transferase. The crystal structure of the ‘glutamine
transamidosome’ of Thermotoga maritima [22] shows
that the anticodon-binding domains of the synthase
recognizes the common features of tRNA
Gln
and
tRNA
Glu
(the second and third bases), whereas the so-
called tail domain of the amidotransferase recognizes
the outer corner of the tRNA
Gln
(specifically for the
tRNA

Gln
). The two enzymes bound to the tRNA
Gln
assume alternative conformations for the two consecu-
tive reactions. The catalytic centers of the two enzymes
compete for the acceptor form of tRNA
Gln
, and there-
fore cannot adopt their productive forms simulta-
neously. Hinge polypeptide regions between the
catalytic and anticodon-binding domains of the syn-
thase, and between the catalytic and tail domains of
the amidotransferase, allow both enzymes to adopt the
productive or the nonproductive forms cooperating in
Gln-tRNA
Gln
synthesis, with a low probability of
releasing the intermediate Glu-tRNA
Gln
species. This
‘alternative conformation’ mechanism may be more
common than expected in consecutive enzymatic
reactions.
The transient positioning of the
nucleosome along the genome
The compaction of DNA molecules inside the cells
occurs by supercoiling and binding to specific proteins.
A supercoiled DNA duplex can form a toroid (the
nucleosome, as in eukaryotes) or a plectoneme (inter-
wound, as in bacteria). This distinction is relevant not

only for DNA packing, but also for transcription, rep-
lication, and repair. The reason for this distinction is
not only how accessible the DNA is, but also the twist-
ing degree (overtwisted in the nucleosome and under-
twisted in the plectoneme). However, it has been
argued that a common topology for bacterial and
eukaryotic DNA-based processes might exist, as the
ejection of a histone octamer would convert the nucle-
osome into a plectoneme [23].
The crystal structure of the nucleosome core particle
[24] shows a compact assembly of 147 bp wrapped
around a disk formed by an octamer of histone proteins
(two copies of each one of the four core histones). How-
ever, this picture is deceptively static, because, in the
chromatin, the nucleosome rotational and translational
positioning is not fixed. Nucleosome rotational posi-
tioning (or register) defines the orientation of the DNA
helix on the histone surface, and a 10-bp periodicity is
observed, reflecting a preference for sequences that face
inwards or outwards with respect to the histones and
optimize DNA bending. Analysis of the minor groove
width along the double helix in 35 high-resolution crys-
tal structures of nucleosomes identified a pattern of 14
minima corresponding to regions where the DNA bends
and has close contacts with histone arginine side chains
[19]. The analysis of DNA sequences bound in vivo by
yeast nucleosomes reveals a periodicity for A-tracts
three bases long, with an average of 10 A-tracts per
nucleosomal DNA. Thus, even though long A-tracts
tend to be excluded from the nucleosome [25], A-tracts

exist and facilitate the bending of the DNA around the
histone core.
Translational positioning [25] is strongly influenced
by the spacing between nucleosomes, but this spacing
is variable, with linker DNA regions in the range of
Transient protein–DNA ⁄ RNA interactions F. J. Blanco and G. Montoya
1646 FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS
10–90 bp, and a given nucleosome can invade its
neighbor’s territory [26]. Recent reports [27,28] indicate
that sequence-dependent histone–DNA interactions
have a predominant influence on the measured nucleo-
some occupancy (average number of histone octamer
levels on a given DNA region in a population of cells)
but not on nucleosome positioning (the extent to
which each of the octamers of the population found in
that DNA region deviates from its consensus location)
[29–31]. Thus, the nucleosomal pattern in DNA coding
regions observed in vivo is not determined by DNA
sequence preferences for octamer binding, but primar-
ily arises by statistical positioning from a barrier near
the promoter, and this barrier involves an unknown
aspect of transcriptional initiation by RNA polymer-
ase II [28]. Nucleosomal assembly in vitro, however, is
sequence-dependent.
Obtaining well-diffracting crystals of nucleosomes
was a difficult task, and was strongly dependent on the
DNA sequence. All structures corresponded to nucleo-
somes assembled from purified histones and human
a-satellite DNA sequences, until very recently, when
two new crystal structures of nucleosomes containing

the strongest known histone octamer-binding sequence
have been reported [32,33]. This sequence is the
Widom 601 DNA, the de facto standard for in vitro
nucleosome reconstitution in chromatin biology
research because of its tight binding, but it is a syn-
thetic repetitive sequence that may not be the best rep-
resentative of real genomic sequences assembled into
nucleosomes. The two structures are very similar to the
former ones, but with increased DNA twisting and a
145-bp core particle instead of the canonical 147-bp
one. The increased twist occurs at two superhelical
regions, which are the same regions where some of the
histone–DNA contacts differ from those in the a-satel-
lite nucleosomes. Therefore, the structure of the nucleo-
some can adapt to small variations in DNA length.
One of these two structures also contains the protein
regulator of chromosome condensation 1 (RCC1, also
known as RanGEF or Ran guanine exchange factor),
with implications for nuclear transport and mitosis.
This structure is the first to show how a nonhistone
protein recognizes and binds to the nucleosome
(Fig. 4). It was found that arginines of the switchback
loop of RCC1 interact with an acidic patch on the
histone H2A–H2B dimer, whereas the DNA-binding
loop interacts with phosphates of the nucleosomal
DNA. These results are consistent with RCC1 being a
non-DNA-sequence specific chromatin factor. Interest-
ingly, the acidic patch on the nucleosome is the same as
that occupied by the histone H4 tail of a neighbor
nucleosome in the crystal lattice of the nucleosome [34].

In prokaryotes, the DNA is condensed with polyam-
ines and proteins. In enterobacteria, the histone-like
nucleoid structuring proteins perform this role and reg-
ulate gene expression in response to environmental
changes. The crystals of the oligomerization domain of
histone-like nucleoid structuring proteins reveal an
assembly of symmetry-related dimers into a superhelix,
establishing a mechanism for the self-association [35].
Although there is no structure for the DNA-binding
domain, the superhelical assembly suggests the forma-
tion of a complex with plectonemic DNA.
Engineering and design of
protein-nucleic acid interactions –
lessons from endonucleases
Restriction endonucleases are the DNA protein bind-
ers with the widest application in biotechnology and
biomedicine [36,37], and large efforts are being
invested in the identification of new nucleases or the
modification of extant ones to give novel or improved
DNA sequence specificities [38,39]. Recent successful
examples of redesigned protein–DNA interfaces illus-
trate our increased ability to achieve these goals by the
manipulation of the direct readout interactions [40]. By
use of the crystal structures of the complexes, these
methods aim at optimizing the amino acids for affinity
Fig. 4. Structure of the nucleosome with bound RCC1 proteins.
Ribbon diagram of the structure of the RCC1Ænucleosome core par-
ticle complex assembled from Drosophila melanogaster RCC1,
Xenopus laevis histones, and the Widom 601 DNA (Protein Data
Bank entry 3MVD). In this view, the DNA superhelix axis lies hori-

zontally and parallel to the plane of the page. The two RCC1 mole-
cules are shown in pale yellow and in magenta. The DNA is
represented by an orange rod (backbone) and blue–green sticks
(bases), and histone H3, H4, H2A and H2B are shown in different
colors. The two RCC1 molecules undergo equivalent interactions
on each side of the nucleosome core particle. Prepared with
PYMOL
().
F. J. Blanco and G. Montoya Transient protein–DNA ⁄ RNA interactions
FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS 1647
at the DNA interface, but they are not efficient for
complexes in which indirect readout is dominant in
DNA sequence recognition [41]. An increase in the
number of the crystal structures of proteinÆDNA com-
plexes should help to overcome this limitation [42,43].
Challenges and the way ahead
Most structural studies are carried out not with full-
length proteins, but with fragments. The most frequent
reasons for this are the difficulty in producing large
amounts of homogeneous material of large proteins,
and the simplification of the system to facilitate crys-
tallization and ⁄ or analysis by other techniques. How-
ever, investing time and effort in preparing and
analyzing the full-length protein can be extremely
rewarding, as shown by the information obtained with
the tumor suppressor protein BRCA2 and its interac-
tion with DNA [44,45]. As compared with protein Æ
protein complexes, there is still little structural infor-
mation on proteinÆnucleic acid complexes, especially
for chromatin enzymes and factors. The transient nat-

ure of many of the interactions is probably one of the
major difficulties in their identification, isolation, and
structural characterization. MS is emerging as a potent
tool for the study of dynamic or heterogeneous pro-
teinÆnucleic acid complexes [46]. Although not a high-
resolution structural technique, it has the bonus of
requiring very little material. Small amounts are also
used in single-molecule techniques, which are becoming
a fruitful approach to answer specific questions,
besides providing spectacular demonstrations of our
prowess in manipulating and observing protein.nucleic
acid complexes. Recent studies have addressed RNA
folding by a helicase [47], DNA transport [48], and
DNA polymerization [49].
Crystallography will continue to be the main tech-
nique for high-resolution studies. Tomographic electron
microscopy can provide structures of proteinÆ
nucleic acid complexes inside the cell [50], and NMR
has the potential to do so [51]. NMR is uniquely suited
to characterize folding–unfolding events that occur at
disordered regions of proteins that become structured
upon recognition of their target nucleic acids, and can
be usefully complemented by small-angle X-ray scatter-
ing [52,53]. Intrinsically disordered proteins or protein
regions will increasingly be the focus of structural stud-
ies as regulators of molecular recognition processes [54].
Acknowledgements
The work in the authors’ laboratories is supported by
grants from the Ministerio de Ciencia e Innovacio
´

n
(CTQ2008-03115 ⁄ BQU to F. J. Blanco, and BFU2008-
01344 to G. Montoya).
References
1 Doudna JA (2000) Structural genomics of RNA. Nat
Struct Biol 7(Suppl), 954–956.
2 Gavin AC, Bosche M, Krause R, Grandi P, Marzioch
M, Bauer A, Schultz J, Rick JM, Michon AM, Cruciat
CM et al. (2002) Functional organization of the yeast
proteome by systematic analysis of protein complexes.
Nature 415, 141–147.
3 Maga G & Hubscher U (2003) Proliferating cell nuclear
antigen (PCNA): a dancer with many partners. J Cell
Sci 116, 3051–3060.
4 Fridman Y, Palgi N, Dovrat D, Ben-Aroya S, Hieter P
& Aharoni A (2010) Subtle alterations in PCNA–part-
ner interactions severely impair DNA replication and
repair. PLoS Biol 8, e1000507.
5 McNally R, Bowman GD, Goedken ER, O’Donnell M
& Kuriyan J (2010) Analysis of the role of PCNA–
DNA contacts during clamp loading. BMC Struct Biol
10, 3, doi:10.1186/1472-6807-10-3.
6 Sanchez R, Torres D, Prieto J, Blanco FJ & Campos-
Olivas R (2007) Backbone assignment of human prolif-
erating cell nuclear antigen. Biomol NMR Assign 1,
245–247.
7 Campos-Olivas R, Sanchez R, Torres D & Blanco FJ
(2007) Backbone assignment of the 98 kDa homotrimer-
ic yeast PCNA ring. J Biomol NMR 38, 167.
8 Sanchez R, Pantoja-Uceda D, Prieto J, Diercks T,

Marcaida MJ, Montoya G, Campos-Olivas R &
Blanco FJ (2010) Solution structure of human growth
arrest and DNA damage 45alpha (Gadd45alpha) and
its interactions with proliferating cell nuclear antigen
(PCNA) and Aurora A kinase. J Biol Chem 285,
22196–22201.
9 De Biasio A, Sanchez R, Prieto J, Villate M, Campos-
Olivas R & Blanco FJ (2011) Reduced stability and
increased dynamics in the human proliferating cell
nuclear antigen (PCNA) relative to the yeast homolog.
PLoS ONE 6, e16600.
10 Remus D, Beuron F, Tolun G, Griffith JD, Morris EP
& Diffley JF (2009) Concerted loading of Mcm2-7 dou-
ble hexamers around DNA during DNA replication
origin licensing. Cell 139, 719–730.
11 Bohrmann B, Haider M & Kellenberger E (1993) Con-
centration evaluation of chromatin in unstained resin-
embedded sections by means of low-dose ratio-contrast
imaging in STEM. Ultramicroscopy 49, 235–251.
12 Clore GM & Iwahara J (2009) Theory, practice, and
applications of paramagnetic relaxation enhancement
for the characterization of transient low-population
states of biological macromolecules and their
complexes. Chem Rev 109 , 4108–4139.
Transient protein–DNA ⁄ RNA interactions F. J. Blanco and G. Montoya
1648 FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS
13 Doucleff M & Clore GM (2008) Global jumping and
domain-specific intersegment transfer between DNA
cognate sites of the multidomain transcription factor
Oct-1. Proc Natl Acad Sci USA 105, 13871–13876.

14 Takayama Y, Sahu D & Iwahara J (2010) NMR studies
of translocation of the Zif268 protein between its target
DNA Sites. Biochemistry 49, 7998–8005.
15 Kalodimos CG, Biris N, Bonvin AM, Levandoski MM,
Guennuegues M, Boelens R & Kaptein R (2004) Struc-
ture and flexibility adaptation in nonspecific and specific
protein–DNA complexes. Science 305, 386–389.
16 Popovych N, Tzeng SR, Tonelli M, Ebright RH &
Kalodimos CG (2009) Structural basis for cAMP-
mediated allosteric control of the catabolite
activator protein. Proc Natl Acad Sci USA 106,
6927–6932.
17 Nikolova EN, Kim E, Wise AA, O’Brien PJ, Andri-
cioaei I & Al-Hashimi HM (2011) Transient Hoogsteen
base pairs in canonical duplex DNA. Nature 470, 498–
502.
18 Farnham PJ (2009) Insights from genomic profiling of
transcription factors. Nat Rev Genet 10, 605–616.
19 Rohs R, Jin X, West SM, Joshi R, Honig B & Mann
RS (2010) Origins of specificity in protein–DNA recog-
nition. Annu Rev Biochem 79, 233–269.
20 Schuwirth BS, Borovinskaya MA, Hau CW, Zhang W,
Vila-Sanjurjo A, Holton JM & Cate JH (2005) Struc-
tures of the bacterial ribosome at 3.5 A resolution. Sci-
ence 310, 827–834.
21 Fischer N, Konevega AL, Wintermeyer W, Rodnina
MV & Stark H (2010) Ribosome dynamics and tRNA
movement by time-resolved electron cryomicroscopy.
Nature 466, 329–333.
22 Ito T & Yokoyama S (2010) Two enzymes bound to

one transfer RNA assume alternative conformations for
consecutive reactions. Nature 467, 612–616.
23 Travers A & Muskhelishvili G (2007) A common
topology for bacterial and eukaryotic transcription
initiation? EMBO Rep 8, 147–151.
24 Luger K, Mader AW, Richmond RK, Sargent DF &
Richmond TJ (1997) Crystal structure of the nucleo-
some core particle at 2.8 A resolution. Nature 389,
251–260.
25 Segal E & Widom J (2009) Poly(dA:dT) tracts: major
determinants of nucleosome organization. Curr Opin
Struct Biol 19, 65–71.
26 Engeholm M, de Jager M, Flaus A, Brenk R, van Noo-
rt J & Owen-Hughes T (2009) Nucleosomes can invade
DNA territories occupied by their neighbors. Nat Struct
Mol Biol 16, 151–158.
27 Kaplan N, Moore IK, Fondufe-Mittendorf Y, Gossett
AJ, Tillo D, Field Y, LeProust EM, Hughes TR, Lieb
JD, Widom J et al. (2009) The DNA-encoded nucleo-
some organization of a eukaryotic genome. Nature 458,
362–366.
28 Zhang Y, Moqtaderi Z, Rattner BP, Euskirchen G,
Snyder M, Kadonaga JT, Liu XS & Struhl K (2009)
Intrinsic histone–DNA interactions are not the major
determinant of nucleosome positions in vivo. Nat Struct
Mol Biol 16, 847–852.
29 Kaplan N, Moore I, Fondufe-Mittendorf Y, Gossett
AJ, Tillo D, Field Y, Hughes TR, Lieb JD, Widom J &
Segal E (2010) Nucleosome sequence preferences influ-
ence in vivo nucleosome organization. Nat Struct Mol

Biol 17, 918–920; author reply 920-912.
30 Zhang Y, Moqtaderi Z, Rattner BP, Euskirchen G,
Snyder M, Kadonaga JT, Liu XS & Struhl K (2010)
Evidence against a genomic code for nucleosome posi-
tioning. Nat Struct Mol Biol 17
, 920–923.
31 Pugh BF (2010) A preoccupied position on nucleo-
somes. Nat Struct Mol Biol 17, 923.
32 Makde RD, England JR, Yennawar HP & Tan S
(2010) Structure of RCC1 chromatin factor bound to
the nucleosome core particle. Nature 467, 562–566.
33 Vasudevan D, Chua EY & Davey CA (2010) Crystal
structures of nucleosome core particles containing the
‘601’ strong positioning sequence. J Mol Biol 403,1–
10.
34 Davey CA, Sargent DF, Luger K, Maeder AW & Rich-
mond TJ (2002) Solvent mediated interactions in the
structure of the nucleosome core particle at 1.9 a resolu-
tion. J Mol Biol 319, 1097–1113.
35 Arold ST, Leonard PG, Parkinson GN & Ladbury JE
(2010) H-NS forms a superhelical protein scaffold for
DNA condensation. Proc Natl Acad Sci USA 107,
15728–15732.
36 Cathomen T & Schambach A (2010) Zinc-finger nuc-
leases meet iPS cells: zinc positive: tailored genome
engineering meets reprogramming. Gene Ther 17, 1–3.
37 Marcaida MJ, Munoz IG, Blanco FJ, Prieto J &
Montoya G (2010) Homing endonucleases: from
basics to therapeutic applications. Cell Mol Life Sci
67, 727–748.

38 Redondo P, Prieto J, Munoz IG, Alibes A, Stricher F,
Serrano L, Cabaniols JP, Daboussi F, Arnould S, Perez
C et al. (2008) Molecular basis of xeroderma pigmento-
sum group C DNA recognition by engineered meganuc-
leases. Nature 456, 107–111.
39 Li T, Huang S, Jiang WZ, Wright D, Spalding
MH, Weeks DP & Yang B (2011) TAL nucleases
(TALNs): hybrid proteins composed of TAL effectors
and FokI DNA-cleavage domain. Nucleic Acids Res
39, 359–372.
40 Thyme SB, Jarjour J, Takeuchi R, Havranek JJ, Ash-
worth J, Scharenberg AM, Stoddard BL & Baker D
(2009) Exploitation of binding energy for catalysis and
design. Nature 461, 1300–1304.
41 Ashworth J & Baker D (2009) Assessment of the opti-
mization of affinity and specificity at protein–DNA
interfaces. Nucleic Acids Res 37, e73.
F. J. Blanco and G. Montoya Transient protein–DNA ⁄ RNA interactions
FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS 1649
42 Marcaida MJ, Prieto J, Redondo P, Nadra AD, Alibes
A, Serrano L, Grizot S, Duchateau P, Paques F, Blanco
FJ et al. (2008) Crystal structure of I-DmoI in
complex with its target DNA provides new insights into
meganuclease engineering. Proc Natl Acad Sci USA
105, 16888–16893.
43 Munoz IG, Prieto J, Subramanian S, Coloma J,
Redondo P, Villate M, Merino N, Marenchino M,
D’Abramo M, Gervasio FL et al. (2011) Molecular
basis of engineered meganuclease targeting of the
endogenous human RAG1 locus. Nucleic Acids Res 39,

729–743.
44 Jensen RB, Carreira A & Kowalczykowski SC (2010)
Purified human BRCA2 stimulates RAD51-mediated
recombination. Nature 467, 678–683.
45 Thorslund T, McIlwraith MJ, Compton SA, Lekomtsev
S, Petronczki M, Griffith JD & West SC (2010) The
breast cancer tumor suppressor BRCA2 promotes the
specific targeting of RAD51 to single-stranded DNA.
Nat Struct Mol Biol 17, 1263–1265.
46 Gordiyenko Y & Robinson CV (2008) The emerging
role of MS in structure elucidation of protein–nucleic
acid complexes. Biochem Soc Trans 36, 723–731.
47 Karunatilaka KS, Solem A, Pyle AM & Rueda D
(2010) Single-molecule analysis of Mss116-mediated
group II intron folding. Nature 467, 935–939.
48 Ptacin JL, Nollmann M, Becker EC, Cozzarelli NR,
Pogliano K & Bustamante C (2008) Sequence-directed
DNA export guides chromosome translocation during
sporulation in Bacillus subtilis. Nat Struct Mol Biol 15,
485–493.
49 Olasagasti F, Lieberman KR, Benner S, Cherf GM,
Dahl JM, Deamer DW & Akeson M (2010) Replication
of individual DNA molecules under electronic control
using a protein nanopore. Nat Nanotechnol 5, 798–806.
50 Brandt F, Carlson LA, Hartl FU, Baumeister W &
Grunewald K (2010) The three-dimensional organiza-
tion of polyribosomes in intact human cells. Mol Cell
39, 560–569.
51 Sakakibara D, Sasaki A, Ikeya T, Hamatsu J, Hanashi-
ma T, Mishima M, Yoshimasu M, Hayashi N, Mikawa

T, Walchli M et al. (2009) Protein structure determina-
tion in living cells by in-cell NMR spectroscopy. Nature
458, 102–105.
52 Mertens HD & Svergun DI (2010) Structural character-
ization of proteins and complexes using small-angle
X-ray solution scattering. J Struct Biol 172, 128–141.
53 Jimenez-Menendez N, Fernandez-Millan P, Rubio-Cosi-
als A, Arnan C, Montoya J, Jacobs HT, Bernado P,
Coll M, Uson I & Sola M (2010) Human mitochondrial
mTERF wraps around DNA through a left-handed
superhelical tandem repeat. Nat Struct Mol Biol 17,
891–893.
54 Garcia-Pino A, Balasubramanian S, Wyns L, Gazit E,
De Greve H, Magnuson RD, Charlier D, van Nuland
NA & Loris R (2010) Allostery and intrinsic disorder
mediate transcription regulation by conditional cooper-
ativity. Cell 142, 101–111.
55 Hodis E, Prilusky J, Martz E, Silman I, Moult J &
Sussman JL (2008) Proteopedia – a scientific ‘wiki’
bridging the rift between three-dimensional structure and
function of biomacromolecules. Genome Biol 9, R121.
Transient protein–DNA ⁄ RNA interactions F. J. Blanco and G. Montoya
1650 FEBS Journal 278 (2011) 1643–1650 ª 2011 The Authors Journal compilation ª 2011 FEBS

×