Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jun 1.
Published in final edited form as: Curr Opin Struct Biol. 2011 Apr 7;21(3):348–357. doi: 10.1016/j.sbi.2011.03.006

Working the kinks out of nucleosomal DNA

Wilma K Olson 1, Victor B Zhurkin 2
PMCID: PMC3112303  NIHMSID: NIHMS288175  PMID: 21482100

Abstract

Condensation of DNA in the nucleosome takes advantage of its double-helical architecture. The DNA deforms at sites where the base pairs face the histone octamer. The largest so-called kink-and-slide deformations occur in the vicinity of arginines that penetrate the minor groove. Nucleosome structures formed from the 601 positioning sequence differ subtly from those incorporating an AT-rich human α-satellite DNA. Restraints imposed by the histone arginines on the displacement of base pairs can modulate the sequence-dependent deformability of DNA and potentially contribute to the unique features of the different nucleosomes. Steric barriers mimicking constraints found in the nucleosome induce the simulated large-scale rearrangement of canonical B-DNA to kink-and-slide states. The pathway to these states shows non-harmonic behavior consistent with bending profiles inferred from AFM measurements.

Introduction

The exquisite management of long, double-helical DNA in the tight confines of the eukaryotic nucleus —a highly condensed chromatin structure that is remarkably accessible to the genetic machinery —takes advantage of structural and energetic signals encoded in the nucleotide sequence. For example, the eight histone proteins that make up the core of the nucleosome —the fundamental packaging unit of ~150 bp of eukaryotic DNA in chromatin —bind preferentially to so-called ‘positioning’ sequences with unique nucleotide repeating patterns. The proteins, however, make few, if any, direct contacts with the DNA bases in known high-resolution structures [1-4], seemingly recognizing the intrinsic conformation and/or the deformability of the DNA wrapped on their surfaces [5,6]. The sequence-neutral aspects of this association allow for the binding of nucleosomes on arbitrary genomic sequences and the concomitant compaction of DNA as a whole. The weak sequence-specificity of nucleosome positioning underlies the precise fold of the nucleosomal array and the access of proteins and other ligands to the encoded genetic information. The latter molecules can take advantage of the exposed nucleotides on the outer surfaces of the nucleosomes or, alternatively, compete with the histones for access to the nucleotides at the DNA-protein interface [7,8].

Until recently much of what was known about the contribution of nucleotide sequence to nucleosome structure and positioning derived from the superhelical pathway of the palindromic, A+T-rich human α-satellite DNA sequence featured in almost every solved crystal complex. The DNAs in these assemblies follow staircase-like routes with fragments of naturally straight, B-like helices interrupted by sharp turns/dislocations in three-dimensional structure [5], i.e., ~5 bp ‘treads’ of B DNA joined by dimeric ‘risers’ generated by concomitant bending and lateral displacement of successive base pairs (Figure 1). Moreover, these ‘kink-and-slide’ deformations occur almost exclusively at regularly spaced CA· TG steps along the DNA, i.e., paired dinucleotides made up of successive C· G and A· T base pairs, and act in concert to set the global curvature and superhelical pitch of the wrapped duplex [5,9,10]. The CA· TG steps are thought to be naturally deformable, not only taking up the major distortions of DNA in many nucleosome core-particle structures but also adopting a wide range of conformational states in other protein-bound complexes [6,11] and converting to highly distorted states at low energetic cost in molecular simulations [4,12].

Figure 1.

Figure 1

(a) Simplified representation, adapted from [5], of the staircase-like superhelical pathway taken by DNA in high-resolution nucleosome crystal structures. The ‘steps’, depicted by yellow rods, span 5-bp stretches of an ideal B-DNA double helix and run in the direction of the local helical axes. Red and blue arrows at the ‘risers’ highlight the lateral displacements associated with the histone-induced deformations that alternatingly convert DNA base-pair steps to A- and C-like helical states. The localized bending of adjacent base pairs into the major and minor grooves (red and blue arrows, respectively) occurs in concert with the lateral displacements and twisting needed to effect the B→A and B→C conformational transformations. The dyad passing through the central base pair is depicted by the diamond and the global axis of the nucleosome by the dotted black line. (b) Atomic-level view of the two helical turns of DNA centered about the dyad of the nucleosomal model. Note the dislocations of the helical axes (yellow lines joining the centers of base pairs) at the sites of DNA deformation and the uniform direction of helical-axis displacement in the detailed and simplified images, i.e., along the global axis of the nucleosome.

Compared to a decade ago when the ‘rules’ of nucleotide deformability were first extracted from the configurations of DNA in high-resolution crystal structures [11] and years earlier when the concept of DNA sequence-dependent helical structure was formulated from the primary sequences of naturally occurring DNAs [13] and the sequence-dependent flexibility of DNA was deduced from the cyclization propensities and gel mobility profiles of short polymeric chains [14,15], there are now more than 3,000 publicly available DNA-containing high-resolution (X-ray and NMR) structures and even larger numbers of low-resolution images of DNA obtained from electron microscopy (EM) and atomic force microscopy (AFM) studies [16,17]. In other words, instead of interpreting an average property of a DNA sequence, such as its gel retardation, in terms of a simplified physical (harmonic) model, one can decipher the sequence-dependent behavior of a specific DNA from large numbers of configurations of the molecule. In addition, computational advances now make it possible to examine the influence of local sequence-induced deformations of DNA helical structure on the spatial configurations and properties of increasingly longer chains [18,19] and to reconsider the precise mechanics of nucleotide-dependent deformations of DNA [4,20]. For example, naturally flexible sequences, such as those containing CA· TG and other pyrimidine-purine, or YR, dimers, do not necessarily follow the ideal harmonic behavior expected from classical physical models [4]. Understanding the precise details of extreme helical distortions helps make sense of the observed positioning of nucleosomes and the binding of other proteins on DNA.

New surprises in the crystal structures of positioned nucleosomes

The past year has witnessed the elucidation of five new nucleosome structures incorporating DNAs other than the popular α-satellite sequence [21-23], and thus provided the opportunity to learn more about the conformational responses of DNA on the nucleosome and the implications of these findings on the positioning of nucleosomes in chromatin. Significantly three of the new structures [21,22] incorporate the synthetic ‘601’ sequence found to bind the histone octamer with very high affinity and widely used to set the positions of nucleosomes in experimental studies of chromatin [24]. Although the ends of the crystallized sequences terminate with different bases, the central 139 nucleotides include the known key repeating patterns found in solution —namely, five TA base-pair steps, i.e., TA· TA dimers, which recur at increments of ~10 bp, i.e., roughly complete double-helical turns, and the alternation, at half-helical turns, of A+T- and G+C-rich motifs long known to characterize native nucleosomal sequences [25] and recently detected in nucleosomes on the yeast genome [26]. Moreover, these sequence preferences play a central role in the organization of nucleosomes in vivo, such as positioning nucleosomes in the vicinity of elements that regulate transcription [27].

The crystallized 601 sequences occupy the same position on the histone core as that mapped to single-nucleotide resolution by site-directed hydroxyl-radical cleavage [28], i.e., the expected base pair lies on the structural dyad. One of the 601-containing structures (PDB_ID 3mvd) includes the Drosophila RCC1 (regulator of chromosome condensation) protein implicated in disparate cellular functions [21]. The other structures (PDB_IDs 3lz0, 3lz1) focus on the subtle differences between the two different asymmetric settings of the non-palindromic 601 sequence on the histone assembly [22].

The deformable nature of the TA dinucleotide steps in many observed and simulated structures [4,6,11,29] has long hinted of a conformational role for the periodically spaced TA steps in the 601 sequence much like that taken by the CA· TG steps in the many nucleosomes incorporating α-satellite DNA. Surprisingly, the crystallized 601 sequences make different use of the naturally deformable, periodically phased TA steps in binding the histone core. Although the phased TA steps lie in the expected rotational settings, i.e., with minor-grove edges facing the histone core, they constitute only a small proportion of the kink-and-slide steps in the structures. Most of the phased TA steps deform more ‘smoothly’, with the histone-induced bends into the minor groove and/or lateral displacement (Slide) of base pairs occurring in adjacent base-pair steps. This deformational ‘smoothing’ leads to greater conformational variability of DNA in the 601-containing nucleosomes compared to that found in nucleosomes containing the α-satellite sequence (Figure 2).

Figure 2.

Figure 2

Scatter plots in the Slide-Roll and Roll-Tilt planes of the observed parameters and derived ‘energy’ contours of base-pair steps from nucleosomes incorporating α-satellite [30,31] and 601 [21,22] DNA (upper and lower images, respectively). Parameters in individual nucleosome structures are denoted by the symbols specified in the inset. Ellipses are projections of the six-dimensional ‘equipotential’ surfaces obtained from covariance analysis [11]. The contours correspond to ‘energies’ of 4.5, i.e., ellipses where parameters deviate from their mean values by no more than 3Δθ, three times the root-mean-square deviation Δθ. Note (i) the broader ranges of Roll and Slide and the more isotropic bending in the 601 dimers compared to the α-satellite steps, (ii) the more extreme deformations in the nucleosome containing 145 vs. 147-bp α-satellite DNA (2nzd vs. 1kx5), and (iii) the more limited deformations in the 601 nucleosome binding RCC1 versus those crystallized in the absence of the chromosomal protein (3mvd vs. 3lz0 and 3lz1). The kink-and-slide steps discussed in the text are those where Roll. −7°, Slide. 1.4 Å (C-like steps) and Roll. +7°, Slide. −1 Å (A-like steps). Block images on the right illustrate positive values of the three parameters. The lightly shaded edges of the blocks correspond to the minor-groove edges of the base pairs. Step parameters and images computed with the 3DNA software [49,50].

Some of the kink-and-slide steps in the 601 sequences entail CA· TG dimers and several involve less naturally deformable RR· YY dimers (see the overlap/non-overlap of kink-and-slide vs. YR steps in the mosaic in Figure 3). In contrast to the isolated CA· TG kink-and slide steps found in nucleosomes containing α-satellite DNA, some of the strongly distorted CA· TG dimers in the 601 nucleosomes abut a comparably distorted base-pair step (noted by ‘double’ bars in Figure 3). The crystallized 601-containing nucleosomes also include a higher than random proportion of CG dimers (~10%), other easily distorted YR steps [11] which are completely missing in the many of the nucleosomes incorporating α-satellite DNA. Most of the CG steps in the 601 nucleosomes occur at sites where the DNA bends into the major groove. Only one CG step (that at step 56 in the 3lz1 structure [22]) adopts a kink-and-slide conformation that narrows the DNA minor groove. As is clear from Figure 3, some of the kink-and-slide steps in the 601 nucleosomes coincide with the corresponding deformations in the 145-bp α-satellite DNA, specifically at the ends of the chain binding the H2A· H2B dimer. The deformations of 601 vs. α-satellite DNA, however, differ substantially in center of the nucleosome, where the DNA is in contact with the (H3· H4)2 tetramer (Figure 3). The binding of the RCC1 chromosomal protein has only a small effect on the kink-and-slide pattern of the 601 sequence (compare entries for 3lz0 and 3mvd in Figure 3). In general, the 601 sequence is more strongly deformed around the histone core than the α-satellite DNA, e.g., more extreme bending between successive base pairs in the nucleosomal 601 sequences compared to the DNA found in the best-resolved core-particle structure (PDB_ID 1kx5) [30] incorporating a 147-bp α-satellite sequence. The areas of the ellipses constructed in Figure 2 from the values of Roll and Tilt in the 601- and α-satellite-containing nucleosomes illustrate these differences.

Figure 3.

Figure 3

Color-coded depictions of the kink-and-slide deformations, YR base-pair steps, and the variation in helical twist in nucleosome core-particle structures incorporating fragments of α-satellite [31] and 601 [21,22] DNA. Nucleotide positions are reported with respect to the central base pairs of the crystallized 145-bp sequences at position 0. The numerals below the graph denote the base-pair position, and the numerals above the mosaic show the superhelical position [1], i.e., the number of helical turns of a given base pair from the dyad. The Protein Data Bank identifiers (PDB_IDs) of individual structures are noted on the left of the mosaic and in the plot inset. The sites of C-like kink-and-slide deformations in individual structures, where Roll. −7° and Slide. 1.4 Å, are denoted by the black lines in the upper half of each structural ‘bar code’. The identities and locations of YR steps are depicted by the differently colored lines in the lower half of each entry —CA· TG (blue), CG (green), TA (red). Note (i) the very different sequence composition and distribution of kink-and-slide steps in the two nucleosomal sequences (601 in 3lz0 and 3mvd vs. α-satellite in 2nzd), (ii) the coincidence of most kink-and-slide states with sites where the DNA is overtwisted, with close to the 9 bp/turn helical repeat characteristic of C DNA, and (iii) the differences in twist at the ends vs. middle of the 601 sequence in rough agreement with the average values of 10.0 and 10.5 bp/turn detected experimentally on the 5S RNA positioning sequence [37]. Helical twist is computed from the twist of supercoiling Twsc [36] and expressed in terms of the number of residues per turn, i.e., 360 /〈Twsc〉, where 〈Twsc〉 is the average value, in degrees, over five successive base-pair steps.

The cost of ‘threading’ DNA [5,6] on the pathways reported in the 601-containing nucleosome structures is higher than that for the DNA found in nucleosomes incorporating α-satellite sequences [9], i.e., there are more 601 base-pair steps far from their minimum-energy rest states. The deformation scores of the 601-bound DNAs, based on sequence-dependent potentials derived from the structures of other protein-bound DNA molecules [6], are over threefold greater than that for the 147-bp α-satellite DNA and twofold higher than that for the ‘stretched’ 145-bp α-satellite DNA with two G· C pairs missing on either side (±3 bps) of the dyad (PDB_ID 2nzd) [31]. In other words, the nucleotide steps along the histone-bound 601 sequences exhibit conformational responses more drastic than those found in other nucleosome structures. The deformation scores of the 601 sequences also exceed those of the DNA in nucleosomes of comparable resolution, e.g., scores of 1624 and 2118 for the two settings of the 601 sequence crystallized at 2.5-Å resolution on Xenopus histones [21] vs. scores of 1086 and 750 for the α-satellite DNA wrapped respectively around chicken [32] and human [33] histone assemblies (PDB_IDs 1eqz and 2cv5). (Note that these scores take no account of the sizable distortions of the Watson-Crick base pairs in the 601 structures; see below.) Whether the greater scores of the 601 DNA reflect real differences in macromolecular sequence and structure or differences in data collection and refinement methods are open questions.

There is yet another difference between the two sets of structures, which is hidden within the double-stranded (base-pair step) parameters described above but visible in the rigid-body parameters describing the spatial arrangements of complementary and/or successive bases, i.e., base-pair and single-stranded base-step parameters. Specifically, the base pairs comprising the histone-bound 601 sequence show signs of ‘melting’ not found in earlier structures of nucleosomes containing α-satellite DNA. The Watson-Crick geometry, particularly that of A· T pairs, is highly distorted. For example, the planes of A and T are systematically staggered (vertically displaced) in the 3′ -direction by ~0.3 Å on average from the ‘natural’ 0 Å rest state found in other nucleosomal DNAs and by as much as 2.5 Å at some locations. Moreover, the angle between base planes, measured in terms of Buckle and/or Propeller, sometimes exceeds 50°. Whether these structural signals mirror the measured sequence-dependent thermal stabilities of DNA [34], and in particular the extremely low stability of TA steps, is an open question. Further analysis of the stacking of bases along the DNA strands, i.e., single-stranded base-step parameters, and the interactions of histones with the individual strands may shed light on this question.

The three 601 sequences form 145-bp core particles spanning the same number of double-helical turns as both the 147-bp α-satellite DNA in the best-resolved crystal structure [30] and the related 145-bp nucleosomal sequence [31]. All five structures are slightly (2-4%) overstretched compared to ideal B DNA —a known numerical consequence [35] of the nonplanar buckling of Watson-Crick geometry that accompanies the insertion of arginines into the narrow DNA minor groove at locations where the minor groove faces the histones. The alteration of sequence, however, has subtle, previously unreported effects on the helical twist and the apparent setting of specific base pairs on the histone core. Although the average twist of the 601 sequences is identical to that of the nucleosomal 145-bp α-satellite DNA (10.4 bp/turn, based on the so-called twist of supercoiling [36]), the local repeating patterns differ appreciably over two stretches centered around superhelical positions ±2.5, i.e., 20-30 bp from the dyad in the vicinity of the dimerization interface between histones H3-H4 and H2A-H2B, and over two stretches at either end of the DNA (Figure 3). Differences such as these, which can be detected with chemical footprinting techniques, may account for the apparent discrepancies between early solution measurements of the helical twist of nucleosomal DNA [37] compared to that of the DNA captured in high-resolution structures.

The 601 and α-satellite sequences share limited identity. Alignment of the crystallized 145-bp sequences with respect to the base pair on the dyad of each structure shows that only 33 bases lie in the same register and that only six base-pair steps are identical, i.e., correspondence at the level of two randomly occurring sequences. The spacing of YR dimers also differs in the two sequences, with a greater concentration of such steps at the ends of the α-satellite DNA and in the middle of 601 DNA (Figure 3). The regularly spaced TA dimers in the 601-nucleosome structure drape over the surface of the (H3· H4)2 tetramer rather than the H2A· H2B dimers. Many of the kink-and-slide deformations of the 601 DNA found at H2A· H2B binding sites entail RR· YY or RY steps. Other sequences with YR steps optimized to fit all the arginine binding sites may thus fit on nucleosomes more tightly than the 601 sequence. Clearly, many more new high-resolution structures with a greater variety of bound DNA sequences are needed to understand the physical basis of DNA deformability and nucleosome positioning.

Insights into sequence-dependent DNA deformations from atomic simulations

Despite the many examples of extreme and complex deformations of DNA in the structures of nucleosomes and other protein-DNA complexes, most of the theories used to interpret the sequence-dependent properties of DNA still rely on oversimplified physical models of the double helix, e.g., DNA treated as an ideal wormlike chain subject to isotropic bending without consideration of either the coupling of bending to twisting and lateral displacement or the effects of localized sequence-dependent deformability [38]. Clear evidence of the breakdown of the classic DNA models comes from the unexpected prevalence of large bends in DNA chains of a few thousand base pairs in AFM images [16]. Interestingly, the cost of duplex bending extracted from the observed deflections of DNA does not follow simple quadratic behavior.

Although the tightly clustered, Gaussian distributions of the rigid-body parameters used to characterize the spatial arrangements of DNA base pairs in high-resolution structures are consistent with a quadratic increase in energy for small thermal fluctuations, e.g., dimeric bend angles of 10-20° (directed toward the grooves), the conformational dependence of the deformation energy is not clear for the isolated, severely distorted states captured in the structures of numerous protein-DNA complexes [6,11] and in specific crystallographic lattices [39]. The increase in energy with sharp bending, when the backbone conformation is altered and neighboring base-stacking interactions are lost, becomes less than quadratic. That is, the second derivative of the energy with respect to bending deflection decreases for large angles compared to small angles. One (out of many possibilities) is that the energy function becomes quasi-linear for angles greater than some threshold, e.g., 40-50° [16].

Indeed, this quasi-linear behavior has surfaced in recent computer simulations of DNA subject to strong bending. The uptake of extreme deformations depends upon the sequence and the assumed force field. In particular, the Lavery group [20] has found that implementation of the updated parmbsc0 force field [40] in molecular dynamics simulations of oligonucleotide duplexes yields different types of kinks in GC-rich versus AT-rich DNA. Whereas the kinks of the GC-rich oligomers are directed toward the major groove, those at the 3′ -end of an A-tract, in the AAC trimer, are directed toward the minor groove. These structural distortions, termed kinks II, span three consecutive base pairs as opposed to a ‘classical’ kink I, which unstacks a single base-pair step. By contrast, use of the older parm94 force field [41] generates both type I and II kinks, which enhance the writhing of small DNA minicircles [18].

The sequence-dependent deformability of DNA also depends on the imposed constraints that mimic the presence of bound protein (Figure 4). For example, modeled, ‘protein-free’ hexameric duplexes [4] bend more easily in the direction of the minor groove at RY steps and in the direction of the major groove at YR steps, a finding obtained many years ago in simulations of DNA tetramers [29] and consistent with the stereochemical rules introduced by Calladine to account for subtle, sequence-dependent conformational effects seen in the earliest DNA crystal structures [42]. That is, major-groove bending relieves purine-purine clashes in the minor groove of YR steps, and minor-groove bending diminishes unfavorable clashes in the major groove of RY steps. If, however, the base pairs are laterally displaced, along their long axes, as at the kink-and-slide sites in the nucleosome core particle [30], the computed conformational preferences switch, with the YR steps easily bent into the minor groove. Moreover, the constraint imposed on Slide [4] also reproduces the large increase in twist and the BI→BII rearrangement of the DNA backbone usually found in combination with the kink-and-slide deformations in high-resolution structures and responsible for the transformation of the canonical B-DNA double helix with 10 bp per helical turn to the C-like form with 9 bp/turn [43]. The coupling of these and other conformational variables may account for the unexpectedly large kink-and-slide deformations of several RR· YY steps and the differences in twist found in the 601 vs. α-satellite nucleosome structures.

Figure 4.

Figure 4

(Top) Composite stick/space-filled atomic representations, adapted from [4], illustrating the relative overlap/repulsions of the purine bases on opposing strands of (left) TA and (right) AT base-pair steps constrained to kink-and-slide arrangements (Roll = −20°, Slide = 2.5 Å) like those found in high-resolution nucleosome core-particle structures. The magenta and blue arrows denote the respective directions of the bending deformation (into the minor groove via negative Roll) and the lateral displacement of base pairs (along their long axes via positive Slide). Note the weaker purine-purine repulsions (lesser overlap between pink and green surfaces) in the TA dimer compared to the AT dimer.

(Bottom) Two-dimensional (Roll vs. Slide) energy contour plots illustrating the greater deformability of the central TA base-pair step of the d(CTTAAG)2 hexamer duplex compared to the corresponding AT dimer in the d(GAATTC)2 structure with the reverse sequence [4]. DNA structures and conformational energies simulated with DNAminiCarlo [51]. Conformational energies are the values minimized over all hexamer arrangements with the specified (Roll, Slide) values. The contour line separation is 2 kcal/mol. Cyan circles: the optimal structures. Note that for TA, the optimal Roll > 0, while for AT, the optimal Roll < 0. Pink circles: the kink-and-slide arrangements illustrated above. White thick lines: boundary between BI and BII forms, i.e. structures with backbone dihedral angles ε and ζ switched from trans, gauche (BI) conformational states to gauche, trans (BII) arrangements.

Interestingly, the predicted uptake of energy depends on the direction of bending. Whereas the cost of deformations into the major groove shows a quadratic increase with (Roll) bend angles up to 20-25°, that for deformations into the minor groove is more linear [4]. The different responses may reflect some kind of ‘asymmetry’ associated with minor- vs. major-groove bends, such as the ‘exposure’ of the glycosidic bonds in the minor groove, but the precise reasons for this behavior are unknown. The bending of TA into the minor groove shows a significant non-harmonicity in the Roll angle and Slide displacement not found in the case of AT (Figure 4). In addition to the correlation between Roll and Slide, there is significant ‘widening’ of the energy ‘ravine’ between positive to negative Roll values (major-groove vs. minor-groove bending). This novel effect may ‘shift’ the Boltzmann equilibrium toward negative Roll values for TA dimeric steps.

Concluding remarks

The positions of nucleosomes on DNA determine the global organization of chromatin and the accessibility of the genomic sequence to the cellular machinery. High-resolution structures of nucleosome-binding sequences offer the unique opportunity to deduce the ‘rules’ that govern nucleosome positioning, i.e., the sites on DNA most likely to take up the severe kinks and lateral displacements of base pairs that give rise to the compact, superhelical fold of the double helix on the surface of the histone octamer. Currently, there is no consensus on the precise ‘rules’ of nucleosome positioning —see, for example, the significant difference of opinions expressed in the recent review by Trifonov [44] and the seven accompanying commentaries. Understanding the different uptake of the 601 sequence on the nucleosome in recent high-resolution structures [21,22] compared to the wrapping of α-satellite DNA found in earlier crystallographic studies is therefore critical for elucidating these ‘rules’ and drawing meaningful conclusions regarding the placement of nucleosomes relative to key elements on genomes.

The unexpected conformational response of the 601 positioning sequence to histone-core binding merits serious investigation. Why RR· YY base-pair steps heretofore characterized as naturally stiff undergo the large deformations expected of the regularly spaced, flexible TA dimers in the 601 sequence and why TA steps behave differently from CA· TG steps, i.e., DNA deformations are sometimes spread over 2-3 steps rather than localized at the TA dimer, are not clear. Re-refinement of the data with different software may lead to different solutions in better agreement with structural expectations much like the independent refinements decades ago of the DNA dodecamer structure [45-47]. For example, some of the CA· TG and CG dimers abutting the currently deformed RR· YY steps may take up the kink-and-slide states, much like the shuffled YR steps found in different α-satellite sequences (where the kink-and-slide steps move to neighboring positions [6]). On the other hand, histone-DNA complexes may not fit the general conformational trends deduced from other protein-DNA structures. That is, the arginines that penetrate the minor groove of nucleosomal DNA may present steric barriers that induce unexpected structural responses in what otherwise appear to be naturally stiff base-pair steps. Computational studies of appropriately constrained model systems along the lines of recent simulations of highly kinked DNA duplexes may shed light on this question.

Given the close spacing of the negatively charged phosphate groups across the DNA minor groove, bending deformations that further narrow the groove are generally considered unfavorable. The neutralization of the DNA phosphates (and the constraint imposed on Slide) by the histone arginines facilitates DNA bending into the minor groove. The highly distorted kink-and-slide states that effect this bending differ significantly from canonical B DNA. The sugar-phosphate backbone typically undergoes a large conformational rearrangement as opposed to randomly fluctuating within the conformation space of the B-DNA energy minimum. The likelihood of the change of DNA helical state, a BI→BII transition, should not necessarily follow the classical mechanics of ideal elastic rods or the sequence-dependent rules of DNA deformability found in ‘intact’ double-helical structures. The complex interplay of nucleotide sequence, DNA backbone architecture, base-pair geometry, groove structure, and histone-DNA interactions along the kink-and-slide pathway is only beginning to be understood [4].

Finally, given that the twist of simulated DNA duplexes increases significantly only if constraints are placed on the lateral displacement (Slide) of adjacent base pairs, the extreme bending of DNA detected in AFM and EM images likely differs in character from the kinked base-pair steps found in nucleosomes, in terms of both the direction of bending (major- vs. minor-groove) and the accompanying rearrangement of the sugar-phosphate backbone. Whether these bends entail BI→BII transitions or other large backbone rearrangements, e.g., sugar repuckering, is an open question. Intriguingly, CG and CA· TG are among the base-pair steps most likely to adopt the BII form in NMR solution studies of protein-free B-DNA duplexes [48]. The transition propensity, which is inferred from P31 NMR chemical shifts, is substantially lower for TA steps. The greater propensity of GC-rich sequences for the BII form may be one of the reasons why the TA steps in the three recent 601 nucleosome structures behave differently from the CA· TG steps in the many nucleosomes formed with α-satellite DNA.

Acknowledgments

We are grateful to Dr. Song Tan for sharing unpublished data and Drs. Andrew V. Colasanti and Difei Wang for assistance with data collections and figures. Support of this work through U.S.P.H.S. grant GM34809 is gratefully acknowledged.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References and recommended reading

  • 1.Luger K, Mä der AW, Richmond RK, Sargent DF, Richmond TJ. Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature. 1997;389:251–260. doi: 10.1038/38444. [DOI] [PubMed] [Google Scholar]
  • 2.García-Pérez M, Pinto M, Subirana JA. Nonsequence-specific arginine interactions in the nucleosome core particle. Biopolymers. 2003;69:432–439. doi: 10.1002/bip.10389. [DOI] [PubMed] [Google Scholar]
  • 3•.Rohs R, West SM, Sosinsky A, Liu P, Mann RS, Honig B. The role of DNA shape in protein-DNA recognition. Nature. 2009;461:1248–1253. doi: 10.1038/nature08473. The enhanced electrostatic potentials of the narrowed minor grooves of A+T-rich DNA suggest a mechanism by which cationic amino acids of various proteins, including the arginines on the surface of the histones in the nucleosome core particle, might recognize DNA sequence without direct base-pair contact. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4•.Wang D, Ulyanov NB, Zhurkin VB. Sequence-dependent kink-and-slide deformations of nucleosomal DNA facilitated by histone arginines bound in the minor groove. J Biomol Struct Dyn. 2010;27:843–859. doi: 10.1080/07391102.2010.10508586. The asymmetric placement of histone arginines in the DNA minor groove appears to stabilize the kink-and-slide deformations responsible for the superhelical folding of DNA on the surface of the nucleosome. Moreover, incorporation of constraints that mimic these interactions in atomic calculations facilitates the bending at YR steps into the minor groove, not only changing the predicted direction of bending in the absence of protein but also introducing a nonharmonic energetic response in the DNA. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5•.Tolstorukov MY, Colasanti AV, McCandlish D, Olson WK, Zhurkin VB. A novel ‘roll-and-slide’ mechanism of DNA folding in chromatin. Implications for nucleosome positioning. J Mol Biol. 2007;371:725–738. doi: 10.1016/j.jmb.2007.05.048. Base-pair level studies of nucleosomal DNA reveal the surprising role of the lateral base-pair displacements (Slide) on the superhelical pitch of DNA and the positioning of nucleosomes on different DNA sequences. The folding of the DNA involves a novel ‘kink-and-slide’ mechanism that is best accommodated by naturally deformable pyrimidine-purine steps. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6•.Balasubramanian S, Xu F, Olson WK. DNA sequence-directed organization of chromatin: structure-based analysis of nucleosome-binding sequences. Biophys J. 2009;96:2245–2260. doi: 10.1016/j.bpj.2008.11.040. The cost of deforming a collection of sequences known to bind or resist uptake in nucleosomes with an updated set of sequence-dependent ‘energy’ functions suggests that sequences which contain periodically spaced pyrimidine-purine steps fold along a kinked pathway and that sequences resistant to deformation follow a smoother spatial route. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kingston RE. A snapshot of a dynamic nuclear building block. Nature Struct Biol. 1997;4:763–666. doi: 10.1038/nsb1097-763. [DOI] [PubMed] [Google Scholar]
  • 8•.Sahu G, Wang D, Chen CB, Zhurkin VB, Harrington RE, Appella E, Hager GL, Nagaich AK. p53 binding to nucleosomal DNA depends on the rotational positioning of DNA response element. J Biol Chem. 2010;285:1321–1332. doi: 10.1074/jbc.M109.081182. The tumor suppressor p53 protein binds its recognition site on nucleosomal DNA when the site is ‘exposed’ on the nucleosome exterior, i.e., when the site deforms around the histone core in the same direction found in the p53-DNA complex in solution. The protein does not bind to its cognate DNA, however, when the orientation of the binding site on the nucleosome changes by 180°. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9•.Bishop TC. Geometry of the nucleosomal DNA superhelix. Biophys J. 2008;95:1007–1017. doi: 10.1529/biophysj.107.122853. Reconstruction of nucleosomal DNA pathways from analytical expressions (finite Fourier series) that capture the chain-length dependent variation of base-pair-step parameters demonstrate the key roles of Slide, Roll, and Twist in the wrapping of the double helix in high-resolution core-particle structures. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10•.Marathe A, Bansal M. An ensemble of B-DNA dinucleotide geometries lead to characteristic nucleosomal DNA structure and provide plasticity required for gene expression. BMC Structural Biology. 2011;11 doi: 10.1186/1472-6807-11-1. http://www.biomedcentral.com/1472-6807/1411/1471. Detailed analyses of the backbone torsions, base-pair geometry, groove widths, and long-range bending of DNA in representative nucleosome structures find Slide to be “the principal dinucleotide parameter” characterizing the chain pathway and point to limitations of elastic treatments of nucleosomal DNA. [DOI] [PMC free article] [PubMed]
  • 11.Olson WK, Gorin AA, Lu X-J, Hock LM, Zhurkin VB. DNA sequence-dependent deformability deduced from protein-DNA crystal complexes. Proc Natl Acad Sci USA. 1998;95:11163–11168. doi: 10.1073/pnas.95.19.11163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12•.Lavery R, Zakrzewska K, Beveridge D, Bishop TC, Case DA, Cheatham T, III, Dixit S, Jayaram B, Lankas F, Laughton C, et al. A systematic molecular dynamics study of nearest-neighbor effects on base pair and base pair step conformations and fluctuations in B-DNA. Nucleic Acids Res. 2010;38:299–313. doi: 10.1093/nar/gkp834. Molecular dynamics simulations of 39 different DNA oligonucleotide duplexes provide new information on the predicted deformations of the 10 unique base-pair steps, including the effects of nearest-neighbors on the computed structures and the equilibria between conformational substates. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Trifonov EN, Sussman JL. The pitch of chromatin DNA is reflected in its nucleotide sequence. Proc Natl Acad Sci USA. 1980;77:3816–3820. doi: 10.1073/pnas.77.7.3816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Levene SD, Crothers DM. Ring closure probabilities for DNA fragments by Monte Carlo simulation. J Mol Biol. 1986;189:61–72. doi: 10.1016/0022-2836(86)90381-5. [DOI] [PubMed] [Google Scholar]
  • 15.Hagerman PJ, Ramadevi VA. Application of the method of phage T4 DNA ligase-catalyzed ring-closure to the study of DNA structure. I. Computational analysis. J Mol Biol. 1990;212:351–362. doi: 10.1016/0022-2836(90)90130-E. [DOI] [PubMed] [Google Scholar]
  • 16•.Wiggins PA, van der Heijden T, Moreno-Herrero F, Spakowitz A, Phillips R, Widom J, Dekker C, Nelson PC. High flexibility of DNA on short length scales probed by atomic force microscopy. Nature Nanotechnol. 2006;1:137–141. doi: 10.1038/nnano.2006.63. High-resolution atomic-force-microscopy images of single DNA molecules show that spontaneous large-angle local DNA bends occur much more frequently than expected from the harmonic bending energies assumed in classical elastic models of DNA. [DOI] [PubMed] [Google Scholar]
  • 17•.Demurtas D, Amzallag A, Rawdon EJ, Maddocks JH, Dubochet J, Stasiak A. Bending modes of DNA directly addressed by cryo-electron microscopy of DNA minicircles. Nucleic Acids Res. 2009;37:2882–2893. doi: 10.1093/nar/gkp137. Cryo-electron-microscopy images of 94-bp covalently closed and nicked DNA minicircles do not show sharp kinks but images of gapped molecules that contain highly flexible single-stranded regions show such kinks. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Lanka F, Lavery R, Maddocks JH. Kinking occurs during molecular dynamics simulations of small DNA minicircles. Structure. 2006;14:1527–1534. doi: 10.1016/j.str.2006.08.004. [DOI] [PubMed] [Google Scholar]
  • 19•.Harris SA, Laughton CA, Liverpool TB. Mapping the phase diagram of the writhe of DNA nanocircles using atomistic molecular dynamics simulations. Nucleic Acids Res. 2008;36:21–29. doi: 10.1093/nar/gkm891. Atomistic molecular-dynamics simulations of DNA minicircles as large as 178 bp show a distinct asymmetry in over- vs. underwinding responses, suggesting that long double-helical molecules are mechanically more complex than a simple elastic rod model would imply. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20•.Curuksu J, Zacharias M, Lavery R, Zakrzewska K. Local and global effects of strong DNA bending induced during molecular dynamics simulations. Nucleic Acids Res. 2009;37 doi: 10.1093/nar/gkp234. Molecular-dynamics studies of the controlled bending of 15-bp B-DNA oligomer duplexes produce local kinks at consecutive base pair steps, the free energies of which vary quadratically with the bending angle for moderate bends and increase linearly at greater deformations. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21•.Makde RD, England JR, Yennawar HP, Tan S. Structure of RCC1 chromatin factor bound to the nucleosome core particle. Nature. 2010;467:562–566. doi: 10.1038/nature09321. The 2.9-Å crystal complex of Xenopus histones, the 601 DNA positioning sequence, and the Drosophila RCC1 (regulator of chromosome condensation) protein shows how the chromatin protein interacts with the histone and DNA components of the nucleosome core particle. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22•.Vasudevan D, Chua EYD, Davey CA. Crystal structures of nucleosome core particles containing the ‘601’ strong positioning sequence. J Mol Biol. 2010;403:1–10. doi: 10.1016/j.jmb.2010.08.039. The crystal structures of the 601 DNA positioning sequence wrapped around Xenopus histones in two different asymmetric nucleosome core-particle structures point to the importance of site-specific DNA stretching in the assemblies. [DOI] [PubMed] [Google Scholar]
  • 23•.Wu B, Mohideen K, Vasudevan D, Davey CA. Structural insight into the sequence dependence of nucleosome positioning. Structure. 2010;18:528–536. doi: 10.1016/j.str.2010.01.015. The structures of nucleosome core particles containing strong TTTAA positioning elements on either side of the nucleosome center show the kink-and-slide deformations expected of the easily deformable TA step if the element is placed 0.5 double-helical turns from the nucleosome center but a smoother deformation when placed 1.5 turns from the center. [DOI] [PubMed] [Google Scholar]
  • 24.Lowary PT, Widom J. New DNA sequence rules for high affinity binding to histone octamer and sequence-directed nucleosome positioning. J Mol Biol. 1998;276:19–42. doi: 10.1006/jmbi.1997.1494. [DOI] [PubMed] [Google Scholar]
  • 25.Satchwell SC, Drew HR, Travers AA. Sequence periodicities in chicken nucleosome core DNA. J Mol Biol. 1986;191:659–675. doi: 10.1016/0022-2836(86)90452-3. [DOI] [PubMed] [Google Scholar]
  • 26.Segal E, Fondufe-Mittendorf Y, Chen L, Thastrom A, Field Y, Moore IK, Wang JP, Widom J. A genomic code for nucleosome positioning. Nature. 2006;442:772–778. doi: 10.1038/nature04979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Albert I, Mavrich TN, Tomsho LP, Qi J, Zanton SJ, Schuster SC, Pugh BF. Translational and rotational settings of H2A.Z nucleosomes across the Saccharomyces cerevisiae genome. Nature. 2007;446:572–576. doi: 10.1038/nature05632. [DOI] [PubMed] [Google Scholar]
  • 28•.Morozov AV, Fortney K, Gaykalova DA, Studitsky VM, Widom J, Siggia ED. Using DNA mechanics to predict in vitro nucleosome positions and formation energies. Nucleic Acids Res. 2009;37:4707–4722. doi: 10.1093/nar/gkp475. This paper focuses on a sequence-dependent elastic model of DNA deformability that accounts for the hydroxyl-radical mapping of nucleosomes on three DNA positioning sequences, also reported in the paper, and for earlier measurements of the free energies of nucleosome formation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ulyanov NV, Zhurkin VB. Sequence-dependent anisotropic flexibility of B-DNA. A conformational study. J Biomol Struct Dyn. 1984;2:361–385. doi: 10.1080/07391102.1984.10507573. [DOI] [PubMed] [Google Scholar]
  • 30.Davey CA, Sargent DF, Luger K, Mä der AW, Richmond TJ. Solvent mediated interactions in the structure of the nucleosome core particle at 1.9 Å resolution. J Mol Biol. 2002;319:1097–1113. doi: 10.1016/S0022-2836(02)00386-8. [DOI] [PubMed] [Google Scholar]
  • 31.Ong MS, Richmond TJ, Davey CA. DNA stretching and extreme kinking in the nucleosome core. J Mol Biol. 2007;368:1067–1074. doi: 10.1016/j.jmb.2007.02.062. [DOI] [PubMed] [Google Scholar]
  • 32.Harp JM, Hanson BL, Timm DE, Bunick GJ. Asymmetries in the nucleosome core particle at 2.5 Ångstrom resolution. Acta Crystallogr D. 2000;56:1513–1534. doi: 10.1107/s0907444900011847. [DOI] [PubMed] [Google Scholar]
  • 33.Tsunaka Y, Kajimura N, Tate S, Morikawa K. Alteration of the nucleosomal DNA path in the crystal structure of a human nucleosome core particle. Nucleic Acids Res. 2005;33:3424–3434. doi: 10.1093/nar/gki663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Yakovchuk P, Protozanova E, Frank-Kamenetskii MD. Base-stacking and base-pairing contributions into thermal stability of the DNA double helix. Nucleic Acids Res. 2006;34:564–574. doi: 10.1093/nar/gkj454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Lu X-J, Olson WK. Resolving the discrepancies among nucleic acid conformational analyses. J Mol Biol. 1999;285:1563–1575. doi: 10.1006/jmbi.1998.2390. [DOI] [PubMed] [Google Scholar]
  • 36•.Britton L, Olson WK, Tobias I. Two perspectives on the twist of DNA. J Chem Phys. 2009;131:245101. doi: 10.1063/1.3273453. This paper introduces a new quantity called the twist of supercoiling, different from the step parameter, called the twist, which is one of six rigid-body parameters used to describe the spatial arrangement of successive base pairs. Unlike the step-parameter twist, the twist of supercoiling gives an integral linking number when summed and combined with the writhing number of a closed DNA. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hayes JJ, Tullius TD, Wolffe AP. The structure of DNA in a nucleosome. Proc Natl Acad Sci USA. 1990;87:7405–7409. doi: 10.1073/pnas.87.19.7405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Shimada J, Yamakawa H. Ring-closure probabilities for twisted wormlike chains. Application to DNA. Macromol. 1984;17:689–698. [Google Scholar]
  • 39.Subirana JA, Faria T. Influence of sequence on the conformation of the B-DNA helix. Biophys J. 1997;73:333–338. doi: 10.1016/S0006-3495(97)78073-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Pérez A, Marchán I, Svozil D, Sponer J, Cheatham TE, III, Laughton CA, Orozco M. Refinement of the AMBER force field for nucleic acids: improving the description of / conformers. Biophys J. 2007;92:3817–3829. doi: 10.1529/biophysj.106.097782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Cornell WD, Cieplak P, Bayly CI, Gould IR, Merz KM, Ferguson DM, Spellmeyer DC, Fox T, Caldwell JW, Kollman PA. A second generation force field for the simulation of proteins, nucleic acids and organic molecules. J Am Chem Soc. 1995;117:5179–5197. [Google Scholar]
  • 42.Calladine CR. Mechanics of sequence-dependent stacking of bases in B-DNA. J Mol Biol. 1982;161:343–352. doi: 10.1016/0022-2836(82)90157-7. [DOI] [PubMed] [Google Scholar]
  • 43.van Dam L, Levitt MH. BII nucleotides in the B and C forms of natural-sequence polymeric DNA: a new model for the C forms of DNA. J Mol Biol. 2000;304:541–561. doi: 10.1006/jmbi.2000.4194. [DOI] [PubMed] [Google Scholar]
  • 44•.Trifonov EN. Cracking the chromatin code: precise rule of nucleosome positioning. Phys Life Rev. 2011;8:39–50. doi: 10.1016/j.plrev.2011.01.004. This retrospective review and the seven accompanying commentaries debate the physical basis of the packaging DNA in chromatin, with emphasis on the structural and energetic signals in the nucleotide sequence responsible for the precise positioning of the histone octamer on a DNA molecule. [DOI] [PubMed] [Google Scholar]
  • 45.Drew HR, Wing RM, Takano T, Broka C, Tanaka S, Itakura K, Dickerson RE. Structure of a B-DNA dodecamer: conformation and dynamics. Proc Natl Acad Sci USA. 1981;78:2179–2183. doi: 10.1073/pnas.78.4.2179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Holbrook SR, Dickerson RE, Kim S-H. Anisotropic thermal-parameter refinement of the DNA dodecamer CGCGAATTCGCG by the segmented rigid-body method. Acta Crystallogr B. 1985;41:255–262. [Google Scholar]
  • 47.Westhof E. Re-refinement of the B-dodecamer d(CGCGAATTCGCG) with a comparative analysis of the solvent in it and in the Z-hexamer d(5BrCG5BrCG5BrCG) J Biomol Struct Dyn. 1987;5:581–600. doi: 10.1080/07391102.1987.10506414. [DOI] [PubMed] [Google Scholar]
  • 48•.Heddi B, Oguey C, Lavelle C, Foloppe N, Hartmann B. Intrinsic flexibility of B-DNA: the experimental TRX scale. Nucleic Acids Res. 2010;38:1034–1047. doi: 10.1093/nar/gkp962. This paper reports the sequence-dependent propensity of the DNA backbone to convert in solution between the BI and BII conformational forms based on a large dataset of NMR 31P chemical shifts. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lu X-J, Olson WK. 3DNA: a software package for the analysis, rebuilding and visualization of three-dimensional nucleic acid structures. Nucleic Acids Res. 2003;31:5108–5121. doi: 10.1093/nar/gkg680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Lu X-J, Olson WK. 3DNA: a versatile, integrated software system for the analysis, rebuilding and visulization of three-dimensional nucleic-acid structures. Nature Protocols. 2008;37:1213–1227. doi: 10.1038/nprot.2008.104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Zhurkin VB, Ulyanov NB, Gorin AA, Jernigan RL. Static and statistical bending of DNA evaluated by Monte Carlo simulations. Proc Natl Acad Sci USA. 1991;88:7046–7050. doi: 10.1073/pnas.88.16.7046. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES