Flexibility of nucleic acids: From DNA to RNA
Bao Lei , Zhang Xi , Jin Lei , Tan Zhi-Jie †,
Department of Physics and Key Laboratory of Artificial Micro- and Nano-structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China

 

† Corresponding author. E-mail: zjtan@whu.edu.cn

Project supported by the National Basic Research Program of China (Grant No. 2011CB933600), the National Natural Science Foundation of China (Grant Nos. 11175132, 11575128, and 11374234), and the Program for New Century Excellent Talents, China (Grant No. NCET 08-0408).

Abstract
Abstract

The structural flexibility of nucleic acids plays a key role in many fundamental life processes, such as gene replication and expression, DNA-protein recognition, and gene regulation. To obtain a thorough understanding of nucleic acid flexibility, extensive studies have been performed using various experimental methods and theoretical models. In this review, we will introduce the progress that has been made in understanding the flexibility of nucleic acids including DNAs and RNAs, and will emphasize the experimental findings and the effects of salt, temperature, and sequence. Finally, we will discuss the major unanswered questions in understanding the flexibility of nucleic acids.

1. Introduction

Nucleic acids are negatively charged biopolymers and their structures are generally stabilized by base pairing/stacking interactions and metal ion-binding. [ 1 4 ] Due to the polymeric nature of nucleic acids and the stabilizing energy on the order of ∼ k B T (thermal energy), nucleic acids generally show strong conformational fluctuations and are rather flexible. The flexibility of nucleic acids is extremely important for their biological functions such as gene replication and expression, protein recognition, and gene regulation. [ 1 5 ]

To evaluate the flexibility of nucleic acids, various experimental methods have been developed, such as atomic force microscopy (AFM), [ 6 , 7 ] fluorescence resonance energy transfer (FRET), [ 8 , 9 ] small-angle x-ray scattering (SAXS), [ 9 11 ] magnetic tweezers (MT), [ 12 14 ] and optical tweezers (OT), [ 14 16 ] among others. Some theoretical models have also been developed and combined with experimental approaches to quantify the flexibility of nucleic acids, such as the worm-like chain (WLC) model and the free-joint chain model. [ 17 , 18 ] Recently, with the development of computational methods, molecular dynamics (MD) simulations [ 19 24 ] have been widely used to examine the flexibility of nucleic acids at the atomic level. Beyond the experiments that can only evaluate the macroscopic properties of flexibility, MD simulations and theoretical modeling at the atomic level can reveal detailed microscopic information, such as conformational changes of nucleic acids and ion binding patterns, as well as related microscopic mechanisms. [ 21 , 24 ] The use of various advanced experimental techniques and atomistic MD/modeling has greatly enhanced the understanding of nucleic acid flexibility. [ 6 22 ]

Due to base pairing/stacking energy on the order of ∼ k B T and the polyanionic nature of nucleic acids, [ 1 5 , 25 ] their flexibilities strongly depend on their sequences, salt ions in solution, and temperature, which would affect the strength of base pairing/stacking, ion binding, and chain conformational entropy, respectively. Drugs or proteins can also interact with nucleic acids and dramatically affect their structures and flexibilities. [ 2 , 3 , 5 , 26 ] In addition, nucleic acids can exhibit distinctively different flexibilities for different structural states depending on the temperature and ionic conditions. [ 27 31 ] Therefore, the nucleic acid flexibility is influenced by several critical factors, such as sequence, salt, and temperature.

In this review, we will focus on recent progress made in understanding the flexibility of nucleic acids. Since the family of nucleic acids includes single-stranded (ss) DNA/RNA, double-stranded (ds) DNA/RNA, and a large number of RNA tertiary folds, the main text is organized as follows. First, we will provide a brief overview of the flexibility of ssDNA/RNA. Second, we will focus on the flexibility of dsDNA, which has already attracted a great deal of attention over many years. Third, we will describe the recent progress in understanding the flexibility of RNAs. Finally, we will discuss the major unanswered questions in understanding the flexibility of nucleic acids.

2. Flexibility of ssDNA and ssRNA

The ss chain is an elementary structural and functional segment of nucleic acids. For example, RNA structures generally consist of different types of loops, and the ss chain also represents the denatured state of nucleic acids. [ 1 , 32 , 33 ] Furthermore, the ss chain is an important intermediate in many key biochemical processes, such as replication, recombination repair, and transcription, and is specifically recognized by many proteins. [ 32 , 34 ] The flexibility of the ss chain plays a significant role in its interactions with other macromolecules, such as proteins. [ 35 ] Generally, under the physiological conditions, the ss nucleic acid chains composed of generic sequences are rather flexible, and can be approximately described using the free-joint chain model, while the ss nucleic acid flexibility may be sensitive to the sequence and ionic environment. [ 36 52 ]

The flexibilities of ss nucleic acids have been quantified using various experimental approaches, such as force-extension curves, [ 36 40 ] FRET, [ 41 , 42 ] SAXS, [ 43 ] AFM, [ 44 ] and fluorescence recovery after photobleaching. [ 52 ] These experimental measurements are summarized in Table  1 . In addition, computer simulations, such as atomistic molecular and Monte Carlo simulations, have been employed to evaluate the flexibility of ss nucleic acids, including the effects of sequence and salt. [ 36 52 ]

Experimental measurements for the flexibility of single-stranded DNAs/RNAs.

ss nucleic acids References Ionic conditions Thermodynamic quantities
13k nt ssDNA Bosco et al . [ 36 ] 10–1000 mM Na + ; S , L , P
0.5–10 mM Mg 2+
20–3500 mM K + ;
McIntosh et al . [ 38 ] 20–2000 mM Na + ; P
0.2-50 mM Mg 2+ ;
0.2–50 mM Ca 2+
dT40, rU40 Chen et al . [ 41 ] 0–800 mM Na + ; R ee , L , P
0–100 mM Mg 2+
dT30 Meisburger et al . [ 42 ] 20 mM Na + ; R g , R ee
0–20 mM Mg 2+
dT8–dT100; dA8–dA50 Sim et al . [ 43 ] 12.5–1000 mM Na + R g , P
12–120 nt ss nucleic acids a) Wang et al . [ 47 ] 1–1000 mM Na + ; R ee , P
0.03–300 mM Mg 2+ ;
0.01-100 mM Co 3+
dT12, dT24, dA12, dA24 Mills et al . [ 50 ] 8–64 mM Na + ; P
0-8 mM Mg 2+
Poly(U) Seol et al . [ 51 ] 5–500 mM Na + S , P
280–5386 nt ssDNA Tinland et al . [ 52 ] 1–100 mM EDTA R g , P

This was a computational study that collected various experimental data for P of ss nucleic acids; R g : radius of gyration; R ee : end-to-end distance; L : contour length; P : bending persistence length; S : stretching modulus.

2.1. Sequence effect

Various experiments have suggested that the structure and flexibility of an ssDNA/RNA chain strongly depend on the intra-chain interactions, such as base-pairing and base stacking, which are highly correlated with the nucleic acid sequence. An ssDNA chain composed of a generic sequence can fold into a secondary structure, such as a hairpin or a helix through base-pairing/stacking; different sequence compositions give generic ss chains different properties, including flexibility. [ 36 , 40 , 45 ] Single-molecule experiments have suggested that an ssDNA can be modeled as the free-joint-chain model under physiological ionic conditions with a small persistence length of ∼ 7.5 Å, [ 49 ] since strong intra-chain base stacking is generally not observed in these molecules.

A homo-polynucleotide ssDNA may exhibit self-stacking interactions between nearby nucleotide bases, which can be sufficiently strong to maintain the rigidity of the ss chain. [ 44 ] The varying strengths of the intra-chain self-stacking interactions result in different flexibilities of the four types of homo-polynucleotide ssDNA molecules, including poly(dA), poly(dT), poly(dC), and poly(dG). Previous experiments have shown that poly(dT) is much more flexible than poly(dC), poly(dG), and poly(dA), since the self-stacking interactions in these molecules are much stronger than that in poly(dT). Actually, poly(dT) behaves as a free polyelectrolyte chain, [ 39 , 44 , 50 ] while poly(A), poly(G), and poly(C) would form ordered ss helices because of strong intra-chain self-stacking. [ 33 , 39 ]

Since thymine (T) and uridine (U) show similar strengths in base-stacking interactions, the effect of the sequence on the flexibility of ssRNA is very similar to that of DNA. [ 39 , 51 ] A previous study showed that the intra-chain self-stacking of ss-RNA is slightly stronger than that of ssDNA because of the difference in the sugar ring. Slightly lower flexibility for poly(rU) than for poly(dT) has been observed under the same ionic conditions. [ 47 ]

However, the details of self-stacking interactions in ssDNA/RNA and the mechanism of the flexibility of ss-DNA/RNA remain unclear. Advanced experimental techniques and improved force fields for atomistic simulations can be used to examine these issues.

2.2. Effects of salt and chain length

Due to the similar charge density on ssDNA and ssRNA, the ion dependence of ssRNA flexibility is very similar to that of ssDNA. [ 41 , 47 , 53 ]

Ions in solution can bind to ssDNA/RNA, which would increase nucleic acid flexibility by neutralizing the negative charge on the phosphates. Numerous experiments have shown that ssDNA/RNA becomes more flexible with increasing concentrations of ions, including Na + and Mg 2+ , and this ion-dependence of flexibility is stronger in longer sequences. [ 36 , 38 , 41 43 , 47 , 48 ] Mg 2+ has more ionic charge than Na + , and experiments and simulations have shown that Mg 2+ is approximately 60–120-fold more efficient than Na + in ionic neutralization. [ 41 , 47 ] Additionally, the ion concentration dependence of ssDNA/RNA flexibility is stronger for Na + than for Mg 2+ . [ 36 , 38 , 41 , 47 ] At high salt concentration, intra-sequence self-stacking interactions would dominate the global flexibility. [ 42 , 47 ] For ss generic sequences, empirical formulas have been derived for ion-dependent persistence length, which are practically useful. [ 43 , 47 , 48 ] For an ssDNA under force, a previously derived formula can describe the force-dependent ssDNA force-extension curves at various concentrations of NaCl. [ 54 ]

Previous studies have also shown that the ion-dependence of ssDNA/RNA flexibility depends on the sequence length, i.e., the persistence length P of longer ssDNA/RNAs increases more rapidly than that of short sequences when the ion concentration is decreased. [ 36 , 38 , 41 , 47 , 53 ] This occurs because ion binding to ssDNA/RNAs strongly depends on the sequence length. [ 36 , 38 , 41 43 , 47 , 52 ] Experiments have also shown that there is a scaling law between the size of ssDNA and its length, and the scaling exponent decreases with increasing monovalent salt concentration. [ 55 ]

However, the effect of multivalent ions on ssDNA/RNA flexibility remains unclear since the strong ion–ion correlation and possible intra-chain base-pairing/stacking can become tightly coupled.

3. Flexibility of dsDNA
3.1. General features in flexibility of dsDNA

Since the discovery of the dsDNA helix, numerous studies have demonstrated that dsDNAs are extensively involved in various life processes. [ 1 , 2 , 56 , 57 ] For example, dsDNA can fold into compact structures to enter into bacteriophage heads [ 56 ] or form chromatin, [ 57 ] and dsDNA can also sharply bend on a local scale to execute its biological functions, such as replication, DNA repair, and transcription, among other functions. [ 1 , 2 ] Therefore, understanding the biological processes related to dsDNAs requires the comprehensive understanding of dsDNA flexibility.

Experimental measurements for the flexibility of dsDNAs/dsRNAs.

dsDNAs or dsRNAs References Ionic conditions & temperature Thermodynamic quantities
4.2k bp & 8.3k bp dsRNA Abels et al . [ 7 ] moderate salt buffer P
16, 21, 66 & 89 bp dsDNA Yuan et al. [ 9 ] 500 mM Na + R g , P
10, 15, 20, 25, 30, 35 bp dsDNA Mathew-Fenn et al . [ 10 ] 100 mM Na + R ee ,
4.2k bp dsRNA & 3.4k bp dsDNA Lipfert et al. [ 14 ] 100 mM &320 mM Na + P , C , S , D
4k bp λ DNA & 4k bp dsRNA Herrero-Galán et al. [ 15 ] 0–500 mM Na + P , S
50k bp λ DNA Baumann et al. [ 28 ] 1.86–586 mM Na + ; P , S
Mg 2+ , Put 2+ , Spd 3+ ,
200 bp λ DNA Geggier et al. [ 31 ] TBE buffer; 5–60 °C P
14.8k bp dsDNA Bryant et al. [ 58 ] 100 mM Na + C
50k bp λ DNA Strick et al. [ 62 ] 10 mM phosphate buffer P , C
T7 DNA Sobel et al. [ 68 ] 5–3000 mM Na + R ee , P
6954±20 bp dsDNA Borochov [ 69 ] 7.3–4000 mM Na + R g , P
dsDNA (125bp-23000 bp) Mantelli et al. [ 73 ] 1 mM Mg 2+ ; 1–100 mM Na + P
3888 bp dsDNA Wang et al. [ 74 ] Na + , K + , Mg 2+ , Spd 3+ P , S , L
pBR322 dsDNA & Φ6 dsRNA Lang et al. [ 76 ] 48–500 mM NH 4 Cl L
685 bp dsDNA Driessen et al. [ 91 ] 60 mM K + ; P
100 & 150 mM Na + ;
23–52 °C
blunt-ended DNA fragments (41–256 bp) Porschke [ 92 ] 2.4–110 mM Na + ; P , R H
0.1 & 10 mM Mg 2+ ;
2–20 °C
∼ 200 bp dsDNA Geggier et al. [ 99 ] TBE buffer P
2743 bp dsDNA Wiggins et al. [ 115 ] 12 mM Mg 2+ bend angle

S : stretching modulus; P : bending persistence length; C : torsional persistence length; L : contour length; R ee : end-to-end distance; R g : radius of gyration; D : twist–stretch coupling parameter; : variance of R ee ; R H : hydrodynamic radius.

The dsDNA, as a highly dynamic structure, is stretchable, bendable, and twistable in vivo and in vitro , and its flexibility can be characterized by three important elastic parameters: stretching modulus S , bending persistence length P , and torsional persistence length C . The S , P , and C describe the stretching, bending, and twisting flexibilities, respectively. In addition, contour length L , end-to-end distance R ee , and radius of gyration R g have also been used to describe the global structural flexibility of dsDNAs. Extensive experiments have been conducted to quantify the flexibility of dsDNA, and the experimental measurements are summarized in Table  2 .

Previous reviews [ 1 , 2 ] have described the bending persistence length P of long dsDNA in buffers of moderate salt concentration, which was 45–50 nm based on early experiments. Recently, advanced single-molecule techniques have enabled accurate measurements of dsDNA flexibility. Herrero-Galán et al. manipulated long dsDNAs using magnetic tweezers and optical tweezers, and they observed a bending persistence length P of ∼ 49 ± 2 nm and a stretch modulus S of ∼ 935 ± 121 pN at 150 mM NaCl. [ 15 ] Other recent force-extension experiments for dsDNA indicated that torsional persistence length C was 100 ± 7 nm, [ 58 ] a slightly higher value than that generally accepted. [ 59 62 ] Very recently, Dekker et al . explored the elastic properties of dsDNA and derived all four elastic constants for dsDNA at 100 mM monovalent salt: 45 ± 2 nm for P , 1000 ± 200 pN for S , 109 ± 4 nm for C , and a negative twist–stretch coupling parameter of ∼ 17 ± 5. [ 14 ]

Since dsDNA is generally stabilized by specific base-pairing/stacking interactions and the binding of metal ions, the flexibility of dsDNA is strongly dependent on salt, sequence, and temperature. Next, we will comprehensively summarize the flexibility of dsDNA in the three aspects. Finally, we will introduce recent findings for short dsDNA and dsDNA under high force.

3.2. Salt effect

Due to the highly negative charge on dsDNA, the flexibility and stability of dsDNA are tightly coupled to the metal ions present in solution. [ 1 , 28 , 63 , 64 ] Numerous experimental and theoretical studies have focused on the role of salt in dsDNA flexibility and have revealed the following major features. [ 15 , 27 , 28 , 65 74 ]

However, on the salt effect on dsDNA flexibility, there are still important questions remaining. First, the strength of the salt dependence of P of dsDNA, particularly above ∼0.1 M monovalent salt, is unclear. Odijk, Skolnick, and Fixman have previously proposed the OSF theory, which divides the bending persistence length into intrinsic and electrostatic contributions. [ 77 , 78 ] The OSF theory predicts that the electrostatic part only contributes less than ∼ 10% to the total persistence length of dsDNA under physiological ionic conditions. Nevertheless, Manning developed a model based on his counter ion condensation theory by defining a hypothetical structure of dsDNA in the absence of DNA residual charges as a “null isomer of DNA.” [ 79 ] Manning’s model predicted that the dependence of P on salt concentration is nearly log-[salt] over the entire concentration range and the electrostatic contribution to the total bending persistence length can reach ∼86%, which has been supported by several experiments. [ 67 69 ] Savelyev et al . developed a two-bead coarse-grained structural model for dsDNA and conducted a reexamination by MD simulations. They found that both electrostatic and nonelectrostatic interactions play comparable roles in dsDNA flexibility. [ 80 ] They also found that the dsDNA bending persistence length decreases by ∼ 25% when the monovalent salt is decreased from 0.1 M to 1 M. [ 81 ] Therefore, there is still no consensus regarding how DNA flexibility depends on salt concentration, particularly at high salt concentrations of > 1 M.

Second, how multivalent ions influence the flexibility of dsDNA is still unclear since multivalent ions interact with DNA much more strongly than monovalent ions. [ 70 , 71 ] As indicated in experiments, multivalent ions, such as Mg 2+ and , have very different effects on DNA flexibility at the same ionic strength compared with monovalent ions. For example, can lead to a value of P for dsDNA as low as 25–30 nm. [ 28 ] Is the effect of multivalent ions coupled to ion-mediated effective interaction between segments in dsDNA? [ 82 , 83 ] Furthermore, what are the roles of divalent ions in the flexibility of dsDNA since some divalent ions can induce effective attractions between dsDNA helices, while other divalent ions cannot? [ 84 ]

In addition to the unsolved issues described above, additional questions remain at the microscopic level: (i) Why does the stretching modulus increase with increasing ion concentration and ion valence? (ii) Why are stretching and twisting negatively coupled?

3.3. Temperature effect

For a polymer, the WLC model describes the relationship between bending rigidity g and bending persistence length P as [ 85 ]

where k B is the Boltzmann constant and T is the absolute temperature. If g does not change with temperature, the temperature dependence of P can be described by Eq. ( 1 ), and P decreases linearly with increasing temperature. However, dsDNA is composed of sequential base pairs unlike an ideal polymer. Since the base pairing/stacking strength is on the order of ∼ k B T , temperature plays an important role in dsDNA flexibility. With increasing temperature, bases and backbone of dsDNA fluctuate more strongly, which may cause local “melted” bubbles. If temperature becomes sufficiently high, dsDNA strands can become completely separated and exhibit a “melted” state of the ss chain. Therefore, the actual temperature dependence of P should be much stronger than that predicted from Eq. ( 1 ).

Approximately 40 years ago, Gray and Hearst measured the sedimentation coefficient of DNA at an infinite dilution and obtained the temperature dependence of P . [ 29 ] However, their data showed that the temperature dependence of P is even weaker than the value predicted from Eq. ( 1 ), which assumes the temperature-independent bending rigidity g . Recently, dsDNA cyclization experiments have indicated that local melting in dsDNA can enhance the flexibility of ds-DNA more significantly than that predicted from the WLC model in the temperature range of 23–42 °C. [ 30 ] A significant decrease in apparent bending persistence length may result from potential excitations of flexible defects, [ 86 , 87 ] which can be excited by high temperature. [ 88 ] In order to more accurately measure the temperature dependence of persistence length, Geggier and Vologodskii employed two different approaches to determine the persistence lengths of dsDNA at different temperatures, and the data from the two independent approaches were highly consistent. The experiments showed that the bending persistence length decreased nearly linearly from 53 nm to 44 nm as temperature decreased from 5 °C to 42 °C, but that it decreased very sharply at higher temperatures, reaching ∼36 nm at 60 °C. [ 31 ] Corresponding theoretical modeling based on the Peyrard–Bishop–Dauxois model has also shown good agreement with the experimental data. Therefore, the experiments by Vologodskii et al . and the Peyrard–Bishop–Dauxois-based model both showed discrepancies with the WLC model. [ 89 , 90 ] Very recent experiments with tethered particle motion by Driessen et al . confirmed the above findings, and also showed that the increase of temperature can lead to a more compact structure of dsDNA and the temperature-dependent P is tightly coupled to the content of GC base pairs. [ 91 ]

It is understandable that dsDNA is more flexible at higher temperature because the ds helix is generally stabilized by base pairing/stacking interactions. However, a previous experiment showed that dsDNA flexibility can become weaker with increasing temperature in [2 °C, 20 °C], which was proposed to be attributed to the sequence-direct DNA curvature. [ 92 ] Such temperature-weakened flexibility for some dsDNA requires further analysis.

3.4. Sequence effect

It is well known that dsDNA flexibility depends on its sequence, which directly affects dsDNA stability. [ 1 3 , 93 , 94 ] Numerous previous studies have shown that different sequence arrangements can greatly influence the stability of dsDNA under bending and its ability to form kinks, which can induce base pair slide to form non-native contacts. [ 1 3 ] Simultaneously, the sequence can also influence the mechanical properties of dsDNA in contact with proteins such as histones. [ 1 3 ]

Olson et al. performed the statistical analysis on x-ray crystal structures and found that different sequences produce distinct flexibility, where the AA·TT step belongs to the rigid class while GG·CC and GC·GC dinucleotides are even more flexible. [ 93 ] However, this conclusion may be invalid because of the choice of system in which the central regions were rich in AT content while the GC steps were segregated at the terminals. [ 95 ] Furthermore, based on the statistical analysis of x-ray crystal structures of protein-oligonucleotide complexes, Olson et al . found that the average twisting of base pair steps increases on the same order within three standard chemical classes: pyrimidine–purine, purine–purine, and purine–pyrimidine. [ 94 ] Ortiz and Pablo proposed a coarse-grained model for the effect of sequence on the overall stability and flexibility of dsDNA under bending constraints. [ 96 ] They found that longer repeated segments such as AAAAAAAA are more likely to form a kink, while short repetitive segments such as CCC are less likely to form a kink. This is because a base in the AA···AA strands can slide more easily to form non-native contacts with neighboring complementary bases in the repeated TT···TT sequence. [ 96 ]

In addition to the above theoretical approaches, scanning force microscopy and AFM have been widely employed to characterize the flexibility of DNA. [ 95 , 97 , 98 ] Scipioni et al. described the intrinsic curvature of DNAs based on scanning force microscopy images, confirming that A·T-rich sequences are more flexible than G·C-rich sequences. [ 95 ] To obtain a more detailed understanding of the sequence-dependent flexibility of dsDNA, Geggier and Vologodskii determined the bending persistence length more accurately based on a cyclization method of short DNA fragments, and their data showed that variations in P induced by different sequences might be sufficiently large to affect their biological activity. Such variations can also affect the binding affinity of DNA-protein complexes, in which the dsDNA segment shows sharp bending. [ 99 ] Moreover, extensive theoretical and experimental analyses showed that DNA fragments containing A-tracts exhibit a tighter bend at the 3′ end than at the 5′ end, where the A-tract refers to the duplex (dA) n ·(dT) n and is equivalent to the T-tract. [ 3 , 96 , 100 ] TA and AG·CT steps show higher roll angle values compared to GC and GG·CC steps, and can also bend DNA even in the absence of the A-tract. [ 100 ] In addition to the specific sequence described above, the sequence GGGCCC also shows a net bend. [ 101 ]

With the rapid development of computational facility, molecular simulation has become an important tool for exploring the effect of sequence on the flexibility of dsDNA at the atomic level. [ 22 , 102 ] Based on an MD study of dsDNA in the gas phase, Xiao and Liang found that all pyrimidine rings are highly flexible in either isolated or paired states, whereas the imidazole rings are relatively more rigid. [ 22 ] Lavery et al. also performed systematic MD simulations to study the nearest-neighbor effects on base pairing. [ 102 ] Their simulations suggested that to predict the sequence dependence of DNA structure and dynamics, next-nearest-neighbor interactions should be taken into account because the effect is significant. [ 102 ]

3.5. DsDNA under high force

As described above, the elastic properties of dsDNA, such as stretching modulus and bending persistence length, can be investigated through single-molecule force–extension experiments by fitting extension–force curves to the (extensible) WLC model. [ 49 , 54 , 103 109 ] Under low stretching force (e.g., < 20 pN), dsDNA generally maintains the B-form, and the measured properties reflect the intrinsic elasticity of B-DNA with specific sequences under specific environmental conditions. However, under high stretching force (> 65 pN, near physiological ionic conditions), apparent overstretching transitions can occur in the dsDNA structure. [ 104 108 ] Such overstretching transitions are based on the early single-molecule experiments of dsDNA and have been proposed to involve a stretching-induced melting transition from dsDNA to ssDNA. [ 28 , 49 , 110 112 ] However, very recent single-molecule experiments have revealed that much more complex overstretching transitions from dsDNA to other overstretched structures occur, such as peeled ssDNA with one strand peeled from another, DNA bubble with two strands separated internally, and S-DNA with elongated base pairs. [ 105 108 ] These experimental findings were also suggested using early theoretical models. [ 112 ] These recent experiments have yielded the following major findings. [ 54 , 104 108 ]

Extensive single-molecule experiments have shown that the overstretching transition is directly coupled to the stability of a dsDNA, which strongly depends on the content of GC base pairs, ionic conditions, and end-constraints. Thus, it is expected that at lower temperature and in multivalent ion solutions, a dsDNA can be stretched to an S-DNA at lower ionic strength because of the higher stability. The atomic structures of overstretched dsDNA can be modeled by all-atom MD simulations. [ 24 ]

3.6. Flexibility of short DNA

Numerous recent experiments have suggested that short DNAs have higher flexibility compared with those of kilo-base pairs. [ 113 117 ] Cyclization experiments have shown that short dsDNAs of ∼ 100 base pairs (bps) form circles much faster than predicted by the WLC model mainly because of the unusually large local bend angle induced by kinking. [ 113 , 114 ] Another series of experiments using FRET and SAXS by Yuan et al. also suggested the higher flexibility of short DNAs of 15–89 bps, which are beyond the description of the conventional WLC model. [ 9 ] In addition, recent SAXS experiments of short DNAs of ≤ 35 bps with two end gold nanocrystals by Mathew-Fenn et al. suggested that short DNAs are at least one order of magnitude more extensible than long DNAs of kilo-bps revealed by previous single-molecule stretching experiments. [ 118 ] This high flexibility of short dsDNA has been proposed to be attributed to defect excitation, which may reduce the local bending energy of dsDNA through local DNA melting of a few base pairs, local DNA kinking, and excitation of a few base pairs of S-DNA. [ 86 , 112 , 115 , 119 , 120 ] Since the local stability of dsDNA depends strongly on sequence, ionic strength, and temperature, [ 2 , 30 , 31 , 63 , 64 , 93 , 94 ] such defect excitation may be sensitive to temperature and ionic strength. [ 112 ]

However, a similar SAXS experiment showed that the flexibility of short DNAs of 42–94 bps with two linked gold nanocrystals could be described by the WLC model with a persistence length of ∼ 50 nm. [ 11 ] On the illusive controversy, the atomic MDs have also been employed to probe the flexibility of short DNAs of 5–50 bps, and the results showed that shorter DNA may have higher apparent flexibility, which is attributed to the higher flexibility of ∼ 6 bps at each end. [ 88 ] Nevertheless, to explain the experiments with labeling nanocrystals developed by Mathew-Fenn et al . and Yuan et al . at the atomic level is still required, which would assist not only the understanding of the experimental findings, but also the understanding of the effect of labeling nanocrystals on the flexibility of short biomolecules.

4. Flexibility of RNA
4.1. Flexibility of dsRNA

Recently, dsRNA has been highly valued because its role in the life cycle of a cell is more crucial than previously considered. [ 7 , 14 , 15 , 121 123 ] In addition to being a central role in RNA interference, [ 121 ] dsRNA may have the potential applications in nanomedicine and nanomaterial. [ 122 , 123 ] Because of these important applications, the flexibility of dsRNA has been studied in various single-molecule experiments, [ 7 , 14 , 15 ] and the experimental measurements are tabulated in Table  2 .

Unlike dsDNA, which is generally in the B-form, dsRNA forms a thicker right-handed duplex in the “A-form”, [ 124 ] and this special helical structure gives dsRNA different flexibilities. In an early experiment of transient electric birefringence (TEB), Hagerman et al. obtained a bending persistence length of ∼ 60±10 nm for dsRNA, [ 125 , 126 ] which is 20%–30% larger than the accepted value for dsDNA. Therefore, dsRNA is somewhat stiffer than dsDNA. Recently, Dekker et al . obtained a mean bending persistence length of 63.8 ± 0.7 nm through force–extension measurements with magnetic tweezers and a value of 62 ± 2 nm using AFM measurements for long dsRNAs. [ 7 ] They also obtained a torsional persistence length of ∼ 99 ± 5 nm at an external stretching force F = 6.5 pN, which is similar to the value for dsDNA, as well as a stretching modulus of 350 ± 100 pN, which is three-fold lower than that of dsDNA. Interestingly, they observed that dsRNA exhibits positive twist–stretch coupling, which is in contrast to dsDNA with negative twist–stretch coupling. [ 14 ] Herrero-Galán et al . systematically measured the mechanical properties of dsRNA under different ion conditions at the single-molecule level. [ 15 ] They found that the bending persistence lengths P for dsRNA are consistently larger than those for dsDNA under the same ionic conditions, and P decreases with increasing salt concentration over the range of 0–500 mM NaCl, which is similar to dsDNA. Furthermore, the OT measurements revealed that the stretching modulus S of dsRNA increases with increasing salt concentration, in agreement with the trend for dsDNA, [ 28 ] while S of dsRNA is much lower than that of dsDNA under the same salt conditions. [ 16 , 28 ]

Therefore, compared with dsDNA, dsRNA has a larger bending persistence length and a similar torsional persistence length. However, the stretching modulus of dsRNA is nearly three-fold lower than that of dsDNA, and a surprising difference between dsRNA and dsDNA is that dsRNA has a positive twist–stretch coupling parameter while dsDNA has a negative one. The microscopic mechanism for the apparent difference in the flexibility between dsRNA and dsDNA remains unclear and requires further investigation.

4.2. Flexibility of structured RNA

Generally, RNAs fold into more complex native structures rather than keeping a denatured ss chain or a perfect duplex. [ 127 , 128 ] The flexibility of complex RNA structures is important for their biological functions, such as RNA-protein recognition and gene regulation. [ 129 , 130 ] Understanding the flexibility of structural RNAs would enable the deep exploration of their biological functions and related applications such as structure-based drug design. [ 131 , 132 ] However, there have been few extensive studies on examining the flexibility of structural RNAs. In Table  3 , we summarized the existing experimental measurements for the elastic properties of structured RNAs beyond the states of the ss chain and helix. [ 133 143 ]

For RNA hairpins with a bulge loop (and an internal loop), Zacharias and Hagerman performed a series of TEB experiments. [ 134 , 137 , 138 ] The bending angles at the junction induced by bulge loops of different sizes and base compositions were determined, and were found to increase monotonically with the increment from ∼ 8° to ∼ 20° for both bulge loops of A n and U n as n was increased from 1 to 6 in the absence of Mg 2+ . Moreover, for bulge loops constituted by U n , the presence of Mg 2+ reduced the increment of the bending angle by a factor of 2 for all n values. [ 137 ] However, for RNA hairpins with symmetric internal loops with the forms A n –A n and U n –U n ( n = 2, 4, 6), it was found that the internal loops could only distort RNAs by values that were much smaller than their bulge loop counterparts. [ 138 ]

Other experimental methods have also been employed to unravel the flexibility of structured RNAs. Al-Hashimi et al . performed a series of studies on the HIV-1 TAR RNA with NMR and residual dipolar coupling (RDC), in combination with coarse-grained modeling by Brooks et al. [ 139 , 144 ] Because of the high resolution of the method, detailed dynamic motions inside RNAs can be captured and some macroscopic quantities such as bend angle can be measured. Their experiments showed that the bending angle of HIV-1 TAR RNA decreases with increasing NaCl concentration, which was also predicted by recent coarse-grained models. [ 140 , 141 , 144 , 145 ] Thirumalai et al. developed and employed an empirical formula for the WLC model to describe the flexibility of the Azoarcus ribozyme at different monovalent and divalent ion concentrations. [ 142 ] Additionally, the corresponding persistence lengths were derived by fitting the empirical formula for the WLC model to the measured experimental data. [ 142 ]

Rather than examining the local details of non-helical RNA elements, Fulle and Gohlke studied whole RNA molecules directly. They first modeled an RNA structure as a topological network representation using a constraint counting method, and then ran a simulation with a framework rigid optimized dynamics algorithm (FRODA). The root-mean-square fluctuations of all atoms can be determined after simulations, and then the flexibility can be estimated both locally and globally. Quantitative comparisons between FRODA simulations and NMR experiments showed good agreement for some RNAs, including tRNA and pseudoknots. [ 131 ] Furthermore, the flexibility of the ribosomal exit tunnel, a large RNA-protein complex, has been studied using this method. [ 143 ] Very recently, a python-based software package named constraint network analysis (CNA) was developed by Gohlke et al . [ 146 ] Using the CNA algorithm, one can obtain both global and local properties for the input biomolecules, including RNAs. However, such constraint counting in the CNA algorithm may require more physical optimization and extensive validation.

Different methods employed to probe the flexibilities of structural RNAs.

RNAs References Methods Thermodynamic quantities
tRNA (Asp) Fulle et al . [ 131 ] constraint counting RMSD
network & FRODA
simulation
tRNA (Phe) Roh et al. [ 133 ] quasielastic neutron R g , P
scattering spectroscopy
HIV-1 TAR RNA Zacharias et al . [ 134 ] gel electrophoresis & TEB bend angle
subsequence of an sRNA (DsrA) De Almeida Ribeiro E et al . [ 135 ] SAXS & NMR & R g
ensemble optimization method
bacterial ribosomal A-site RNA Fulle et al . [ 136 ] normal mode analysis & MD simulation binding free energies, RMSD
segments of bulge loops Zacharias et al . [ 137 ] TEB bend angle
segments of symmetric internal loops Zacharias et al . [ 138 ] TEB bend angle
HIV-1 TAR RNA Al-Hashimi et al . [ 139 ] NMR & RDC & MD RMSD
HIV-1 TAR RNA Al-Hashimi et al . [ 140 , 141 ] NMR & RDC RMSD
Azoarcus ribozyme & RNase P Caliskan et al. [ 142 ] SAXS & WLC P
ribosomal exit tunnel Fulle et al. [ 143 ] constraint counting RMSD
network & FRODA simulation

P : bending persistence length; R g : radius of gyration; RMSD: root-mean-square deviation.

Since structured RNAs differ from DNA and RNA helices whose flexibility can be well quantified by persistence length, stretching modulus, and twisting modulus, characterizing the flexibility of structured RNAs in a straightforward and quantitative manner remains unclear.

5. Conclusion and perspective

As described above, extensive experiments and theoretical modeling have revealed that the flexibility of nucleic acids is tightly coupled to several critical factors. First, the states of structures can dominate the flexibility of nucleic acids, such as the states of the ss chain, ds helices, partially melted helices, and more complicated tertiary folds, corresponding to significantly different flexibilities. Second, the sequences of nucleic acids determine their structures and stabilities, and thus strongly influence their flexibility. Third, temperature can directly determine the state of structures of nucleic acids and thus greatly influence their flexibility. Additionally, solution conditions such as metal ions, which can strongly interact with nucleic acids, significantly affect nucleic acid flexibility, particularly multivalent ions. Recent developments in computation facility and molecular force fields have enabled extensive explorations of the flexibility of nucleic acids at the atomic level. However, despite this great progress in understanding the flexibility of nucleic acids, many important elusive problems must be explored. We will discuss several major challenging issues in the following.

5.1. Flexibility of DNA on a short length scale

A recent AFM experiment showed that spontaneous large-angle bends are many times more prevalent than predicted by the WLC model, [ 115 ] suggesting that dsDNA may have much higher flexibility on the short length scale. [ 147 ] To examine the mechanism for such higher flexibility on the short length scale, atomic-level MD simulations have been employed for two short dsDNAs, and the calculated apparent persistence length can be ∼ 20 nm on a very short scale (∼ 1–2 bp), and exhibit the oscillation periodically between 20 nm and 100 nm. [ 148 ] A correlated WLC model has also been developed to explain the experimental findings, while the microscopic mechanism for such proposed correlation remains unclear. [ 149 ] Very recently, high-resolution AFM in solution has been used to analyze the effect on a short length scale, which showed that dsDNA can be well described by the WLC model on a scale beyond 2–3 helical turns. [ 116 , 117 ] However, on a length scale below the threshold of 2–3 helical turns, quantification of the flexibility remains limited by the limitations of AFM. Furthermore, how the local kinking and disruption of hydrogen bond in base pairing affect the flexibility of dsDNA is still unknown. Additionally, the effect of substrate in the AFM experiments on the flexibility of dsDNA must be examined, since a theoretical modeling indicated that Mg 2+ -mediated attraction between DNA and substrate can cause DNA softening. [ 120 ] On a short length scale, atomistic MD can become a powerful tool, while reliable force fields in MD are essentially required.

5.2. Flexibility of helices of different conformations

Two typical conformations of nucleic acids include B-form DNA (B-DNA) and A-form RNA (A-RNA). The dsDNA and dsRNA are generally present in B-form and A-form, respectively. B-DNA and A-RNA helices show similar helical structures and some of their elastic properties are qualitatively similar, such as bending persistence length and torsional modulus. [ 14 , 15 , 27 ] However, B-DNA and A-RNA are significantly different in some elastic properties, such as stretching modulus and twist–stretch coupling. The stretching modulus of B-DNA is ∼ 3 times higher than that of A-RNA. [ 14 ] More surprisingly, single-molecule stretching experiments showed that the twist–stretch coupling parameter of B-DNA was negative, i.e., the DNA stretched by pulling force (4–8 pN) was accompanied by overwinding of the helix, while that of A-RNA was positive. However, very recently, Manning analyzed existing experimental data and concluded that DNA stretching by environmental change such as the decrease of salt concentration is accompanied by helix unwinding. [ 72 ] Therefore, several unanswered questions remain: (i) Why are B-DNA and A-RNA different in stretching modulus and twist–stretch coupling? (ii) How can we unify the results involving the pulling of B-DNA and the analyses based on experiments involving free B-DNA? (iii) Is the flexibility of other conformations of nucleic acid helices, such as A-DNA and Z-DNA, also different from the flexibility of B-DNA? Further studies, particularly those on the microscopic level, are required to answer these questions.

5.3. Effect of high salt and multivalent salt

Numerous experiments have shown that the persistence length of dsDNA decreases and the stretching modulus increases with increasing salt concentration. However, the available data regarding salt-dependent P can be categorized in two ways: (i) P will not continue to decrease after NaCl exceeds 0.1 M; [ 28 , 66 ] (ii) P will continue to decrease after NaCl exceeds 0.1 M. [ 68 , 69 ] The classic OSF theory supports the former, while recent coarse-grained simulations and theoretical analysis based on the counterion condensation theory agree with the latter. Therefore, for the salt-dependent flexibility of DNA, several questions remain unanswered. (i) To what extent does salt in the solution influence the flexibility of DNA, and what are the relative fractions of electrostatic and intrinsic contributions to the global flexibility of DNA? (ii) Why does P of DNA continue to decrease or become nearly invariant after NaCl exceeds 0.1 M? (iii) Why does the stretching modulus of DNA increase at higher salt concentration? (iv) Further studies are required to understand the salt dependence of other elastic properties such as torsional modulus and twist–stretch coupling. A series of experiments in combination with molecular modeling at the atomic level is required to resolve these issues.

The limited experiments on the flexibility of DNA in multivalent salt have demonstrated the dramatic roles of multivalent ions. Multivalent ions can cause an apparent decrease in the persistence length and an apparent increase in the stretching modulus of dsDNA compared with monovalent salt. [ 28 ] Such decrease in P has been attributed to a multivalent ion-mediated intra-chain attractive force, [ 82 , 83 , 150 , 151 ] while the increase in S has not been thoroughly explained. Therefore, to systematically quantify the effect of multivalent ions such as Mg 2+ and on DNA flexibility is still required, particularly on the local deformation of the DNA helix induced by multivalent ions. In addition, the effect of multivalent ions on the flexibility of A-RNA has not been widely examined, while previous experiments have shown that ions can bind to an A-RNA in a very different manner to B-DNA. [ 152 , 153 ] This suggests that the (multivalent) ion effect on A-RNA flexibility may be significantly different from that of B-DNA and thus is highly desirable.

5.4. Flexibility of RNA tertiary folds

The flexibility of a DNA or RNA helix can be well described by the parameters of the elastic theory of linear polymers or elastic rods. However, RNAs are generally in the folded state of complex native structures beyond a perfect helix, and thus to quantitatively describe the flexibility of non-helix RNAs is beyond the description of the elastic parameters for a helix. The bending angle, persistence length, the distribution of radius of gyration, and the distribution of root of mean square deviation (RMSD) have been used to characterize the flexibility of non-helix RNAs. Nevertheless, there are limitations to these methods. For example, the bending angle only works well for the local bending of a helix and two-way junction. The use of persistence length for structured RNAs (e.g., tRNA) is not feasible, and may only be considered as a relative quantity compared with the denatured ss state or secondary state. The distribution of the radius of gyration can be easily measured, but this value only gives a global description of the flexibility. Since a structured RNA is non-uniform in its flexibility over the entire molecule, the distribution of the radius of gyration may be inadequate for completely describing flexibility. The distribution of RMSD can describe the dynamics of all atoms and can be tracked by experiments and atomic-level modeling, while this description may be only convenient for small RNAs. A combination of these quantities may provide a thorough description of the flexibility of an RNA tertiary structure.

Since RNA tertiary structures are more sensitive to temperature and ionic conditions than helices, [ 154 158 ] the flexibilities of RNA tertiary structures are more strongly coupled to temperature and ionic conditions, particularly multivalent ions. Furthermore, dehydrated Mg 2+ and small molecules such as metabolites can interact specifically with RNAs and alter the flexibility of RNAs to aid their functions. [ 159 162 ] Therefore, understanding the flexibility of RNA tertiary structures is not well understood and requires further investigation.

In summary, the results of previous studies have greatly enhanced the understanding of nucleic acid flexibility, but many questions remain and require further comprehensive investigation. In the next decade, we expect more surprising findings regarding the flexibility of nucleic acids as well as more related applications.

Reference
1 Hagerman P J 1988 Annu. Rev. Biophys. Biophys. Chem. 17 265
2 Peters J P Maher L J 2010 Q. Rev. Biophys. 43 23
3 Hagerman P J 1990 Annu. Rev. Biochem. 59 755
4 Tan Z J Chen S J 2011 Met. Ions Life Sci. 9 101
5 Lebrun A Shakked Z Lavery R 1997 Proc. Natl. Acad. Sci. USA 94 2993
6 Mazur A K Maaloum M 2014 Phys. Rev. Lett. 112 068104
7 Abels J A Moreno-Herrero F Van der Heijden T Dekker C Dekker N H 2005 Biophys. J. 88 2737
8 Kang J Jung J Kim S K 2014 Biophys. Chem. 195 49
9 Yuan C Chen H Lou X W Archer L A 2008 Phys. Rev. Lett. 100 018102
10 Mathew-Fenn R S Das R Harbury P A 2008 Science 322 446
11 Mastroianni A J Sivak D A Geissler P L Alivisatos A P 2009 Biophys. J. 97 1408
12 Wang X L Zhang X H Cao M Zheng H Z Xiao B Wang Y Li M 2009 J. Phys. Chem. B 113 2328
13 Li W Wang P Y Yan J Li M 2012 Phys. Rev. Lett. 109 218102
14 Lipfert J Skinner G M Keegstra J M Hensgens T Jager T Dulin D Köbera M Yu Z B Donkers S P Chou F C Das R Dekker N H 2014 Proc. Natl. Acad. Sci. USA 111 15408
15 Herrero-Galán E Fuentes-Perez M E Carrasco C Valpuesta J M Carrascosa J L Moreno-Herrero F Arias-Gonzalez J R 2013 J. Am. Chem. Soc. 135 122
16 Lipfert J Kerssemakers J W Jager T Dekker N H 2010 Nat. Methods 7 977
17 Marko J F Siggia E D 1995 Macromolecules 28 8759
18 Toan N M Thirumalai D 2012 J. Chem. Phys. 136 235103
19 Song T J Liang H J 2012 J. Am. Chem. Soc. 134 10803
20 Zhang Y J Zhang J Wang W 2011 J. Am. Chem. Soc. 133 6882
21 Xiao S Y Zhu H Wang L Liang H J 2014 Soft Matter 10 1045
22 Xiao S Y Liang H J 2012 J. Chem. Phys. 136 205102
23 Zhao Y J Huang Y Y Gong Z Wang Y J Man J F Xiao Y 2012 Sci. Rep. 2 734
24 Qi W P Lei X L Fang H P 2010 ChemPhysChem 11 2146
25 Tan Z J Chen S J 2008 Biophys. J. 94 3137
26 Hou X M Zhang X H Wei K J Ji C Dou S X Wang W C Li M Wang P Y 2009 Nucleic Acids Res. 37 1400
27 Wenner J R Williams M C Rouzina I Bloomfield V A 2002 Biophys. J. 82 3160
28 Baumann C G Smith S B Bloomfield V A Bustamante C 1997 Proc. Natl. Acad. Sci. USA 94 6185
29 Gray H B Hearst J E 1968 J. Mol. Biol. 35 111
30 Forties R A Bundschuh R Poirier M G 2009 Nucleic Acids Res. 37 4580
31 Geggier S Kotlyar A Vologodskii A 2011 Nucleic Acids Res. 39 1419
32 Hannon G J 2002 Nature 418 244
33 Saenger W 1984 Principles of Nucleic Acid Structure New York Springer-Verlag
34 Diebold S S Kaisho T Hemmi H Akira S e Sousa C R 2004 Science 303 1529
35 Sung P Robberson D L 1995 Cell 82 453
36 Bosco A Camunas-Soler J Ritort F 2013 Nucleic Acids Res. 42 2064
37 Landy J McIntosh D B Saleh O A 2012 Phys. Rev. Lett. 109 048301
38 McIntosh D B Saleh O A 2011 Macromolecules 44 2328
39 Seol Y Skinner G M Visscher K Buhot A Halperin A 2007 Phys. Rev. Lett. 98 158103
40 Zhang Y Zhou H J Ou-Yang Z C 2001 Biophys. J. 81 1133
41 Chen H Meisburger S P Pabit S A Sutton J L Webb W W Pollack L 2012 Proc. Natl. Acad. Sci. USA 109 799
42 Meisburger S P Sutton J L Chen H Pabit S A Kirmizialtin S Elber R Pollack L 2013 Biopolymers 99 1032
43 Sim A Y Lipfert J Herschlag D Doniach S 2012 Phys. Rev. E 86 021901
44 Ke C Humeniuk M Hanna S Marszalek P E 2007 Phys. Rev. Lett. 99 018302
45 Toan N M Thirumalai D 2012 J. Chem. Phys. 136 235103
46 Sing C E Alexander-Katz A 2011 Macromolecules 44 6962
47 Wang F H Wu Y Y Tan Z J 2013 Biopolymers 99 370
48 Dessinges M N Maier B Zhang Y Peliti M Bensimon D Croquette V 2002 Phys. Rev. Lett. 89 248102
49 Smith S B Cui Y Bustamante C 1996 Science 271 795
50 Mills J B Vacano E Hagerman P J 1999 J. Mol. Biol. 285 245
51 Seol Y Skinner G M Visscher K 2004 Phys. Rev. Lett. 93 118102
52 Tinland B Pluen A Sturm J Weill G 1997 Macromolecules 30 5763
53 Tan Z J Chen S J 2008 Biophys. J. 95 738
54 Cocco S Yan J Léger J F Chatenay D Marko J F 2004 Phys. Rev. E 70 011910
55 Sim A Y Lipfert J Herschlag D Doniach S 2012 Phys. Rev. E 86 021901
56 Riemer S C Bloomfield V A 1978 Biopolymers 17 785
57 Richmond T J Finch J T Rushton B Rhodes D Klug A 1984 Nature 311 532
58 Bryant Z Stone M D Gore J Smith S B Cozzarelli N R Bustamante C 2003 Nature 424 338
59 Horowitz D S Wang J C 1984 J. Mol. Biol. 173 75
60 Shore D Baldwin R L 1983 J. Mol. Biol. 170 983
61 Vologodskii A V Marko J F 1997 Biophys. J. 73 123
62 Strick T R Bensimon D Croquette V 1999 Genetica 106 57
63 Tan Z J Chen S J 2006 Biophys. J. 90 1175
64 Tan Z J Chen S J 2007 Biophys. J. 92 3615
65 Peters J P Mogil L S McCauley M J Williams M C Maher L J 2014 Biophys. J. 107 448
66 Maret G Weill G 1983 Biopolymers 22 2727
67 Podesta A Indrieri M Brogioli D Manning G S Milani P Guerra R Finzi L Dunlap D 2005 Biophys. J. 89 2558
68 Sobel E S Harpst J A 1991 Biopolymers 31 1559
69 Borochov N Eisenberg H Kam Z 1981 Biopolymers 20 231
70 Tan Z J Chen S J 2005 J. Chem. Phys. 122 044903
71 Tan Z J Chen S J 2009 Methods Enzymol. 469 465
72 Manning G S 2015 Biopolymers 103 223
73 Mantelli S Muller P Harlepp S Maaloum M 2011 Soft Matter 7 3412
74 Wang M D Yin H Landick R Gelles J Block S M 1997 Biophys. J. 72 1335
75 Anderson P Bauer W 1978 Biochemistry 17 594
76 Lang D Steely H T Kao C Y Ktistakis N T 1987 Biochim. Biophys. Acta. 910 271
77 Odijk T 1977 J. Polym. Sci. Polym. Phys. Ed. 15 477
78 Skolnick J Fixman M 1977 Macromolecules 10 944
79 Manning G S 2006 Biophys. J. 91 3607
80 Savelyev A Materese C K Papoian G A 2011 J. Am. Chem. Soc. 133 19290
81 Savelyev A 2012 Phys. Chem. Chem. Phys. 14 2250
82 Tan Z J Chen S J 2006 Nucleic Acids Res. 34 6629
83 Tan Z J Chen S J 2012 Biophys. J. 103 827
84 Rau D C Parsegian V A 1992 Biophys. J. 61 246
85 Schellman J A 1974 Biopolymers 13 217
86 Chen H Yan J 2008 Phys. Rev. E 77 041907
87 Chen H Liu Y Zhou Z Hu L Ou-Yang Z C Yan J 2009 Phys. Rev. E 79 041926
88 Wu Y Y Bao L Zhang X Tan Z J 2015 J. Chem. Phys. 142 125103
89 Kochoyan M Leroy J L Guéron M 1987 J. Mol. Biol. 196 599
90 Coman D Russu I M 2005 Biophys. J. 89 3285
91 Driessen R P Sitters G Laurens N Moolenaar G F Wuite G J Goosen N Dame R T 2014 Biochemistry 53 6430
92 Porschke D 1991 Biophys. Chem. 40 169
93 Olson W K Marky N L Jernigan R L Zhurkin V B 1993 J. Mol. Biol. 232 530
94 Olson W K Gorin A A Lu X J Hock L M Zhurkin V B 1998 Proc. Natl. Acad. Sci. USA 95 11163
95 Scipioni A Anselmi C Zuccheri G Samori B De Santis P 2002 Biophys. J. 83 2408
96 Ortiz V de Pablo J J 2011 Phys. Rev. Lett. 106 238107
97 Rivetti C Guthold M Bustamante C 1996 J. Mol. Biol. 264 919
98 Le Cam E Culard F Larquet E Delain E Cognet J A 1999 J. Mol. Biol. 285 1011
99 Geggier S Vologodskii A 2010 Proc. Natl. Acad. Sci. USA 107 15421
100 MacDonald D Herbert K Zhang X Polgruto T Lu P 2001 J. Mol. Biol. 306 1081
101 Brukner I Dlakic M Savic A Susic S Pongor S Suck D 1993 Nucleic Acids Res. 21 1025
102 Lavery R Zakrzewska K Beveridge D Bishop T C Case D A Cheatham III T Dixit S Jayaram B Lankas F Laughton C Maddocks J H Michon A Osman R Orozco M Perez A Singh T Spackova N Sponer J 2010 Nucleic Acids Res. 38 299
103 Cluzel P Lebrun A Heller C Lavery R Viovy J L Chatenay D Caron F 1996 Science 271 792
104 Bustamante C Bryant Z Smith S B 2003 Nature 421 423
105 Fu H Chen H Zhang X Qu Y Marko J F Yan J 2011 Nucleic Acids Res. 39 3473
106 Zhang X Chen H Fu H Doyle P S Yan J 2012 Proc. Natl. Acad. Sci. USA 109 8103
107 Zhang X Chen H Le S Rouzina I Doyle P S Yan J 2013 Proc. Natl. Acad. Sci. USA 110 3865
108 Zhang X Qu Y Chen H Rouzina I Zhang S Doyle P S Yan J 2014 J. Am. Chem. Soc. 136 16073
109 Li J Wijeratne S S Qiu X Kiang C H 2015 Nanomaterials 5 246
110 Rouzina I Bloomfield V A 2001 Biophys. J. 80 882
111 Rouzina I Bloomfield V A 2001 Biophys. J. 80 894
112 Yan J Marko J F 2004 Phys. Rev. Lett. 93 108108
113 Cloutier T E Widom J 2004 Mol. Cell. 14 355
114 Vafabakhsh R Ha T 2012 Science 337 1097
115 Wiggins P A Van Der Heijden T Moreno-Herrero F Spakowitz A Phillips R Widom J Dekker C Nelson P C 2006 Nat. Nanotech. 1 137
116 Mazur A K Maaloum M 2014 Phys. Rev. Lett. 112 068104
117 Mazur A K Maaloum M 2014 Nucleic Acids Res. 42 14006
118 Mathew-Fenn R S Das R Harbury P A 2008 Science 322 446
119 Yan J Kawamura R Marko J F 2005 Phys. Rev. E 71 061905
120 Padinhateeri R Menon G I 2013 Biophys. J. 104 463
121 Timmons L Fire A 1998 Nature 395 854
122 Guo P 2010 Nat. Nanotech. 5 833
123 Bumcrot D Manoharan M Koteliansky V Sah D W 2006 Nat.Chem. Biol. 2 711
124 Metzler D 2003 Biochemistry: the Chemical Reactions of Living Cells 2nd edn. Amsterdam Elsevier
125 Gast F U Hagerman P J 1991 Biochemistry 30 4268
126 Kebbekus P Draper D E Hagerman P 1995 Biochemistry 34 4354
127 Shi Y Z Wu Y Y Wang F H Tan Z J 2014 Chin. Phys. B 23 078701
128 Tan Z J Zhang W B Shi Y Z Wang F H 2015 Adv. Expt. Med. Biol. 827 143
129 Frank J Agrawal R K 2000 Nature 406 318
130 Mandal M Breaker R R 2004 Nat. Rev. Mol. Cell Biol. 5 451
131 Fulle S Gohlke H 2008 Biophys. J. 94 4202
132 Murchie A I Davis B Isel C Afshar M Drysdale M J Bower J Potter A J Starkey I D Swarbrick T M Mirza S Prescott C D Vaglio P Aboul-ela F Karn J 2004 J. Mol. Biol. 336 625
133 Roh J H Tyagi M Briber R M Woodson S A Sokolov A P 2011 J. Am. Chem. Soc. 133 16406
134 Zacharias M Hagerman P J 1995 Proc. Natl. Acad. Sci. USA 92 6052
135 De Almeida Ribeiro E Beich-Frandsen M Konarev P V Shang WF Večerek B Kontaxis G Hänmerle H Peterlik H Svergun D I Bläsi U Djinović-Carugo K 2012 Nucleic Acids Res. 40 8072
136 Fulle S Gohlke H 2010 J. Mol. Recognit. 23 220
137 Zacharias M Hagerman P J 1995 J. Mol. Biol. 247 486
138 Zacharias M Hagerman P J 1996 J. Mol. Biol. 257 276
139 Salmon L Bascom G Andricioaei I Al-Hashimi H M 2013 J. Am. Chem. Soc. 135 5457
140 Al-Hashimi H M Gosser Y Gorin A Hu W Majumdar A Patel D J 2002 J. Mol. Biol. 315 95
141 Zhang Q Stelzer A C Fisher C K Al-Hashimi H M 2007 Nature 450 1263
142 Caliskan G Hyeon C Perez-Salas U Briber R M Woodson S A Thirumalai D 2005 Phys. Rev. Lett. 95 268303
143 Fulle S Gohlke H 2009 Methods 49 181
144 Mustoe A M Brooks C L Al-Hashimi H M 2014 Annu. Rev. Biochem. 83 441
145 Shi Y Z Wang F H Wu Y Y Tan Z J 2014 J. Chem. Phys. 141 105102
146 Pfleger C Rathi P C Klein D L Radestock S Gohlke H 2013 J. Chem. Inf. Model. 53 1007
147 Wiggins P A Phillips R Nelson P C 2005 Phys. Rev. E 71 021909
148 Noy A Golestanian R 2012 Phys. Rev. Lett. 109 228101
149 Xu X Thio B J R Cao J 2014 J. Phys. Chem. Lett. 5 2868
150 Widom J Baldwin R L 1980 J. Mol. Biol. 144 431
151 Bloomfield V A 1997 Biopolymers 44 269
152 Tolokh I S Pabit S A Katz A M Chen Y Drozdetski A Baker N Pollack L Onufriev A V 2014 Nucleic Acids Res. 42 10823
153 Wu Y Y Zhang Z L Zhang J S Zhu X L Tan Z J 2015 Nucleic Acids Res. 43 6156
154 Lipfert J Doniach S Das R Herschlag D 2014 Annu. Rev. Biochem. 83 813
155 Draper D E Grilley D Soto A M 2005 Annu. Rev. Biophys. Biomol. Struct. 34 221
156 Woodson S A 2005 Curr. Opin. Chem. Biol. 9 104
157 Tan Z J Chen S J 2010 Biophys. J. 99 1565
158 Tan Z J Chen S J 2011 Biophys. J. 101 176
159 Serganov A Nudler E 2013 Cell 152 17
160 Cech T R Steitz J A 2014 Cell 157 77
161 Chen J W Zhang W B 2012 J. Chem. Phys. 137 225102
162 Gong S Wang Y J Zhang W B 2015 J. Chem. Phys. 142 015103