Next Article in Journal
Breast Cancer Mortality in the Americas and Australasia over the Period 1980–2017 with Predictions for 2025
Next Article in Special Issue
Pressure Perturbation Studies of Noncanonical Viral Nucleic Acid Structures
Previous Article in Journal
Polyculture of Juvenile Dog Conch Laevistrombus canarium Reveals High Potentiality in Integrated Multitrophic Aquaculture (IMTA)
Previous Article in Special Issue
A Pathfinder in High-Pressure Bioscience: In Memoriam of Gaston Hui Bon Hoa
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Volumetric Properties of Four-Stranded DNA Structures

by
Tigran V. Chalikian
* and
Robert B. Macgregor, Jr.
Department of Pharmaceutical Sciences, Leslie Dan Faculty of Pharmacy, University of Toronto, 144 College Street, Toronto, ON M5S 3M2, Canada
*
Author to whom correspondence should be addressed.
Biology 2021, 10(8), 813; https://doi.org/10.3390/biology10080813
Submission received: 15 July 2021 / Revised: 18 August 2021 / Accepted: 19 August 2021 / Published: 22 August 2021

Abstract

:

Simple Summary

The volumetric properties of biomolecules define their pressure stability, while also characterizing their intrinsic and hydration properties. In this paper, we review the recent progress in volumetric investigations of G-quadruplexes and i-motifs, four-stranded secondary structures of DNA that have been found in the cell and implicated in regulatory genomic functions. Although the volumetric studies of G-quadruplexes and i-motifs are still in their nascent state, the data on volume, expansibility, and compressibility accumulated to date have begun to provide insights into the balance of forces governing the stability of these non-canonical structures. We present the available volumetric data and discuss how they can be rationalized in terms of intra-and intermolecular interactions involving G-quadruplexes and i-motifs including their solute-solvent interactions.

Abstract

Four-stranded non-canonical DNA structures including G-quadruplexes and i-motifs have been found in the genome and are thought to be involved in regulation of biological function. These structures have been implicated in telomere biology, genomic instability, and regulation of transcription and translation events. To gain an understanding of the molecular determinants underlying the biological role of four-stranded DNA structures, their biophysical properties have been extensively studied. The limited libraries on volume, expansibility, and compressibility accumulated to date have begun to provide insights into the molecular origins of helix-to-coil and helix-to-helix conformational transitions involving four-stranded DNA structures. In this article, we review the recent progress in volumetric investigations of G-quadruplexes and i-motifs, emphasizing how such data can be used to characterize intra-and intermolecular interactions, including solvation. We describe how volumetric data can be interpreted at the molecular level to yield a better understanding of the role that solute–solvent interactions play in modulating the stability and recognition events of nucleic acids. Taken together, volumetric studies facilitate unveiling the molecular determinants of biological events involving biopolymers, including G-quadruplexes and i-motifs, by providing one more piece to the thermodynamic puzzle describing the energetics of cellular processes in vitro and, by extension, in vivo.

1. Introduction

DNA molecules rich in guanine are prone to folding into four-stranded G-quadruplex structures, while cytosine-rich molecules tend to fold into four-stranded i-motif structures at slightly acidic pH [1,2,3,4,5,6,7,8,9]. G-quadruplexes are formed by stacking of two or more G-tetrads on top of each other. A G-tetrad represents a cyclic planar construct in which four guanine bases are linked together via Hoogsteen hydrogen bonds as shown in Figure 1a. The stacking results in the formation of a central cavity in which mono- or divalent cations are coordinated to the O6 atoms of guanines [1,2,3,7,10,11,12]. Sodium and potassium are the two biologically most relevant cations stabilizing G-quadruplex structures. The four consecutive G-runs involved in the formation of stacked tetrads in an intramolecular G-quadruplex are connected to each other via three single stranded linkers known as loops. The polarity of the loops defines the specific topology assumed by the G-quadruplex. G-quadruplexes can assume a parallel, antiparallel, or hybrid topologies with the molecularity ranging from mono to tetra (Figure 1b) [7,8,13].
Cytosine-rich DNA molecules can fold into the four-stranded i-motif conformation in which two parallel duplexes interact in an anti-parallel manner through mutual intercalation of hemiprotonated cytosine base pairs (Figure 2a,b) [4,5,6,14,15,16]. Although i-motif structures are favored at slightly acidic pH, they may exist at neutral pH, making the study of i-motifs biologically relevant [4,14,17,18,19,20]. In fact, there is increasing evidence suggesting that four-stranded nucleic acid structures, including G-quadruplex and i-motif structures, exist in the cell and are involved in regulation of genomic events including telomere control, gene expression, and DNA replication [10,21,22,23,24,25,26,27,28,29,30]. The existence of i-motifs in vivo may be related to an increase in the pKa of cytosine protonation in the crowded environment of the cell [31]. In addition, given the excluded volume effect, crowders may stabilize the compact i-motif conformation relative the extended unfolded conformation [31].
The thermodynamic and kinetic properties of interconversions between the duplex, tetraplex, and single-stranded conformations will necessarily constrain and define the role these structures play in vivo [7,32,33]. The stability characteristics of G-quadruplex and i-motif structures have been extensively and systematically studied by varying parameters such as temperature, pH, salt, and the concentration of cosolvent [4,5,7,31,32,34,35,36,37,38,39,40,41,42]. These studies have provided a wealth of information about the modulation of the differential free energy of the folded and unfolded conformations of the two tetrahelical DNA structures as a function of temperature, pH, salt, and the identity and concentration of cosolvents [32]. Comparative analyses of thermodynamic data have provided valuable insights into the contributions of intra- and intermolecular interactions (e.g., counterion-DNA interactions) and individual structural features (e.g., the length and nucleotide content of loops) to the stability of G-quadruplexes and i-motifs [32].
The volumetric data obtained from pressure-dependent measurements complement the stability data afforded by more conventional temperature-, pH-, salt-, and cosolvent-dependent studies [43,44,45,46,47,48,49]. Analysis of the effect of hydrostatic pressure on the equilibrium between the folded and unfolded DNA structures yield molar changes in volume, expansibility, and compressibility that accompany formation of the folded structures. The volumetric properties of solutes are determined by the entire ensemble of intra- and intermolecular interactions involving the solute, including solute–solvent interactions, and volume and energy fluctuations of the solute molecule [45,46,47,50,51,52,53,54,55]. The volumetric characteristics of G-quadruplexes and i-motifs reflect their structural and hydration properties, which differ significantly from those of other nucleic acid secondary structures, such as the B-form DNA duplex. The four-stranded structures are globular in shape with a surface charge density lower than that of other DNA structures; in addition, the G-quadruplex exhibits a compressible and expandable internal cavity [34,35,56,57].
In this work, we describe the current state of the art and give an overview of the studies that have dealt with the volumetric characterization of G-quadruplex and i-motif structures. We begin by defining volumetric properties. Next, we outline the experimental methods that have been used in volumetric investigations and explain how macroscopic volumetric properties can be rationalized to gain microscopic insights. We subsequently proceed to reviewing published data on changes in volume, expansibility, and compressibility accompanying conformational transitions involving tetraplex DNA structures. Finally, we discuss the use of volumetric properties to construct the pressure–temperature diagram of G-quadruplex stability.

2. Definitions and Experimental Methods

2.1. Observables

The partial molar volume of a solute, , is the pressure slope of its chemical potential, μ:
V ° = ( μ P ) T = lim C 0 ( V N ) T , P
where P is the pressure; T is the temperature; V is the volume of solution; N is the number of moles of a solute in solution; and C is the concentration of a solute.
The partial molar expansibility of a solute is the temperature derivative of its partial molar volume:
E ° = ( V ° T ) P = ( 2 μ P T ) = lim C 0 ( α V N ) T , P
where α = 1 V ( V T ) P is the coefficient of thermal expansibility of solution; and E = αV is the expansibility of solution.
The partial molar isothermal compressibility of a solute is the negative pressure derivative of partial molar volume:
K ° T = ( V ° P ) T = ( 2 μ P 2 ) T = lim C 0 ( β T V N ) T , P
where βT = 1 V ( V P ) T is the coefficient of isothermal compressibility of solution; and KT = βTV is the isothermal compressibility of solution.
The partial molar adiabatic compressibility of a solute is given by:
K ° S = lim C 0 ( β S V N ) T , P
where βS = − 1 V ( V P ) S is the coefficient of adiabatic compressibility of solution; S is the entropy; and KS = βSV is the adiabatic compressibility of solution.
Partial molar isothermal and adiabatic compressibilities are related to each other via the relationship [58,59]:
K ° T = K ° S + T α 0 2 ρ 0 C P 0 · ( 2 E ° α 0 C ° P ρ 0 c P 0 )
where α0, ρ0, and cP0 are the coefficient of thermal expansion, density, and specific heat capacity of the neat solvent, respectively; and P is the partial molar heat capacity of a solute.
According to scaled particle theory, the partial molar volume, , can be broken down into the following terms [60,61]:
= VC + VI + βT0RT
where VC is the volume of the cavity comprising a solute; it is given by VC = VM + VT; VM is the molecular volume of a solute; VT is the void volume around a solute (the thermal volume); VI is the interaction volume that is a change in solvent volume under the influence of solute-solvent interactions; and βT0 is the coefficient of isothermal compressibility of solvent.
Interaction volume, VI is related to the properties of water of solute hydration via VI = nh(V°h − V°0), where nh is the hydration number (the number of water molecules influenced by the solute), and h and 0 are the partial molar volumes of water of hydration and bulk water, respectively.
The partial molar expansibility, , and adiabatic compressibility, S, of a solute are related to its intrinsic and hydration properties as follows:
E° = EM + nh(E°h − E°0)
S = KM + nh(K°Sh − K°S0)
where EM and KM are, respectively, the intrinsic expansibility and compressibility of the solute molecule; h and 0 are the partial molar expansibilities of water of hydration and bulk water, respectively; and Sh and S0 are the partial molar adiabatic compressibilities of water of hydration and bulk water, respectively.
The volumetric properties of a solute can be expressed more rigorously based on the concepts of statistical thermodynamics in which water of hydration is represented by a heterogeneous network of solvent molecules with varying affinities for the solute [62,63,64]. The statistical thermodynamic formalism has been extended to the analysis of the volumetric properties of solutes in binary solvents consisting of water and water-miscible cosolvents [51,65].

2.2. Experimental Techniques

Differential measurements of density of solution and solvent have been widely employed to measure the partial molar volume, , of solutes and changes in volume, ΔV, accompanying their binding events and conformational transitions [66,67,68,69,70,71,72,73,74,75,76,77]. The partial molar volume, , of a solute can be determined from density data as follows:
V ° = M ρ 0 ρ ρ 0 ρ 0 C = M ρ ρ ρ 0 ρ ρ 0 m
where ρ and ρ0 are the densities of solution and solvent, respectively; M is the molecular mass of a solute; and C and m are the molar and molal concentrations of a solute, respectively.
If measurements of volume are carried out as a function of temperature, the resulting data can be used to determine the partial molar expansibility, , of a solute or a change in expansibility, ΔE, accompanying a reaction involving a solute such as a conformational transition or ligand binding [78]. Pressure-perturbation calorimetry (PPC) offers an alternative way to determine the partial molar expansibility, , of a solute as a function of temperature [79,80,81,82,83]. If a change in temperature within the experimental range causes a conformational transition of the solute, the measured E°(T) profile will display a characteristic peak. Given E° = ( V ° T )P [see Equation (2)], a change in volume, ΔV, accompanying the conformational transition equals the area under the peak:
Δ V = T 1 T 2 E o T d T
A combination of density and sound velocity measurements can be used to determine the partial molar adiabatic compressibility, S, of solutes and changes in adiabatic compressibility, ΔKS, accompanying binding events and conformational transitions [84,85]. Sound velocity, U, in a medium is related to its density, ρ, and coefficient of adiabatic compressibility, βS, via the Newton-Laplace equation: U2 = (ρβS)−1. Differentiation of this equation with respect to the concentration of a solute yields the following relationship for its partial molar adiabatic compressibility for the limit of infinite dilution [84,86,87]:
K ° S = β S 0 2 V ° 2 U M ρ 0
where βS0 is the coefficient of adiabatic compressibility of solvent; [U] = U U 0 U 0 C is the relative molar sound velocity increment of a solute; U and U0 are the sound velocities in solution and solvent, respectively.
Pressure-dependent measurements of conformation-sensitive spectroscopic parameters (typically, light absorption, fluorescence, NMR, and, more recently, circular dichroism) have been used to monitor pressure-induced shifts in the conformational equilibria of proteins and nucleic acids [88,89,90,91]. Provided that the solute population is restricted to two conformations (folded and unfolded), such measurements enable one to determine the equilibrium constant, K, as a function of pressure. In turn, the pressure dependence of K can be used to calculate a change in volume, ΔV, accompanying the conformational transition:
Δ V = R T l n K P T
Alternatively, ΔV can be determined from the Clausius–Clapeyron relation:
d T M d P = T M Δ V Δ H
where TM is the transition temperature; and ΔH is the transition enthalpy.

3. Differential Volume of Four-Stranded and Single-Stranded Conformations

3.1. G-quadruplexes

Table 1 presents a compilation of literature data on changes in volume accompanying unfolding transitions of G-quadruplexes differing in topology and sequence. Inspection of Table 1 reveals that G-quadruplex-to-single strand transitions are accompanied by negative changes in volume, ΔV [48,92,93,94,95,96,97]. To rationalize a change in volume accompanying G-quadruplex unfolding, one needs to carefully consider the entire set of molecular interactions that may contribute to the change. Equation (6) can be modified to analyze the molecular origins underlying the negative values of ΔV observed for G-quadruplex unfolding:
ΔV = ΔVM + ΔVT + ΔVI + nM+VM+
where nM+ is the number of stabilizing cations released from the central cavity to the bulk; and VM+ is the partial molar volume of the stabilizing cation.
The differential molecular, ΔVM, and thermal, ΔVT, volumes of the G-quadruplex and coiled states of DNA can be computed from the X-ray or NMR structure of the G-quadruplex conformation and the molecular dynamics-simulated single-stranded conformations. A change in thermal volume, ΔVT, is related to a change in solvent-accessible surface area, ΔSA, via ΔVT = δΔSA, where δ = 0.5 Å is the thickness of the thermal volume [98].
Table 2 shows the computed changes in intrinsic (molecular) volume, ΔVM, and solvent accessible surface area, ΔSA, associated with the unfolding transitions of three G-quadruplex structures, specifically, the Na+-stabilized antiparallel human telomeric G-quadruplex Tel22 [d(A(G3T2A)3G3)], the K+-stabilized hybrid human telomeric G-quadruplex Tel26 [d(A3(G3T2A)3G3A2)], and the K+-stabilized parallel c-MYC G-quadruplex [d(TGAG3TG3TAG3TG3T2)]. It is tempting to ascribe the negative change in volume to the presence of the intramolecular cavity within the G-quadruplex, which makes it distinct from other secondary structures (e.g., double- and triple-stranded DNA). However, elimination of the central cavity due to unfolding provides just one of the negative contributions to the change in volume. This source of negative volume change must be considered in concert with other contributions such as the release of counterions and changes in hydration and thermal volume.
The release of cations internally bound inside the central cavity contributes to the volumetric properties of G-quadruplex transitions. The nM+VM+ term in Equation (14) serves to take this contribution into account. On the other hand, G-quadruplex unfolding is not accompanied by a pronounced release of externally bound (condensed) counterions [34,35,56]. In this respect, G-quadruplexes are distinct from other DNA structures, such as the double-stranded B-DNA. Depending on the specific G-quadruplex topology and loop and flanking sequences, the unfolding transition may lead to a very slight release, no release, or significant uptake of external counterions [34,35,56]. Thus, the uptake or release of these counterions should contribute only modestly to the observed changes in volume.
Inspection of data in Table 2 reveals a change in molecular volume, ΔVM, which involves the contribution due to the elimination of the central cavity, is not the main component of the experimentally measured change in volume accompanying the unfolding transitions of G-quadruplexes. The thermal, ΔVT, and interaction, ΔVI, contributions are large and play a decisive role in determining the magnitude and the sign of the overall change in volume, ΔV, associated with G-quadruplex unfolding.
Changes in volume, ΔV, accompanying G-quadruplex unfolding transitions have been combined with structural data on ΔVM and ΔSA to estimate the values of ΔVI and changes in hydration, Δnh [92,93]. The estimates of Δnh were 103, 432, and 170 for the Tel22, Tel26, and c-MYC G-quadruplexes, respectively [92,93]. Thus, G-quadruplex-to-coil transitions are all accompanied by an increase in hydration with a considerable uptake of water from the bulk. The estimated changes in hydration represent the differential hydration of the folded and unfolded states. A recent crystallographic study revealed elaborate, topology-dependent, organized networks of water molecules in the grooves and loop regions of G-quadruplexes [100]. It was found that the primary sphere water molecules make direct contacts with groove and loop atomic groups, thereby contributing to the stability of the specific G-quadruplex topology and loop conformations [100]. Disruption of such water networks contributes to the changes in hydration that are observed in and evaluated from the results of volumetric measurements.

3.2. Influence of the Bases in the Loops

The Sugimoto and Macgregor groups simultaneously published systematic studies of the extent to which the loops of G-quadruplex structures influence the effect of hydrostatic pressure on the stability of these tetrahelical structures [97,101]. The experimental space in such measurements is very large; for each simple, monomolecular oligonucleotide that can fold into a G-quadruplex, there are three different loops, each of which can have a different sequence and number of bases (or other linking moieties). In this respect, loops are highly polymorphic; and we are still a long way from an understanding of how the loops influence molar changes in volume that accompany folding of intramolecular G-quadruplexes.
Two systems have been studied; these are the thrombin binding aptamer (TBA) and the human telomeric sequence (Tel22) [97,101]. In order to attempt to isolate the factors that arise solely from the loops, the sequence of the loops has been systematically altered while preserving the topology of the folded structure [97,101]. The original and altered G-quadruplex-forming oligonucleotides that have been investigated are shown in Table 3 and Table 4. In addition to nucleic acid base substitutions, Takahashi and Sugimoto examined the behavior of TBA derivatives in which the loops were replaced by a 12-carbon methylene linker, (-CH2-)12 [101]. These substitutions remove the ability of the bases in the loops to stack with the G-tetrads in TBA.
Inspection of the data in Table 3 and Table 4 reveals that the response of the TBA- and Tel22-based systems to hydrostatic pressure is qualitatively similar. The unfolding transitions of the studied G-quadruplexes shift to lower temperatures with increasing pressure regardless of the sequence of the loops, even when a (-CH2-)12 linker substitutes the bases of the loops; changes in volume, ΔV, accompanying unfolding of the G-quadruplexes are all negative. However, the magnitude of ΔV depends on the nature of the substitution. In some cases, the substitution leads to a greater pressure sensitivity, while, in other cases, it does not result in an appreciable change in ΔV. Note that Takahashi and Sugimoto [101] reported their ΔV values at the TM of G-quadruplex unfolding, whereas the data reported by Li et al. [97] were extrapolated to a common temperature of 57 °C (near the transition temperatures). Although the TM values of the G-quadruplexes in the two studies appear to be similar, a more accurate picture of the role of the loops in the volume change of G-quadruplex unfolding would emerge if the values of ΔV from the two studies are extrapolated to the same temperature.
Takahashi and Sugimoto have also studied the pressure dependence of the stability of the TBA-based G-quadruplexes in aqueous solutions containing ethylene glycol, PEG200, and PEG4000 [94,95,101]. The values of ΔV for the TBA G-quadruplex and its derivatives in the presence of PEG200 are shown in Table 3. The purpose of including cosolvents is to study the role of solvation of the loops in the observed pressure dependences. Independent of the type of the cosolvent, the effect of pressure on the stability was significantly reduced in solutions containing cosolvents relative to the values of ΔV obtained in an aqueous buffer. For example, a change in volume, ΔV, accompanying the unfolding of TBA G-quadruplex decreases from −54.6 cm3 mol−1 in water to −12.9 cm3 mol−1 in PEG200 [94]. This finding was rationalized in terms of changes in interaction volume, ΔVI, in Equation (14); it appears that the effect of the loop sequence on the molar changes in volume arises predominantly from the differential hydration of the loops.
It is apparent that other factors such as topological differences (i.e., parallel, anti-parallel, etc.) may also contribute to the observed behavior. In addition, it seems reasonable to propose that the stacking of the G-tetrads or the stacking of the bases in the loops with the G-tetrads might also significantly contribute to the molar volume change of unfolding. The role of the stacking of the G-tetrads has not been directly assessed to date. The data originating from the -(CH2)12-substituted oligonucleotides provide insights into how the stacking of the bases in the loops with the two terminal G-tetrads influence the volume change. Currently, however, it is difficult to assign with any degree of confidence the differential volumetric properties of the oligonucleotide constructs with and without the -(CH2)12- links to any one specific molecular interaction. Additional measurements are required in order to parse the roles of stacking and folding topology in the pressure-dependent stability of G-quadruplexes.

3.3. i-Motifs

Table 5 lists changes in volume, ΔV, measured for heat-induced and pH-induced i-motif-to-coil transitions. In contrast to G-quadruplexes, which are characterized by large negative changes in volume upon unfolding, i-motif-to-single strand transitions exhibit near-zero changes in volume [57,102,103]. Because of the small value of the volume change, the stability of i-motifs is nearly insensitive to hydrostatic pressure. In agreement with this expectation, the transition temperatures of the heat-induced i-motif-to-coil transitions do not change or change very weakly with pressure [57,102,103]. When treating the volumetric properties of i-motif structures derived from measurements at high pressures, it is important to account for the pressure-induced changes in the pH of the solution, as the stability of an i-motif critically depends on pH. It is easy to confound the resulting pH-induced change in i-motif stability for its pressure dependence with the resulting change in volume.
In the absence of structural data on i-motifs, one cannot reliably estimate the magnitude or even the sign of ΔVM in Equation (14). As mentioned above, the value of ΔVT correlates with a change in solvent-accessible surface area, SA, of the DNA associated with the i-motif-to-coil transition. Since the i-motif conformation is more compact than the unfolded conformation (ΔSA is negative), the change, ΔVT, should be negative. The magnitude and the sign of ΔVI are more difficult to assess. On the one hand, polar groups of hemiprotonated cytosine residues that are hydrogen-bonded with water in the coil state become buried within the interior of the i-motif, thereby diminishing the extent of solute−solvent interactions. On the other hand, the proximity of negatively charged phosphate groups within i-motif conformation should increase the charge density and enhance the volume-reducing effect of solute−solvent interactions, ΔVI. The experimentally measured value of ΔV ≈ 0 suggests a near perfect compensation between the ΔVM, ΔVT, and ΔVI terms in Equation (14).

4. Differential Expansibility

Table 6 presents changes in expansibility, ΔE, associated with G-quadruplex-to-coil transitions of the Tel22 and Tel26 telomeric sequences. The partial molar expansibilities, , of the two G-quadruplexes increase upon unfolding. The relationship for ΔE can be obtained by modifying Equation (7):
ΔE = ΔEM + Δnh(E°h − E°0) + nM+EM+ + ΔErel
where EM+ is the is the partial molar expansibility of the stabilizing cation; ΔErel is the change in the relaxation contribution, Erel = (<ΔHΔV> − <ΔH><ΔV>)/RT2; <ΔH> and <ΔV> are, respectively, the ensemble average changes in enthalpy and volume relative to a “ground state” conformation.
The intrinsic term in Equation (15) is given by ΔEM = αMUVMU − αMFVMF, where αMU and αMF are the intrinsic coefficients of thermal expansibility of the unfolded (coil) and folded (G-quadruplex) states, respectively; and VMU and VMF are the intrinsic volumes of the unfolded (coil) and folded (G-quadruplex) states, respectively. The intrinsic coefficient of thermal expansibility of the unfolded state, αMU, is close to zero. There are currently no data on the value of αMF; however, one would expect the intrinsic coefficient of thermal expansibility of the folded state, αMF, to be a sizeable quantity owing to the expandable central cavity. Therefore, ΔEM should be negative for G-quadruplex unfolding. In the absence of data, it seems reasonable to propose that the value of αMF is on the order of the intrinsic coefficient of thermal expansibility of globular proteins. In common with G-quadruplexes, globular proteins are characterized by potentially expandable internal voids [104,105]. The average intrinsic coefficient of thermal expansibility of globular proteins is ~1 × 10−4 K−1 [50].
The relaxation component, ΔErel, originates from the existence of a broadly distributed isoenergetic population of unfolded conformations differing in enthalpy and volume. An increase in temperature shifts the population of coil-like conformations towards the species with a greater enthalpy, which would result in a positive or negative value of ΔErel depending on the sign of the volume difference between the high and low enthalpy subpopulations [93].
At room temperature, all low molecular weight model compounds studied to date exhibit positive partial molar expansibilities, [106,107,108,109,110,111,112,113]. The intrinsic expansibility, EM, of low molecular weight compounds is close to zero. Hence, according to Equation (7), the positive values of of small molecules are suggestive of the positive values of the differential expansibility of water of hydration and bulk water, (h − E°0), for all functional groups independent of their chemical nature. The hydration term in Equation (15), Δnh(E°h − E°0), is positive given the positive values of Δnh (water is taken up by the hydration shell of the DNA upon its G-quadruplex-to-coil transition) and (h − E°0).

5. Differential Compressibility

5.1. G-quadruplexes

Table 6 presents changes in adiabatic compressibility, ΔKS, accompanying unfolding transitions of the Tel22, Tel26, and c-MYC G-quadruplexes. The partial molar adiabatic compressibilities, S, of the three G-quadruplexes all decrease upon their unfolding. The relationship for ΔKS can be obtained by modifying Equation (8):
ΔKS = ΔKM + Δnh(K°Sh − K°S0) + nM+KSM+ + ΔKSrel
where KSM+ is the partial molar adiabatic compressibility of the stabilizing ion. The relaxation contribution in Equation (16) is given by KSrel = (<ΔV2> − <ΔV>2) / RT, where <ΔV> is the ensemble average changes in volume relative to a “ground state” conformation. The intrinsic compressibility, KM, of a solute is given by KM = βMVM, where βM is the intrinsic coefficient of adiabatic compressibility.
Owing to the presence of the central cavity, a G-quadruplex possesses a compressible interior that is absent in its unfolded state. In other words, the value of βM of the native G-quadruplex is sizeable, but reduces to near zero in the unfolded state, which is devoid of a compressible interior. There are no estimates of the value of βM for G-quadruplexes. However, given the structural similarity of G-quadruplexes and globular proteins, the βM of a G-quadruplex can be expected to be on the order of 25 × 10−6 bar−1, the average coefficient of adiabatic compressibility of a globular protein [46,50,114,115,116]. Further studies, particularly, the pressure-dependent NMR characterization of native G-quadruplexes, may help evaluate the value of β M = 1 V M V M P T , which, in turn, would assist in a more reliable estimate of the molecular determinants of changes in compressibility accompanying conformational transitions involving G-quadruplexes.
For nucleic acids, the hydration term, Δnh(K°Sh − K°S0), in Eq.(16) is negative given the positive sign of Δnh and the negative sign of (Sh − K°S0) [44,47,117,118,119]. The volumetric properties of water of hydration of nucleic acids are dominated by the hydration of charged and clustered polar groups, which exhibit partial molar adiabatic compressibilities, Sh, of water of hydration lower than that of bulk water, S0 [45,120]. In contrast, isolated polar groups and nonpolar groups may exhibit partial molar compressibilities that are greater than that of bulk water [45,120]. Currently, it is difficult to come up with a reliable estimate of the value of (Sh − K°S0) for parsing the measured changes in compressibility, ΔKS, in terms of specific components using Equation (16). For charged solutes, such as DNA, the partial molar adiabatic compressibility, Sh, of water of hydration is ~25% smaller than that of bulk water [64]. However, the functional groups that become solvent exposed and subsequently solvated when a G-quadruplex unfolds are not all charged. In contrast, many of them are polar uncharged or even nonpolar; for example, the functional groups of nucleic acid bases that, due to their participation in base pairing and stacking interactions in the folded state, are shielded from the solvent, but become solvent-exposed and hydrated when the structure unfolds. Consequently, there is no clarity about the specific value of (Sh − K°S0) that should be used in the analysis.
The relaxation component, ΔKrel, originates from the broadly distributed ensemble of nearly isoenergetic unfolded single-stranded conformations that vary in volume. An increase in pressure shifts the ensemble of unfolded conformations towards the species with smaller volumes that gives rise to an additional positive contribution to the observed change in compressibility associated with G-quadruplex unfolding [93].

5.2. i-Motifs

There is only a single work that reports a change in adiabatic compressibility, ΔKS, accompanying the helix-to-coil transition of an i-motif [57]. For the c-MYC i-motif, the value of ΔKS for the pH-induced helix-to-coil transition at 25 °C is nearly zero (see Table 5) [57].
As mentioned above, the intrinsic compressibility term, KM, in Equation (16) represents the compressibility of the water-inaccessible interior core of the solute. While the intrinsic compressibility, KM, of the ensemble of unfolded conformations is close to zero, that of the folded i-motif conformation is less certain and, in principle, may be significant. The sign and the magnitude of hydration contribution to compressibility in Equation (16), Δnh(Kh − K0), is also unclear. The differential partial molar compressibility of hydration and bulk water, (Kh − K0), is negative for the unfolded, coil-like state. On the other hand, in the folded i-motif conformation, the value of (Kh − K0) may be more negative relative to the coil state (because the negatively charged phosphates are brought closer together), it may be the same, or it may be less negative (e.g., due to a more effective mutual neutralization of the positively charged cytosines and negatively charged phosphates). Should the change in intrinsic compressibility, ΔKM, in Equation (16) be near zero, that is the interior of the i-motif conformation is rigid, the observation that ΔKS ≈ 0 suggests the similarity of the hydration of the i-motif and coil states with ΔΔKh ≈ 0. Although the current data do not enable us to discriminate between the various scenarios, the observed ΔKS ≈ 0 is an indication of a near-perfect offsetting of the ΔKM and ΔΔKh terms in Equation (16). In addition, it should be noted that there may be a relaxation component in analogy with G-quadruplexes.
In the aggregate, the current volumetric data on i-motifs collected to date suggest that fortuitous compensations between the intrinsic and hydration contributions to volume and compressibility result in ΔV and ΔKS of zero. A compressibility change of zero implies that the compensations leading to ΔV being zero are not restricted to ambient pressure, but also act at elevated pressures.

6. Pressure-Temperature Phase Diagram

The stability of a G-quadruplex (or any other biopolymer) as a function of temperature and pressure can be presented analytically as follows [121,122,123,124,125,126]:
Δ G P , T = Δ H M 1 T T M + Δ C P T T M T l n T T M + Δ V T R + Δ E T T R P P R 0.5 Δ K T P 2 P R 2
where ΔHM is the differential enthalpy of the folded and unfolded states at the transition temperature, TM, and reference pressure, PR; ΔCP is the differential heat capacity of the folded and unfolded states; ΔV(TR) is the differential volume of the folded and unfolded states at the reference temperature, TR, and pressure, PR; and ΔE and ΔKT are, respectively, the differential expansibility and isothermal compressibility of the folded and unfolded states.
The relationship for pressure-temperature phase diagram can be derived by equating Equation (17) to zero and solving it with respect to the denaturation pressure, PM:
Δ H M 1 T T M + Δ C P T T M T l n T T M + Δ V T R + Δ E T T R P P R 0.5 Δ K T P 2 P R 2 = 0
The reference pressure, PR, is generally set equal to ambient pressure (1 bar). It can be ignored relative to pressures, PM, at which proteins and nucleic acids denature. The latter are, typically, on the order of ~1 kbar and higher:
0.5 Δ K T P M 2 Δ V T R + Δ E T T R P M Δ H M 1 T T M Δ C P T T M T l n T T M = 0
Solving Equation (19) with respect to PM, one obtains the following:
P M = Δ V T R + Δ E T T R ± D Δ K T
where D = Δ V T R + Δ E T T R 2 + 2 Δ K T Δ H M 1 T T M Δ C P T T M T l n T T M .
Equation (20) has been used to compute the pressure-temperature phase diagram of the c-MYC G-quadruplex [99]. In the computation, the change in heat capacity, ΔCP, has been evaluated based on the results reported by Majhi et al. [127].
Figure 3 presents the pressure–temperature stability phase diagram for the c-MYC G-quadruplex [99]. As is seen from Figure 3, the diagram is elliptic. It resembles that of a globular protein [121,122,124,128], while being distinct from that of duplex DNA [129]. We propose that the similarity of the pressure–temperature stability phase diagram of a G-quadruplex and a globular protein reflects their shared structural and volumetric features. In particular, both G-quadruplexes and globular proteins are compact and characterized by compressible intramolecular voids. The order-disorder transitions of both structures are accompanied by an increase in expansibility, ΔE, and reductions in volume, ΔV, and compressibility, ΔKT.
Pressure-induced structural changes in nucleic acid structures and the resulting inhibition of genomic processes may contribute to pressure-induced cell death and injury in microorganisms. We have previously suggested that the fine-tuned, differential pressure sensitivity of the duplex, G-quadruplex, and i-motif conformations adopted by specific genomic loci may be involved in regulation of genomic processes in barophilic organisms thereby governing their survival at high pressures [57]. The differential stability phase diagram of G-quadruplex and duplex DNA [129] may provide an additional platform for developing this hypothesis.

7. Conclusions

There is only a decade of volumetric studies of G-quadruplex and i-motif structures. In some ways, our understanding of the volumetric characteristics of these four-stranded structures is reminiscent of the situation with the volumetric studies of proteins in the 1980s. Despite their scarcity, the volumetric data accumulated so far have established several regularities that are common for each of these four-stranded structures. In particular, all G-quadruplex-to-coil transitions studied to date are accompanied by negative changes in volume, ΔV, and compressibility, ΔKS, and positive changes in expansibility, ΔE. The few i-motif-to-coil transitions studied are accompanied by near zero changes in volume, ΔV, and compressibility, ΔKS. While it is still difficult to reliably rationalize volumetric observations in terms of intrinsic and hydration contributions, they establish an experimental framework for deriving the pressure–temperature stability diagrams of tetraplex DNA structures. Further studies involving a wide range of G-quadruplex and i-motif structures are needed to understand the generality and molecular origins of these results.

Author Contributions

T.V.C. and R.B.M.J. performed research, analyzed data, and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from NSERC to T.V.C. and R.B.M.J.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Shafer, R.H.; Smirnov, I. Biological aspects of DNA/RNA quadruplexes. Biopolymers 2000, 56, 209–227. [Google Scholar] [CrossRef]
  2. Huppert, J.L. Four-stranded nucleic acids: Structure, function and targeting of G-quadruplexes. Chem. Soc. Rev. 2008, 37, 1375–1384. [Google Scholar] [CrossRef] [PubMed]
  3. Huppert, J.L. Structure, location and interactions of G-quadruplexes. FEBS J. 2010, 277, 3452–3458. [Google Scholar] [CrossRef]
  4. Day, H.A.; Pavlou, P.; Waller, Z.A. i-Motif DNA: Structure, stability and targeting with ligands. Bioorg. Med. Chem. 2014, 22, 4407–4418. [Google Scholar] [CrossRef] [PubMed]
  5. Benabou, S.; Avino, A.; Eritja, R.; Gonzalez, C.; Gargallo, R. Fundamental aspects of the nucleic acid i-motif structures. RSC Adv. 2014, 4, 26956–26980. [Google Scholar] [CrossRef] [Green Version]
  6. Alba, J.J.; Sadurni, A.; Gargallo, R. Nucleic acid i-motif structures in analytical chemistry. Crit. Rev. Anal. Chem. 2016, 46, 443–454. [Google Scholar] [CrossRef] [Green Version]
  7. Lane, A.N.; Chaires, J.B.; Gray, R.D.; Trent, J.O. Stability and kinetics of G-quadruplex structures. Nucleic Acids Res. 2008, 36, 5482–5515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Balasubramanian, S.; Hurley, L.H.; Neidle, S. Targeting G-quadruplexes in gene promoters: A novel anticancer strategy? Nat. Rev. Drug Discov. 2011, 10, 261–275. [Google Scholar] [CrossRef] [Green Version]
  9. Dolinnaya, N.G.; Ogloblina, A.M.; Yakubovskaya, M.G. Structure, properties, and biological relevance of the DNA and RNA G-quadruplexes: Overview 50 Years after their discovery. Biochemistry 2016, 81, 1602–1649. [Google Scholar] [CrossRef]
  10. Hansel-Hertsch, R.; Di Antonio, M.; Balasubramanian, S. DNA G-quadruplexes in the human genome: Detection, functions and therapeutic potential. Nat. Rev. Mol. Cell Biol. 2017, 18, 279–284. [Google Scholar] [CrossRef]
  11. Qin, Y.; Hurley, L.H. Structures, folding patterns, and functions of intramolecular DNA G-quadruplexes found in eukaryotic promoter regions. Biochimie 2008, 90, 1149–1171. [Google Scholar] [CrossRef] [Green Version]
  12. Sen, D.; Gilbert, W. Formation of parallel four-stranded complexes by guanine-rich motifs in DNA and its implications for meiosis. Nature 1988, 334, 364–366. [Google Scholar] [CrossRef]
  13. Burge, S.; Parkinson, G.N.; Hazel, P.; Todd, A.K.; Neidle, S. Quadruplex DNA: Sequence, topology and structure. Nucleic Acids Res. 2006, 34, 5402–5415. [Google Scholar] [CrossRef] [Green Version]
  14. Brooks, T.A.; Kendrick, S.; Hurley, L. Making sense of G-quadruplex and i-motif functions in oncogene promoters. FEBS J. 2010, 277, 3459–3469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Collie, G.W.; Parkinson, G.N. The application of DNA and RNA G-quadruplexes to therapeutic medicines. Chem. Soc. Rev. 2011, 40, 5867–5892. [Google Scholar] [CrossRef] [PubMed]
  16. Gehring, K.; Leroy, J.L.; Gueron, M. A Tetrameric DNA structure with protonated cytosine·cytosine base pairs. Nature 1993, 363, 561–565. [Google Scholar] [CrossRef]
  17. Sun, D.; Hurley, L.H. The importance of negative superhelicity in inducing the formation of G-quadruplex and i-motif structures in the c-Myc promoter: Implications for drug targeting and control of gene expression. J. Med. Chem. 2009, 52, 2863–2874. [Google Scholar] [CrossRef] [Green Version]
  18. Kang, H.J.; Kendrick, S.; Hecht, S.M.; Hurley, L.H. The transcriptional complex between the BCL2 i-motif and hnRNP LL is a molecular switch for control of gene expression that can be modulated by small molecules. J. Am. Chem. Soc. 2014, 136, 4172–4185. [Google Scholar] [CrossRef] [PubMed]
  19. Kendrick, S.; Kang, H.J.; Alam, M.P.; Madathil, M.M.; Agrawal, P.; Gokhale, V.; Yang, D.; Hecht, S.M.; Hurley, L.H. The dynamic character of the BCL2 promoter i-motif provides a mechanism for modulation of gene expression by compounds that bind selectively to the alternative DNA hairpin structure. J. Am. Chem. Soc. 2014, 136, 4161–4171. [Google Scholar] [CrossRef] [PubMed]
  20. Zhou, J.; Wei, C.; Jia, G.; Wang, X.; Feng, Z.; Li, C. Formation of i-motif structure at neutral and slightly alkaline pH. Mol. Biosyst. 2010, 6, 580–586. [Google Scholar] [CrossRef]
  21. Tateishi-Karimata, H.; Sugimoto, N. Chemical biology of non-canonical structures of nucleic acids for therapeutic applications. Chem. Commun. 2020, 56, 2379–2390. [Google Scholar] [CrossRef]
  22. Spiegel, J.; Adhikari, S.; Balasubramanian, S. The structure and function of DNA G-quadruplexes. Trends Chem. 2020, 2, 123–136. [Google Scholar] [CrossRef] [Green Version]
  23. Varshney, D.; Spiegel, J.; Zyner, K.; Tannahill, D.; Balasubramanian, S. The regulation and functions of DNA and RNA G-quadruplexes. Nat. Rev. Mol. Cell Biol. 2020, 21, 459–474. [Google Scholar] [CrossRef]
  24. Bryan, T.M. Mechanisms of DNA replication and repair: Insights from the study of G-quadruplexes. Molecules 2019, 24, 3439. [Google Scholar] [CrossRef] [Green Version]
  25. Armas, P.; David, A.; Calcaterra, N.B. Transcriptional control by G-quadruplexes: In vivo roles and perspectives for specific intervention. Transcription 2017, 8, 21–25. [Google Scholar] [CrossRef] [Green Version]
  26. Biffi, G.; Tannahill, D.; McCafferty, J.; Balasubramanian, S. Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 2013, 5, 182–186. [Google Scholar] [CrossRef] [PubMed]
  27. Lam, E.Y.; Beraldi, D.; Tannahill, D.; Balasubramanian, S. G-quadruplex structures are stable detectable in human genomic DNA. Nat. Commun. 2013, 4, 1796. [Google Scholar] [CrossRef] [Green Version]
  28. Biffi, G.; Di, A.M.; Tannahill, D.; Balasubramanian, S. Visualization and selective chemical targeting of RNA G-quadruplex structures in the cytoplasm of human cells. Nat. Chem. 2014, 6, 75–80. [Google Scholar] [CrossRef] [PubMed]
  29. Zeraati, M.; Langley, D.B.; Schofield, P.; Moye, A.L.; Rouet, R.; Hughes, W.E.; Bryan, T.M.; Dinger, M.E.; Christ, D. i-motif DNA structures are formed in the nuclei of human cells. Nat. Chem. 2018, 10, 631–637. [Google Scholar] [CrossRef]
  30. Abou Assi, H.; Garavis, M.; Gonzalez, C.; Damha, M.J. i-Motif DNA: Structural features and significance to cell biology. Nucleic Acids Res. 2018, 46, 8038–8056. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, B.G.; Chalikian, T.V. Thermodynamic linkage analysis of pH-induced folding and unfolding transitions of i-motifs. Biophys. Chem. 2016, 216, 19–22. [Google Scholar] [CrossRef] [PubMed]
  32. Chalikian, T.V.; Liu, L.; Macgregor, R.B., Jr. Duplex-tetraplex equilibria in guanine- and cytosine-rich DNA. Biophys. Chem. 2020, 267, 106473. [Google Scholar] [CrossRef] [PubMed]
  33. Takahashi, S.; Sugimoto, N. Stability prediction of canonical and non-canonical structures of nucleic acids in various molecular environments and cells. Chem. Soc. Rev. 2020, 49, 8439–8468. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, B.G.; Evans, H.M.; Dubins, D.N.; Chalikian, T.V. Effects of salt on the stability of a G-quadruplex from the human c-MYC promoter. Biochemistry 2015, 54, 3420–3430. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, B.G.; Long, J.; Dubins, D.N.; Chalikian, T.V. Ionic effects on VEGF G-quadruplex stability. J. Phys. Chem. B 2016, 120, 4963–4971. [Google Scholar] [CrossRef] [PubMed]
  36. Aslanyan, L.; Ko, J.; Kim, B.G.; Vardanyan, I.; Dalyan, Y.B.; Chalikian, T.V. Effect of urea on G-quadruplex stability. J. Phys. Chem. B 2017, 121, 6511–6519. [Google Scholar] [CrossRef]
  37. Sugimoto, N. Noncanonical structures and their thermodynamics of DNA and RNA under molecular crowding: Beyond the Watson-Crick double helix. Int. Rev. Cell Mol. Biol. 2014, 307, 205–273. [Google Scholar]
  38. Nakano, S.; Miyoshi, D.; Sugimoto, N. Effects of molecular crowding on the structures, interactions, and functions of nucleic acids. Chem. Rev. 2014, 114, 2733–2758. [Google Scholar] [CrossRef]
  39. Boncina, M.; Lah, J.; Prislan, I.; Vesnaver, G. Energetic basis of human telomeric DNA folding into G-quadruplex structures. J. Am. Chem. Soc. 2012, 134, 9657–9663. [Google Scholar] [CrossRef]
  40. Boncina, M.; Vesnaver, G.; Chaires, J.B.; Lah, J. Unraveling the thermodynamics of the folding and interconversion of human telomere G-quadruplexes. Angew. Chem. Int. Ed. Engl. 2016, 55, 10340–10344. [Google Scholar] [CrossRef] [Green Version]
  41. Olsen, C.M.; Gmeiner, W.H.; Marky, L.A. Unfolding of G-quadruplexes: Energetic, and ion and water contributions of G-quartet stacking. J. Phys. Chem. B 2006, 110, 6962–6969. [Google Scholar] [CrossRef]
  42. Kaushik, M.; Suehl, N.; Marky, L.A. Calorimetric unfolding of the bimolecular and i-motif complexes of the human telomere complementary strand, d(C3TA2)4. Biophys. Chem. 2007, 126, 154–164. [Google Scholar] [CrossRef] [PubMed]
  43. Macgregor, R.B., Jr. Effect of hydrostatic pressure on nucleic acids. Biopolymers 1998, 48, 253–263. [Google Scholar] [CrossRef]
  44. Chalikian, T.V.; Macgregor, R.B. Nucleic acid hydration: A volumetric perspective. Phys. Life Rev. 2007, 4, 91–115. [Google Scholar] [CrossRef]
  45. Chalikian, T.V.; Sarvazyan, A.P.; Breslauer, K.J. Hydration and partial compressibility of biological compounds. Biophys. Chem. 1994, 51, 89–107. [Google Scholar] [CrossRef]
  46. Chalikian, T.V.; Breslauer, K.J. Thermodynamic analysis of biomolecules: A volumetric approach. Curr. Opin. Struct. Biol. 1998, 8, 657–664. [Google Scholar] [CrossRef]
  47. Chalikian, T.V.; Breslauer, K.J. Volumetric properties of nucleic acids. Biopolymers 1998, 48, 264–280. [Google Scholar] [CrossRef]
  48. Takahashi, S.; Sugimoto, N. Pressure-dependent formation of i-motif and G-quadruplex DNA structures. Phys. Chem. Chem. Phys. 2015, 17, 31004–31010. [Google Scholar] [CrossRef] [PubMed]
  49. Winter, R. Interrogating the structural dynamics and energetics of biomolecular systems with pressure modulation. Annu. Rev. Biophys. 2019, 48, 441–463. [Google Scholar] [CrossRef] [PubMed]
  50. Chalikian, T.V. Volumetric properties of proteins. Annu. Rev. Biophys. Biomol. Struct. 2003, 32, 207–235. [Google Scholar] [CrossRef] [PubMed]
  51. Chalikian, T.V. Volumetric measurements in binary solvents: Theory to experiment. Biophys. Chem. 2011, 156, 3–12. [Google Scholar] [CrossRef]
  52. Cooper, A. Protein fluctuations and the thermodynamic uncertainty principle. Prog. Biophys. Mol. Biol. 1984, 44, 181–214. [Google Scholar] [CrossRef]
  53. Chen, C.R.; Makhatadze, G.I. Molecular determinants of temperature dependence of protein volume change upon unfolding. J. Phys. Chem. B 2017, 121, 8300–8310. [Google Scholar] [CrossRef]
  54. Chen, C.R.; Makhatadze, G.I. Molecular determinant of the effects of hydrostatic pressure on protein folding stability. Nat. Commun. 2017, 8, 14561. [Google Scholar] [CrossRef] [PubMed]
  55. Krobath, H.; Chen, T.; Chan, H.S. Volumetric physics of polypeptide coil-helix transitions. Biochemistry 2016, 55, 6269–6281. [Google Scholar] [CrossRef] [PubMed]
  56. Kim, B.G.; Shek, Y.L.; Chalikian, T.V. Polyelectrolyte effects in G-quadruplexes. Biophys.Chem. 2013, 184, 95–100. [Google Scholar] [CrossRef] [PubMed]
  57. Liu, L.; Kim, B.G.; Feroze, U.; Macgregor, R.B., Jr.; Chalikian, T.V. Probing the ionic atmosphere and hydration of the c-MYC i-motif. J. Am. Chem. Soc. 2018, 140, 2229–2238. [Google Scholar] [CrossRef]
  58. Blandamer, M.J.; Davis, M.I.; Douheret, G.; Reis, J.C.R. Apparent molar isentropic compressions and expansions of solutions. Chem. Soc. Rev. 2001, 30, 8–15. [Google Scholar] [CrossRef]
  59. Desnoyers, J.E.; Philip, P.R. Isothermal compressibilities of aqueous solutions of tetraalkylammonium bromides. Can. J. Chem. 1972, 50, 1094–1096. [Google Scholar] [CrossRef] [Green Version]
  60. Kharakoz, D.P. Partial molar volumes of molecules of arbitrary shape and the effect of hydrogen bonding with water. J. Solut. Chem. 1992, 21, 569–595. [Google Scholar] [CrossRef]
  61. Pierotti, R.A. Scaled particle theory of aqueous and non-aqueous solutions. Chem. Rev. 1976, 76, 717–726. [Google Scholar] [CrossRef]
  62. Chalikian, T.V. On the molecular origins of volumetric data. J. Phys. Chem. B 2008, 112, 911–917. [Google Scholar] [CrossRef]
  63. Patel, N.; Dubins, D.N.; Pomes, R.; Chalikian, T.V. Parsing partial molar volumes of small molecules: A molecular dynamics study. J. Phys. Chem. B 2011, 115, 4856–4862. [Google Scholar] [CrossRef] [PubMed]
  64. Chalikian, T.V. Structural thermodynamics of hydration. J. Phys. Chem. B 2001, 105, 12566–12578. [Google Scholar] [CrossRef]
  65. Lee, S.Y.; Chalikian, T.V. Volumetric properties of solvation in binary solvents. J. Phys. Chem. B 2009, 113, 2443–2450. [Google Scholar] [CrossRef]
  66. Han, F.; Chalikian, T.V. Hydration changes accompanying nucleic acid intercalation reactions:volumetric characterizations. J. Am. Chem. Soc. 2003, 125, 7219–7229. [Google Scholar] [CrossRef] [PubMed]
  67. Han, F.; Taulier, N.; Chalikian, T.V. Association of the minor groove binding drug Hoechst 33258 with d(CGCGAATTCGCG)2: Volumetric, calorimetric, and spectroscopic characterizations. Biochemistry 2005, 44, 9785–9794. [Google Scholar] [CrossRef] [PubMed]
  68. Filfil, R.; Chalikian, T.V. The thermodynamics of protein-protein recognition as characterized by a combination of volumetric and calorimetric techniques: The binding of turkey ovomucoid third domain to α-chymotrypsin. J. Mol. Biol. 2003, 326, 1271–1288. [Google Scholar] [CrossRef]
  69. Filfil, R.; Chalikian, T.V. Volumetric and spectroscopic characterizations of glucose-hexokinase association. FEBS Lett. 2003, 554, 351–356. [Google Scholar] [CrossRef] [Green Version]
  70. Son, I.; Shek, Y.L.; Dubins, D.N.; Chalikian, T.V. Volumetric characterization of tri-N-acetylglucosamine binding to lysozyme. Biochemistry 2012, 51, 5784–5790. [Google Scholar] [CrossRef]
  71. Son, I.; Selvaratnam, R.; Dubins, D.N.; Melacini, G.; Chalikian, T.V. Ultrasonic and densimetric characterization of the association of cyclic AMP with the cAMP-binding domain of the exchange protein EPAC1. J. Phys. Chem. B 2013, 117, 10779–10784. [Google Scholar] [CrossRef]
  72. Liu, L.; Stepanian, L.; Dubins, D.N.; Chalikian, T.V. Binding of L-argininamide to a DNA aptamer: A volumetric study. J. Phys. Chem. B 2018, 122, 7647–7653. [Google Scholar] [CrossRef] [PubMed]
  73. Filfil, R.; Chalikian, T.V. Volumetric and spectroscopic characterizations of the native and acid-induced denatured states of staphylococcal nuclease. J. Mol. Biol. 2000, 299, 827–842. [Google Scholar] [CrossRef]
  74. Taulier, N.; Chalikian, T.V. Characterization of pH-induced transitions of β-lactoglobulin: Ultrasonic, densimetric, and spectroscopic studies. J. Mol. Biol. 2001, 314, 873–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Taulier, N.; Beletskaya, I.V.; Chalikian, T.V. Compressibility changes accompanying conformational transitions of apomyoglobin. Biopolymers 2005, 79, 218–229. [Google Scholar] [CrossRef] [PubMed]
  76. Taulier, N.; Chalikian, T.V. Hydrophobic hydration in cyclodextrin complexation. J. Phys. Chem. B 2006, 110, 12222–12224. [Google Scholar] [CrossRef] [PubMed]
  77. Taulier, N.; Chalikian, T.V. γ-Cyclodextrin forms a highly compressible complex with 1-adamantanecarboxylic acid. J. Phys. Chem. B 2008, 112, 9546–9549. [Google Scholar] [CrossRef]
  78. Chalikian, T.V.; Volker, J.; Anafi, D.; Breslauer, K.J. The native and the heat-induced denatured states of α-chymotrypsinogen A: Thermodynamic and spectroscopic studies. J. Mol. Biol. 1997, 274, 237–252. [Google Scholar] [CrossRef] [PubMed]
  79. Lin, L.N.; Brandts, J.F.; Brandts, J.M.; Plotnikov, V. Determination of the volumetric properties of proteins and other solutes using pressure perturbation calorimetry. Anal. Biochem. 2002, 302, 144–160. [Google Scholar] [CrossRef]
  80. Suladze, S.; Kahse, M.; Erwin, N.; Tomazic, D.; Winter, R. Probing volumetric properties of biomolecular systems by pressure perturbation calorimetry (PPC)—The effects of hydration, cosolvents and crowding. Methods 2015, 76, 67–77. [Google Scholar]
  81. Heerklotz, H. Pressure Perturbation Calorimetry. In Methods in Molecular Biology; Dopica, A.M., Ed.; Humana Press, Inc.: Totowa, NJ, USA, 2007; pp. 197–206. [Google Scholar]
  82. Mitra, L.; Smolin, N.; Ravindra, R.; Royer, C.; Winter, R. Pressure perturbation calorimetric studies of the solvation properties and the thermal unfolding of proteins in solution—Experiments and theoretical interpretation. Phys. Chem. Chem. Phys. 2006, 8, 1249–1265. [Google Scholar] [CrossRef]
  83. Schweiker, K.L.; Makhatadze, G.I. Use of pressure perturbation calorimetry to characterize the volumetric properties of proteins. Methods Enzymol. 2009, 466, 527–547. [Google Scholar]
  84. Sarvazyan, A.P. Ultrasonic velocimetry of biological compounds. Annu. Rev. Biophys. Biophys. Chem. 1991, 20, 321–342. [Google Scholar] [CrossRef]
  85. Sarvazian, A.P. Ultrasonic velocimetry of biological compounds. Mol. Biol. 1983, 17, 916–927. [Google Scholar] [CrossRef]
  86. Owen, B.B.; Simons, H.L. Standard partial molal compressibilities by ultrasonics. 1. Sodium chloride and potassium chloride at 25 °C. J. Phys. Chem. 1957, 61, 479–482. [Google Scholar] [CrossRef]
  87. Barnartt, S. The velocity of sound in electrolytic solutions. J. Chem. Phys. 1952, 20, 278–279. [Google Scholar] [CrossRef]
  88. Lerch, M.T.; Horwitz, J.; McCoy, J.; Hubbell, W.L. Circular dichroism and site-directed spin labeling reveal structural and dynamical features of high-pressure states of myoglobin. Proc. Natl. Acad. Sci. USA 2013, 110, E4714–E4722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Dellarole, M.; Royer, C.A. High-pressure fluorescence applications. Methods Mol. Biol. 2014, 1076, 53–74. [Google Scholar] [PubMed]
  90. Akasaka, K. Probing conformational fluctuation of proteins by pressure perturbation. Chem. Rev. 2006, 106, 1814–1835. [Google Scholar] [CrossRef] [PubMed]
  91. Wu, J.Q.; Macgregor, R.B., Jr. Pressure dependence of the melting temperature of dA·dT polymers. Biochemistry 1993, 32, 12531–12537. [Google Scholar] [CrossRef] [PubMed]
  92. Fan, H.Y.; Shek, Y.L.; Amiri, A.; Dubins, D.N.; Heerklotz, H.; Macgregor, R.B., Jr.; Chalikian, T.V. Volumetric characterization of sodium-induced G-quadruplex formation. J. Am. Chem. Soc. 2011, 133, 4518–4526. [Google Scholar] [CrossRef]
  93. Shek, Y.L.; Noudeh, G.D.; Nazari, M.; Heerklotz, H.; Abu-Ghazalah, R.M.; Dubins, D.N.; Chalikian, T.V. Folding thermodynamics of the hybrid-1 type intramolecular human telomeric G-quadruplex. Biopolymers 2014, 101, 216–227. [Google Scholar] [CrossRef]
  94. Takahashi, S.; Sugimoto, N. Effect of pressure on the stability of G-quadruplex DNA: Thermodynamics under crowding conditions. Angew. Chem.-Int. Ed. 2013, 52, 13774–13778. [Google Scholar] [CrossRef] [PubMed]
  95. Takahashi, S.; Sugimoto, N. Effect of pressure on thermal stability of G-quadruplex DNA and double-stranded DNA structures. Molecules 2013, 18, 13297–13319. [Google Scholar] [CrossRef] [Green Version]
  96. Molnar, O.R.; Somkuti, J.; Smeller, L. Negative volume changes of human G-quadruplexes at unfolding. Heliyon 2020, 6, 6. [Google Scholar] [CrossRef]
  97. Li, Y.Y.; Dubins, D.N.; Le, D.; Leung, K.; Macgregor, R.B., Jr. The role of loops and cation on the volume of unfolding of G-quadruplexes related to HTel. Biophys. Chem. 2017, 231, 55–63. [Google Scholar] [CrossRef]
  98. Chalikian, T.V.; Macgregor, R.B., Jr. On empirical decomposition of volumetric data. Biophys. Chem. 2019, 246, 8–15. [Google Scholar] [CrossRef]
  99. Liu, L.; Scott, L.; Tariq, N.; Kume, T.; Dubins, D.N.; Macgregor, R.B., Jr.; Chalikian, T.V. Volumetric interplay between the conformational states adopted by guanine-rich DNA from the c-MYC promoter. J. Phys. Chem. B 2021, 125, 7406–7416. [Google Scholar] [CrossRef]
  100. Li, K.; Yatsunyk, L.; Neidle, S. Water spines and networks in G-quadruplex structures. Nucleic Acids Res. 2021, 49, 519–528. [Google Scholar] [CrossRef] [PubMed]
  101. Takahashi, S.; Sugimoto, N. Volumetric contributions of loop regions of G-quadruplex DNA to the formation of the tertiary structure. Biophys. Chem. 2017, 231, 146–154. [Google Scholar] [CrossRef] [PubMed]
  102. Lepper, C.P.; Williams, M.A.K.; Edwards, P.J.B.; Filichev, V.V.; Jameson, G.B. Effects of pressure and pH on the physical stability of an i-motif DNA structure. Chem. Phys. Chem. 2019, 20, 1567–1571. [Google Scholar] [CrossRef]
  103. Somkuti, J.; Molnar, O.R.; Smeller, L. Revealing unfolding steps and volume changes of human telomeric i-motif DNA. Phys. Chem. Chem. Phys. 2020, 22, 23816–23823. [Google Scholar] [CrossRef]
  104. Rashin, A.A.; Iofin, M.; Honig, B. Internal cavities and buried waters in globular proteins. Biochemistry 1986, 25, 3619–3625. [Google Scholar] [CrossRef]
  105. Liang, J.; Dill, K.A. Are proteins well-packed? Biophys. J. 2001, 81, 751–766. [Google Scholar] [CrossRef] [Green Version]
  106. Chalikian, T.V.; Sarvazyan, A.P.; Breslauer, K.J. Partial molar volumes, expansibilities, and compressibilities of α,ω-aminocarboxylic acids in aqueous solutions between 18 and 55 °C. J. Phys. Chem. 1993, 97, 13017–13026. [Google Scholar] [CrossRef]
  107. Chalikian, T.V.; Sarvazyan, A.P.; Funck, T.; Breslauer, K.J. Partial molar volumes, expansibilities, and compressibilities of oligoglycines in aqueous solutions at 18–55 °C. Biopolymers 1994, 34, 541–553. [Google Scholar] [CrossRef]
  108. Lee, A.; Chalikian, T.V. Volumetric characterization of the hydration properties of heterocyclic bases and nucleosides. Biophys. Chem. 2001, 92, 209–227. [Google Scholar] [CrossRef]
  109. Kharakoz, D.P. Volumetric properties of proteins and their analogs in diluted water solutions. 1. Partial volumes of amino acids at 15–55 °C. Biophys. Chem. 1989, 34, 115–125. [Google Scholar] [CrossRef]
  110. Hedwig, G.R.; Hastie, J.D.; Høiland, H. Thermodynamic properties of peptide solutions. 14. Partial molar expansibilities and isothermal compressibilities of some glycyl dipeptides in aqueous solution. J. Solut. Chem. 1996, 25, 615–633. [Google Scholar] [CrossRef]
  111. Hedwig, G.R.; Jameson, G.B.; Høiland, H. The partial molar heat capacity, expansion, isentropic, and isothermal compressions of thymidine in aqueous solution at T=298.15 K. J. Chem. Thermodyn. 2011, 43, 1936–1941. [Google Scholar] [CrossRef]
  112. Chalikian, T.V. Ultrasonic and densimetric characterizations of the hydration properties of polar groups in monosaccharides. J. Phys. Chem. B 1998, 102, 6921–6926. [Google Scholar] [CrossRef]
  113. Lee, S.; Tikhomirova, A.; Shalvardjian, N.; Chalikian, T.V. Partial molar volumes and adiabatic compressibilities of unfolded protein states. Biophys. Chem. 2008, 134, 185–199. [Google Scholar] [CrossRef]
  114. Chalikian, T.V.; Totrov, M.; Abagyan, R.; Breslauer, K.J. The hydration of globular proteins as derived from volume and compressibility measurements: Cross correlating thermodynamic and structural data. J. Mol. Biol. 1996, 260, 588–603. [Google Scholar] [CrossRef] [Green Version]
  115. Kharakoz, D.P. Protein compressibility, dynamics, and pressure. Biophys. J. 2000, 79, 511–525. [Google Scholar] [CrossRef] [Green Version]
  116. Taulier, N.; Chalikian, T.V. Compressibility of protein transitions. Biochim. Biophys. Acta 2002, 1595, 48–70. [Google Scholar] [CrossRef]
  117. Chalikian, T.V.; Sarvazyan, A.P.; Plum, G.E.; Breslauer, K.J. Influence of base composition, base sequence, and duplex structure on DNA hydration: Apparent molar volumes and apparent molar adiabatic compressibilities of synthetic and natural DNA duplexes at 25 °C. Biochemistry 1994, 33, 2394–2401. [Google Scholar] [CrossRef]
  118. Tikhomirova, A.; Chalikian, T.V. Probing hydration of monovalent cations condensed around polymeric nucleic acids. J. Mol. Biol. 2004, 341, 551–563. [Google Scholar] [CrossRef]
  119. Chalikian, T.V.; Volker, J.; Srinivasan, A.R.; Olson, W.K.; Breslauer, K.J. The hydration of nucleic acid duplexes as assessed by a combination of volumetric and structural techniques. Biopolymers 1999, 50, 459–471. [Google Scholar] [CrossRef]
  120. Kharakoz, D.P. Volumetric properties of proteins and their analogs in diluted water solutions. 2. Partial adiabatic compressibilities of amino acids at 15–70 °C. J. Phys. Chem. 1991, 95, 5634–5642. [Google Scholar] [CrossRef]
  121. Doster, W.; Friedrich, J. Pressure-temperature phase diagrams of proteinsC. In Protein Folding Handbook; Buchner, J., Kiefhaber, T., Eds.; Wiley-V C H Verlag Gmbh: Weinheim, Germany, 2005; pp. 99–126. [Google Scholar]
  122. Hawley, S.A. Reversible pressure-temperature denaturation of chymotrypsinogen. Biochemistry 1971, 10, 2436–2442. [Google Scholar] [CrossRef]
  123. Ravindra, R.; Winter, R. On the temperature-pressure free-energy landscape of proteins. Chem. Phys. Chem. 2003, 4, 359–365. [Google Scholar] [CrossRef] [PubMed]
  124. Winter, R.; Lopes, D.; Grudzielanek, S.; Vogtt, K. Towards an understanding of the temperature/pressure configurational and free-energy landscape of biomolecules. J. Non-Equilib. Thermodyn. 2007, 32, 41–97. [Google Scholar] [CrossRef]
  125. Daniel, I.; Oger, P.; Winter, R. Origins of life and biochemistry under high-pressure conditions. Chem. Soc. Rev. 2006, 35, 858–875. [Google Scholar] [CrossRef]
  126. Luong, T.Q.; Kapoor, S.; Winter, R. Pressure-a gateway to fundamental insights into protein solvation, dynamics, and function. Chem. Phys. Chem. 2015, 16, 3555–3571. [Google Scholar] [CrossRef]
  127. Majhi, P.R.; Qi, J.Y.; Tang, C.F.; Shafer, R.H. Heat capacity changes associated with guanine quadruplex formation: An isothermal titration calorimetry study. Biopolymers 2008, 89, 302–309. [Google Scholar] [CrossRef]
  128. Scharnagl, C.; Reif, M.; Friedrich, J. Stability of proteins: Temperature, pressure and the role of the solvent. Biochim. Biophys. Acta 2005, 1749, 187–213. [Google Scholar] [CrossRef]
  129. Dubins, D.N.; Lee, A.; Macgregor, R.B., Jr.; Chalikian, T.V. On the stability of double stranded nucleic acids. J. Am. Chem. Soc. 2001, 123, 9254–9259. [Google Scholar] [CrossRef]
Figure 1. (a) Structure of a G-quartet with a coordinated ion; (b) Schematic representation of antiparallel, parallel, and hybrid intramolecular G-quadruplexes.
Figure 1. (a) Structure of a G-quartet with a coordinated ion; (b) Schematic representation of antiparallel, parallel, and hybrid intramolecular G-quadruplexes.
Biology 10 00813 g001
Figure 2. (a) A hemiprotonated cytosine-cytosine+ base pair; (b) Schematic representation of an i-motif structure.
Figure 2. (a) A hemiprotonated cytosine-cytosine+ base pair; (b) Schematic representation of an i-motif structure.
Biology 10 00813 g002
Figure 3. The pressure-temperature phase diagram for the stability c-MYC G-quadruplex at 50 mM CsCl and 0.1 mM KCl computed with Equation (20) from ref. [99].
Figure 3. The pressure-temperature phase diagram for the stability c-MYC G-quadruplex at 50 mM CsCl and 0.1 mM KCl computed with Equation (20) from ref. [99].
Biology 10 00813 g003
Table 1. Changes in volume, ΔV, accompanying the unfolding transitions of G-quadruplexes differing in sequence, topology, and stabilizing cation determined at temperature, T.
Table 1. Changes in volume, ΔV, accompanying the unfolding transitions of G-quadruplexes differing in sequence, topology, and stabilizing cation determined at temperature, T.
Sequence (DNA)TopologyCationT, °CΔV, cm3 mol−1
d(G2T2G2TGT-G2T2G2) (TBA) aAntiparallelK+58.1 ± 1.4−54.6 ± 4.2
d[A(G3T2A)3G3](Tel22) bAntiparallelNa+54.6 ± 0.9−38.4 ± 10.1
d[A(G3T2A)3G3](Tel22) cAntiparallelNa+40.0 ± 0.6−66 ± 3
d[A(G3T2A)3G3](Tel22) bHybridK+64.6 ± 2.2−42.7 ± 6.7
d[TGA(G3TG3TA)2A](c-MYC) dParallelK+83.4 ± 1.1−16.9 ± 1.8
d(AG3AG3CGCTG3-AG2AG3) (KIT) dParallelK+58.5 ± 0.4−6.2 ± 0.9
d(T2G4CG3C2G5C-G4T2) (VEGF) dParallelK+78.8 ± 1.1−18.1 ± 4.6
d[A3(G3T2A)3G3A2](Tel26) eHybridK+25.0−69 ± 7
d[TGA(G3TG3TA)2A](c-MYC) fParallelK+25.0−34 ± 15
a from ref. [94]; b from ref. [97]; c from ref. [92]; d from ref. [96]; e from ref. [93]; f from ref. [99].
Table 2. Changes in volume, ΔV (cm3 mol−1), determined at 25 °C, intrinsic volume, ΔVM (cm3 mol−1), solvent accessible surface area, ΔSA (Å2), thermal volume, ΔVT (cm3 mol−1), and interaction volume, ΔVI, accompanying G-quadruplex unfolding transitions.
Table 2. Changes in volume, ΔV (cm3 mol−1), determined at 25 °C, intrinsic volume, ΔVM (cm3 mol−1), solvent accessible surface area, ΔSA (Å2), thermal volume, ΔVT (cm3 mol−1), and interaction volume, ΔVI, accompanying G-quadruplex unfolding transitions.
G-QuadruplexΔVΔVMΔSAΔVT = δΔSAΔVI a
Tel22−67−233 b1230 b370−186
Tel26−69−4 c2348 c707−779
c-MYC−342 d879 d265−306
a ΔVI = ΔV − (ΔVM + ΔVT + nM+VM+); b computed based on PDB entry 143D [92]; c computed based on PDB entry 2HY9 [93]; d computed based on PDB entry 1XAV [34].
Table 3. Molar chanes in volume (ΔV) accompanying the unfolding transitions of thrombin binding aptamer (TBA) and thrombin binding aptamers with altered loops determined from the pressure dependences of the unfolding temperature, TM. The values were calculated from analysis of the effect of hydrostatic pressure on the thermal unfolding of the G-quadruplex using the Clapeyron equation. For TBA 1LC12 and TBA 2LC12, a 12-carbon methylene chain replaces the nucleobases in the respective loops. The changes in the original TBA sequence are underlined.
Table 3. Molar chanes in volume (ΔV) accompanying the unfolding transitions of thrombin binding aptamer (TBA) and thrombin binding aptamers with altered loops determined from the pressure dependences of the unfolding temperature, TM. The values were calculated from analysis of the effect of hydrostatic pressure on the thermal unfolding of the G-quadruplex using the Clapeyron equation. For TBA 1LC12 and TBA 2LC12, a 12-carbon methylene chain replaces the nucleobases in the respective loops. The changes in the original TBA sequence are underlined.
DNASequenceTM, °C eΔV, cm3 mol−1
TBA a,cd(G2T2G2TGTG2T2G2)52.6 ± 3.4−54.6 ± 4.2
TBA T3A b,cd(G2ATG2TGTG2T2G2)45.1 ± 0.1−75.5 ± 2.2
TBA G8T b,cd(G2T2G2TTTG2T2G2)47.1 ± 6.6−41.1 ± 2.4
TBA 1LC12 b,cd[G2-(CH2)12-G2TGTG2T2G2]49.0 ± 3.9−57.8 ± 8.4
TBA 2LC12 b,cd[G2T2G2-(CH2)12-G2T2G2]36.5 ± 0.6−103.4 ± 8.0
TBA a,dd(G2T2G2TGTG2T2G2)59.3 ± 2.3−12.9 ± 0.9
TBA T3A b,dd(G2ATG2TGTG2T2G2)56.7 ± 2.2−14.7 ± 4.9
TBA G8T b,dd(G2T2G2TTTG2T2G2)54.5 ± 0.7−13.2 ± 2.1
TBA 1LC12 b,dd[G2-(CH2)12-G2TGTG2T2G2] 56.6 ± 6.0−9.7 ± 5.2
TBA 2LC12 b,dd[G2T2G2-(CH2)12-G2T2G2]62.4 ± 0.3−5.6 ± 1.7
a from ref. [94]; b from ref. [101]; c the solutions contained 30 mM Tris-HCl at pH 7.0 and 100 mM KCl; d the solutions contained 30 mM Tris-HCl at pH 7.0, 100 mM KCl, and 40 wt% PEG200; e the unfolding transition at atmospheric pressure.
Table 4. Volume changes associated with the unfolding of the original and modified human telomeric G-quadruplexes in the presence of Na+ or K+ at 57 °C, ΔV57 (cm3 mol−1), presented by Li et al. [97]. The changes to the original Tel22 sequence are underlined. Values were obtained by analyzing the effect of hydrostatic pressure on the equilibrium between the folded and unfolded states of the respective G-quadruplexes assuming a two-state monomolecular reaction. The values were extrapolated to a common temperature of 57 °C which corresponds to the average TM of the G-quadruplexes studied. All samples contained 10 mM Tris at pH 7.4, 0.1 mM EDTA, and either 100 mM NaCl or 100 mM KCl.
Table 4. Volume changes associated with the unfolding of the original and modified human telomeric G-quadruplexes in the presence of Na+ or K+ at 57 °C, ΔV57 (cm3 mol−1), presented by Li et al. [97]. The changes to the original Tel22 sequence are underlined. Values were obtained by analyzing the effect of hydrostatic pressure on the equilibrium between the folded and unfolded states of the respective G-quadruplexes assuming a two-state monomolecular reaction. The values were extrapolated to a common temperature of 57 °C which corresponds to the average TM of the G-quadruplexes studied. All samples contained 10 mM Tris at pH 7.4, 0.1 mM EDTA, and either 100 mM NaCl or 100 mM KCl.
DNASequenceV, cm3 mol−1)
Na+K+
Tel22 d(AG3T2AG3T2AG3T2AG3)−38.4 ± 10.1−42.7 ± 6.7
L1AATd(AG3AATG3T2AG3T2AG3)−29.4 ± 5.6−38.0 ± 7.3
L2AATd(AG3T2AG3AATG2T2AG3)−29.8 ± 10.1−35.6 ± 8.0
L3AATd(AG3T2AG3T2AG3AATG3)−34.5 ± 1.1−27.2 ± 6.4
L1TTTd(AG3TTTG3T2AG3T2AG3)−35.2 ± 3.0−35.2 ± 3.9
L2TTTd(AG3T2AG3TTTG3T2AG3)−26.2 ± 8.0−30.5 ± 11.1
L3TTTd(AG3T2AG2T2AG3TTTG3)−38.6 ± 4.0−21.9 ± 7.9
L1AAAd(AG3AAAG3T2AG3T2AG3)−37.7 ± 9.0−37.7 ± 4.8
L2AAAd(AG3T2AG3AAAG3T2AG3)−30.2 ± 13.1−37.8 ± 1.0
L3AAAd(AG3T2AG3T2AG3AAAG3)−41.4 ± 6.8−31.7 ± 1.0
Table 5. Changes in volume, ΔV, and adiabatic compressibility, ΔKS, accompanying the unfolding transitions of i-motif structures determined at temperature, T.
Table 5. Changes in volume, ΔV, and adiabatic compressibility, ΔKS, accompanying the unfolding transitions of i-motif structures determined at temperature, T.
Sequence (DNA)pHT, °C ΔV, cm3 mol−1ΔKS, 10−4 cm3 mol−1bar−1
d(C3TA2)3C3 (Tel22-iM) a4.636~0
d(C3TA2)3C3 (Tel22-iM) b5.1545.5−11 ± 2
d(T2AC3AC3TAC3A-C3TCA) (c-MYC-iM) c5.025.0~0~0
a from ref. [102]; b from ref. [103]; c from ref. [57].
Table 6. Changes in adiabatic compressibility, ΔKS, and expansibility, ΔE, accompanying the unfolding transitions of G-quadruplexes varying in sequence, topology, and stabilizing cation at 25 °C.
Table 6. Changes in adiabatic compressibility, ΔKS, and expansibility, ΔE, accompanying the unfolding transitions of G-quadruplexes varying in sequence, topology, and stabilizing cation at 25 °C.
Sequence (DNA) TopologyCationΔKS, 10−4 cm3 mol−1bar−1ΔE, cm3 mol−1K−1
d[A(G3T2A)3G3](Tel22) aAntiparallelNa+−236 ± 200.87 ± 0.16
d[A3(G3T2A)3G3A2](Tel26) bHybridK+−332 ± 180.92 ± 0.07
d[TGA(G3TG3TA)2A](c-MYC) cParallelK+−304 ± 26
a from ref. [92]; b from ref. [93]; c from ref. [99].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chalikian, T.V.; Macgregor, R.B., Jr. Volumetric Properties of Four-Stranded DNA Structures. Biology 2021, 10, 813. https://doi.org/10.3390/biology10080813

AMA Style

Chalikian TV, Macgregor RB Jr. Volumetric Properties of Four-Stranded DNA Structures. Biology. 2021; 10(8):813. https://doi.org/10.3390/biology10080813

Chicago/Turabian Style

Chalikian, Tigran V., and Robert B. Macgregor, Jr. 2021. "Volumetric Properties of Four-Stranded DNA Structures" Biology 10, no. 8: 813. https://doi.org/10.3390/biology10080813

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop