You are on page 1of 10

3242 Biophysical Journal Volume 77 December 1999 3242–3251

Heat Capacity Effects on the Melting of DNA. 1. General Aspects

Ioulia Rouzina and Victor A. Bloomfield†


Department of Biochemistry, Molecular Biology, and Biophysics, University of Minnesota, St. Paul, Minnesota 55108 USA

ABSTRACT In this paper we analyze published data on DH and DS values for the DNA melting transition under various
conditions. We show that there is a significant heat capacity increase DCp associated with DNA melting, in the range of
40 –100 cal/mol K per base pair. This is larger than the transition entropy per base pair, DS0 ' 25 cal/mol K. The ratio of
DCp/DS0 determines the importance of heat capacity effects on melting. For DNA this ratio is 2– 4, larger than for many
proteins. We discuss how DCp values can be extracted from experimental data on the dependence of DH and DS on the
melting temperature Tm. We consider studies of DNA melting as a function of ionic strength and show that while polyelec-
trolyte theory provides a good description of the dependence of Tm on salt, electrostatics alone cannot explain the
accompanying strong variation of DH and DS. While Tm is only weakly affected by DCp, its dependence on one parameter
(e.g., salt) as a function of another (e.g., DNA composition) is determined by DCp. We show how this accounts for the stronger
stabilization of AT relative to GC base pairs with increasing ionic strength. We analyze the source of discrepancies in DH as
determined by calorimetry and van’t Hoff analysis and discuss ways of analyzing data that yield valid van’t Hoff DH. Finally,
we define a standard state for DNA melting, the temperature at which thermal contributions to DH and DS vanish, by analyzing
experimental data over a broad range of stabilities.

INTRODUCTION
DNA melting is the process of separating the two strands deviations from a standard temperature, thermal contribu-
wound in a double helix into two single strands. Conceptu- tions to DH and TDS cancel each other, when summed to
ally, this phenomenon is similar to the much-studied protein give the free energy DG. But over a broader temperature
unfolding. In both cases the macromolecule cooperatively range, the different functional forms of the enthalpy and
loses its secondary structure and exposes a large fraction of entropy can result in a significant contribution of the heat
its internal surface to aqueous solution. The analogy goes capacity to the overall transition free energy. It can some-
further, because in both cases the macromolecular second- times dominate the stability of proteins and can even lead to
ary structure has marginal stability under physiological con- the striking phenomenon of cold denaturation (Franks,
ditions. This means that the transition free energy is so small 1995; Makhatadze and Privalov, 1994).
that even moderate changes in environmental conditions can Although they are commonplace in protein denaturation,
shift the equilibrium between the two states. This marginal heat capacity effects are rarely considered important in
macromolecular stability does not imply that the interac- DNA thermal denaturation. In fact, it is usually assumed
tions that hold the double helix or a native protein together that DH and DS for DNA melting are essentially indepen-
are weak. Rather, it is the result of the compensation of two dent of the temperature (Breslauer et al., 1986; Grosberg
large quantities: a favorable enthalpy DH and unfavorable and Khokhlov, 1994; Klump, 1988; Privalov et al., 1969).
entropy DS. The reason for this assumption, which we show below to be
The stability of a protein is not only determined by direct erroneous, is primarily historical. Differential scanning cal-
interactions between its atoms or the conformational free- orimeters in the 1970s and 1980s were not able to detect
dom of its backbone and side chains. Another very impor- DCp directly in the differential Cp versus T melting curves,
tant factor is the increase in hydrophobicity as the buried on a background of sloping baselines and significant noise.
amino acid side chains are exposed to water in the unfolded Moreover, the DCp value, if any, constituted less than 1% of
state. Hydrophobicity is not really an interaction; rather it is its overall change in the course of the melting transition,
the ability of nonpolar residues to affect the structure and making it unsurprising that it escaped quantitation. There-
fluctuations of the aqueous solvent. This is most character- fore it was assumed that, being below the detection level,
istically manifested in a large increase in the heat capacity DCp was not important for the transition thermodynamics.
DCp of the system, resulting in a strong temperature depen-
At the same time, when DH and DS values obtained in the
dence of the transition enthalpy and entropy. At small
same studies were presented as a function of the transition
temperature, a significant positive DCp became apparent.
In this paper we analyze published data on DH and DS
Received for publication 21 May 1999 and in final form 17 August 1999.
values for the DNA melting transition under various con-
Address reprint requests to Dr. Victor A. Bloomfield, Department of
Biochemistry, University of Minnesota, 1479 Gortner Avenue, St. Paul,
ditions. We show that there is a significant heat capacity
MN 55108. Tel.: 612-625-2268; Fax: 612-625-5780; E-mail: victor.a. increase associated with DNA melting, in the range of
bloomfield-1@tc.umn.edu. 40 –100 cal/mol K per base pair. “Significant” means that
© 1999 by the Biophysical Society DCp is larger than the transition entropy per base pair,
0006-3495/99/12/3242/10 $2.00 DS0 ' 25 cal/mol K. On a mass basis, DCp is less for DNA
Rouzina and Bloomfield Heat Capacity Effects on DNA Melting 1 3243

than for proteins (Sturtevant, 1977), but it is the ratio of DCp where T0 5 DH0/DS0 is the melting temperature in the
to DS0 that determines the importance of heat capacity reference state. The quantities denoted by d refer to pertur-
effects on melting. For DNA this ratio is 2– 4, larger than for bations from standard conditions, such as might be intro-
many proteins. Whether this means that hydrophobic effects duced by a change in salt concentration.
are important in DNA melting, or whether there is some These equations can conveniently be written in terms of
other explanation for the large heat capacity change, re- reduced variables:
mains to be seen.
The structure of this paper is as follows. In the next DH 5 DH0@1 1 gH 1 at#, (4)
section we discuss how DCp values can be extracted from
DS 5 DS0@1 1 gS 1 aln~1 1 t!#, (5)
experimental data on the dependence of DH and DS on the
melting temperature Tm. In the third section we consider where
studies of DNA melting as a function of ionic strength. We
show that while polyelectrolyte theory provides a good a 5 DCp/DS0, (6)
description of the dependence of Tm on salt, electrostatics
alone cannot explain the accompanying strong variation of t 5 ~T 2 T0!/T0, (7)
DH and DS. In the fourth section we show that while Tm is
and
only weakly affected by DCp, its dependence on one pa-
rameter (e.g., salt) as a function of another (e.g., DNA gH 5 dH/DH0, gS 5 dS/DS0. (8)
composition) is determined by DCp. This explains the stron-
ger stabilization of AT relative to GC base pairs with In practice the relative deviations from standard conditions
increasing ionic strength. In the fifth section we address the of t, gH, and gS are all much less than 1. The smallness of
question of whether DH and DS as determined by calorim- these quantities leads to the expansion for the free energy to
etry and van’t Hoff analysis are equivalent and assess the leading order in the reduced variables,
relative error in determination of DH in the latter. In the
sixth section we attempt to find the “standard state” for DG 5 DH0@gH 2 gS 2 ~1 1 gS!t 2 ~a/2!t2#. (9)
DNA melting, i.e., the temperature at which thermal con- The melting temperature, determined by the condition
tributions to DH and DS vanish, by analyzing experimental DG 5 0, is
data over a broad range of stabilities. Our summary and
conclusions are presented in the seventh section.
In the accompanying paper (Rouzina and Bloomfield,
1999) we apply this analysis to the extensive compilations
S
tm 5 t0m 1 2
a/2 0
D
t ,
1 1 gS m
(10)

of data on the thermal melting of oligonucleotides of de- where


fined sequence, the aim of which has been to tabulate
standard values of DG, DH, and DS for various nearest- gH 2 gS
t0m 5 or T0m 5 T0~1 1 gH 2 gS! (11)
neighbor base pairs. While the free energies determined in 1 1 gS
various studies generally agree with each other, there is little
consensus on the enthalpy and entropy values. The most is the melting temperature if DCp 5 0, or a 5 0. Equations
striking differences are between the results obtained on 10 and 11 are correct to first order in small quantities.
oligomeric versus polymeric DNA. We show that no special The key feature of these equations is that DH and DS vary
concepts are needed to explain these results, if the temper- to first order as at, which, because a . 1, is a stronger
ature dependence of the thermodynamic parameters is prop- temperature variation than that of DG or Tm, which vary
erly taken into account. only as the smaller quantities t or g. Examples of such
behavior from several experimental studies on DNA melt-
ing, in which equilibrium was altered by varying the ionic
strength I, are presented in Fig. 1. It is clear that the
THERMODYNAMICS OF DNA MELTING WITH
moderate (;10%) variation in free energy, as compared to
HEAT CAPACITY CHANGE
the large (;40%) variation in enthalpy and entropy, is the
Basic formulation result of mutual compensation of significant heat capacity
contributions.
The changes in free energy DG, enthalpy DH, entropy DS,
and heat capacity DCp are related to temperature T and to
each other by the familiar equations Determination of DCp and DS0 from
experimental data
DG 5 DH 2 TDS, (1)
General thermodynamic principles indicate that DCp at Tm
DH~T! 5 DH0 1 dH 1 DCp~T 2 T0!, (2) should be given equivalently by either of the two slopes
­DH/­Tm or ­DS/­ln Tm. However, when the data of Fig. 1
DS~T! 5 DS0 1 dS 1 DCp ln~T/T0!, (3) are plotted as a function of Tm, as in Fig. 2, the enthalpy
3244 Biophysical Journal Volume 77 December 1999

FIGURE 1 Thermodynamic parameters for melting of polymeric B-


DNA double helix. Enthalpy DH(Tm) (A), entropic component of the FIGURE 2 The same data and symbols as in Fig. 1, but presented as a
transition free energy T37DS(Tm) (B), free energy of melting DG 5 function of melting temperature Tm, rather than solution ionic strength.
DH(Tm) 2 T37DS(Tm) (C), corresponding to physiological temperature T 5 Note significantly stronger DH as compared to T37DS dependence. The
378C 5 310 K, as a function of solution ionic strength, I, in monovalent DG(37) dependence is even weaker and is quite similar for all of the cases
salt. DH(Tm) and DS(Tm) were obtained by calorimetry in all studies except studied.
for Karapetian et al. (1990), where differential ligand binding method was
employed. F, Calf thymus DNA, (42% GC), in Na2SO4, at pH 6.8
(Gruenwedel, 1974). ■, Poly(A1U) and poly(A12U) in Na1 and K1
salts, at pH 7 (Krakauer and Sturtevant, 1968). Œ and , C. perfringens and
31% GC and M. lysodeictikus 71% GC DNA, respectively, in NaCl at pH
7 (Karapetian et al., 1990). }, T2 DNA, 35% GC, in NaCl, at pH 7 ­DS ­d S
(Privalov et al., 1969). Note similar scales but the different ranges of the 5 1 DCp . (13)
panels. Clearly the experimental DG values vary significantly from one
­ln Tm ­ln Tm
study to another. This can be related to different DNA sequences and
Thus the experimental slopes have contributions from the
slightly different solution conditions. Independently of these differences
variation of DG with salt in each case was less then 1 kcal/mol, while it was thermal dependence of the perturbations dH and dS, as well
#3 kcal/mol for DH and #2 kcal/mol for T37DS(Tm). as from the common DCp.
We can relate the two slopes by noting that at Tm,
DG(Tm) 5 0, and this remains the case if solution conditions
slope is always larger (see Table 1). From Eqs. 2 and 3, we and temperature are adjusted simultaneously to remain at
obtain DG 5 0, so that
­DH ­dH ­DG~Tm! ­DH~Tm! ­DS~Tm!
5 1 DCp (12) 5 2 2 DS 5 0 (14)
­Tm ­Tm ­Tm ­Tm ­ ln Tm
Rouzina and Bloomfield Heat Capacity Effects on DNA Melting 1 3245

TABLE 1 Experimental slopes and derived DS0 and DCp parameters (in cal/mol K per base pair) for melting of various
genomic DNAs
­S/
DNA ­H/­Tm ­ ln Tm DS0 DCp Reference
ct, 42%GC 65.3 40.4 24.9 55.3 Gruenwedel (1974)
poly(AU) 59.5 35.5 24.0 49.9 Krakauer and Sturtevant (1968)
ct, pH 7 41.0 21.0 20.0 33.0 Shiao and Sturtevant (1973)
T2, 35%GC 30.0 3.2 26.8 19.3 Privalov et al. (1969)
C. perf., 31%GC 80.5 55.2 25.3 70.4 Karapetian et al. (1990)
M. lys., 71%GC 90.3 65.2 25.1 80.3 Karapetian et al. (1990)
DNA stability and Tm were altered by varying the ionic strength between 0.001 and 1 M in pH 7 buffers.

or magnitudes of 70 –100 cal/mol K per base pair but are


different by less than 5%. In this case the small difference
­DH ­DS Tm 2 T0 between the two slopes comes from the large negative
2 5 DS0 1 dS 1 DCp 5 DS. (15)
­Tm ­ ln Tm T0 contribution to the melting entropy due to the low Tm of
The difference between the two slopes always equals the oligomeric relative to polymeric DNA (see the third term in
transition entropy. The individual slopes contain contribu- the middle equality of Eq. 15).
tions from DCp as well as from DS, the latter being parti- Knowledge of the actual partitioning of the perturbation
tioned between the slopes in the same proportion p by which between enthalpy and entropy requires a physical model. In
the perturbation dG is partitioned between enthalpy and the next section we study the particular case of perturbation
entropy: DH 5 pDG, TDS 5 2(1 2 p)DG. That is, by varying the ionic strength.

­DH~Tm!
5 pDS 1 DCp , IONIC STRENGTH DEPENDENCE OF
­Tm DNA MELTING
(16)
­DS~Tm! An understanding of the effect of ionic strength on DNA
5 2~1 2 p!DS 1 DCp .
­ ln Tm melting temperature has been one of the major accomplish-
ments of polyelectrolyte theory. However, as we shall show
If the perturbation is purely enthalpic, e.g., in the case of
in this section, although polyelectrolyte theory accounts
DNA composition change, then p 5 1 and the apparent
very well for the dependence of DG and Tm on ionic
enthalpy slope is larger than DCp. On the other hand, if the
strength I, it can explain almost nothing about the depen-
perturbation is purely entropic, e.g., in the case of variation
dence of DH and DS. We couch our treatment, for the sake
of oligomer strand concentration, then dH 5 0, p 5 0, and
of simplicity, in terms of counterion condensation theory
the enthalpy slope reflects only the heat capacity change.
(Manning, 1975, 1978), but the same results emerge from
Thus the accurate determination of DCp from experimental
various applications of Poisson-Boltzmann theory (Bond et
enthalpy and entropy slopes requires knowledge of p, i.e.,
al., 1994; Frank-Kamenetskii et al., 1987).
how the perturbation is partitioned into its components.
In counterion condensation theory, the polyelectrolyte is
Unfortunately, because there are only two slopes and three
modeled as a cylindrical rod with length per unit charge b.
unknowns (DS, DCp, and p), this partitioning cannot be
A key parameter that characterizes the electrostatic behavior
performed uniquely. An estimate of DCp must then be
is the ratio j of b to the Bjerrum length lB:
obtained as the arithmetic average of the enthalpy and
entropy slopes; this estimate will be most adequate if the j 5 lB/b, (17)
slopes are close to each other.
A number of experimental results for the slopes and the where
derived DS and DCp are summarized in Table 1, in which e2
Tm was manipulated by varying the salt concentration. We lB 5 , (18)
ekBT
see from this table that all of the entropies are close to
DS0 5 25 6 2 cal/mol K. (It is interesting that this is very
close to the value estimated from the number of degrees of TABLE 2 Temperature dependence of oligomer melting
freedom frozen upon double helix formation (Cantor and Sequence DCp
Schimmel, 1980; DeVoe and Tinoco, 1962; Klump, 1988).) CCGC 98
The heat capacity changes are in the range DCp ' 30 –100 ACCGGp 81
cal/mol K, so that a (Eq. 6) is generally in the range of 2– 4. CCGGUp 73
In contrast, data obtained for low concentrations of short ACCGGUp 75
oligomeric DNA (Table 2) shows values of DS # 5 cal/mol CCGGAp 55
K. That is, the enthalpy and entropy slopes have absolute Petersheim and Turner (1983). In all cases, DS # 5 cal/mol K.
3246 Biophysical Journal Volume 77 December 1999

e is the electron charge, e is the solvent dielectric constant, Values of dGel, dHel, and TdSel calculated in this way are
and kB is the Boltzmann constant. Because of the different plotted in Fig. 3 B and may be compared with typical
linear charge densities of helix and coil, they territorially experimental data (Gruenwedel, 1974) in Fig. 3 A. The
bind different numbers of counterions, so that upon the experimental values were obtained by subtracting from the
helix-coil transition a number Dn of counterions is released total thermodynamic quantities DG, DH, and TDS their
per DNA phosphate charge. This leads to the free energy values at I 5 1 M, a procedure that yields the salt-dependent
difference between helix (H) and coil (C), part of the thermodynamic parameters. It is evident that the
calculated and experimental values of dGel are in close
dGel 5 kBTDn ln~I!, (19) agreement. However, the experimental dependence of DH
where on salt is much stronger than that calculated, while the salt
dependence of TDS even has a slope of opposite sign.
1 1 This discrepancy becomes even more pronounced when
Dn 5 2 . (20)
jC jH the data are replotted as a function of melting temperature,
as is done in Fig. 4, using Eq. 21 with T0 5 377.5 K. The
Equation 19 is strictly valid only in the limit of zero ionic root of the discrepancy is made clear by substituting p 5
strength, but for polyelectrolytes with the high charge den- n 2 1 into Eq. 16. If DCp 5 0, the purely polyelectrolyte
sity of DNA it holds up to at least 0.1 M (Rouzina and components of the slopes in Fig. 4 would be 10 cal/mol K
Bloomfield, 1996). If this equation is substituted into Eq. for dHel and 215 cal/mol K for TdSel, while the experimen-
11, it results in the expression for the dependence of Tm on I, tal values are 65.2 and 40.4 cal/mol K, respectively. Clearly,

S
Tm 5 T0 1 1
RDn ln~I!
DS0 D
, (21)
the difference is due to a heat capacity change of nonpoly-
electrolyte origin of substantial magnitude, ;55 cal/mol K.
Analyzing other data in a similar way (Table 1) similarly
where we have converted to molar quantities by replacing yields nonelectrostatic contributions to DCp of 30 –100 cal/
kB with the gas constant R. Using the values bH 5 0.17 nm,
bC 5 0.34 nm, and lB 5 0.714 nm, we find that Dn 5 0.24.
Substituting DS0 5 25 cal/mol K we get Tm 5 T0[1 1 0.044
log(I)], which is essentially identical to the Tm 5 377.5[1 1
0.045 log(I)] obtained from a summary of experimental
results for DNAs of normal composition (Blake and Del-
court, 1998).
To obtain the polyelectrolyte contributions to the en-
thalpy and entropy, we differentiate dGel with respect to
temperature according to the standard equations:
­~dGel/T! ­~dGel!
dH 5
el
, dS 5 2
el
. (22)
­~1/T! ­T
When dGel (Eq. 19) is divided by T, the only remaining T
dependence is in Dn through the dependence on T of lB. But
lB (Eq. 18) varies as the reciprocal of eT, and it is commonly
assumed that e is inversely proportional to T. If this were the
case, dGel/T would be independent of T, leading to dHel 5
0 and dSel 5 2dGel/T. This is, in fact, the common assump-
tion for DNA and other polyelectrolytes—that electrostatic
effects are purely entropic.
However, accurate measurements of e for water over a
wide range of T (Eisenberg and Kauzmann, 1964; Weast
and Astle, 1979) yield the more complex behavior
e~T! 5 e*~T*/T!n, (23)
FIGURE 3 Electrostatic contribution to thermodynamic parameters of
where e* 5 78.54 at T 5 298 K and n 5 1.4. Manipulation DNA melting as a function of ionic strength, I. (A) Experimental (–■–)
of the above equations then yields points (Gruenwedel, 1974) were obtained as a difference between the
measured values of thermodynamic parameters at the given salt and at I 5
dHel 5 ~n 2 1!dGel, TdSel 5 2~2 2 n!dGel. (24) 1 M. (B) Theoretical (——) dependencies are calculated as described in the
text with DCp 5 55 cal/mol K. Note that the polyelectrolyte theory predicts
Thus for helix-coil transitions in DNA, dHel ' 0.4dGel and a much weaker dH(Tm) dependence and the wrong sign of the slope of
TdSel ' 0.6dGel, where dGel ' 163 log(I) cal/mol per base dS(Tm) dependence as compared to the experimental behavior of these
pair at room temperature. quantities.
Rouzina and Bloomfield Heat Capacity Effects on DNA Melting 1 3247

contribution to the heat capacity cannot have a significant


effect on the thermodynamics of DNA melting.

BASE COMPOSITION DEPENDENCE OF


DNA MELTING
Our discussion of heat capacity effects on DNA melting can
be used to gain insight into the base composition depen-
dence of Tm. It is well known that GC-rich DNA melts at a
higher temperature than AT-rich DNA, because GC base
pairs have three hydrogen bonds, while AT base pairs have
only two. It is less well known, and not well understood,
that this difference in stability becomes smaller with in-
creasing ionic strength.
An early survey led to the empirical equation for the
dependence of Tm on mole fraction XGC and ionic strength
I (Frank-Kamenetskii, 1971),

Tm 5 355.4 1 36.0XGC 1 ~18.3 2 7.04XGC!log~I!, (26)


and a more recent summary (Blake and Delcourt, 1998)
gave an only moderately different equation

Tm 5 360.31 1 34.47XGC 1 ~20.15 2 6.52XGC!log~I!.


(27)
From Eq. 27 we calculate, for example, that the difference
in Tm between d(GC) and pure d(AT) is 65°C in 1025 M
salt, but only 35°C in 1 M salt.
If we try to understand this result using the simple poly-
electrolyte result (Eq. 21), we must conclude that either the
change in the number of counterions bound, Dn, or the
intrinsic entropy of melting, DS0, is different for AT and GC
base pairs. However, neither of these seems a very satisfac-
tory explanation.
The assumption that Dn differs for different base pairs is
commonly made (Blake and Delcourt, 1998; Santalucia,
1998) and is interpreted as a difference in linear charge
density of the various bases in the melted form (see Eq. 20).
FIGURE 4 Electrostatic contribution to thermodynamic parameters of This interpretation has even been extended to obtain infor-
DNA melting as a function of Tm. These are the same data as in Fig. 3, but mation on the structure of the single-stranded DNA as a
replotted using the relationship between Tm and log(I) given by Eqs. 3–9 function of composition (Korolev et al., 1994). However,
with T0 5 377.5 K. Note the inability of polyelectrolyte theory to repro-
duce the experimental behavior of the enthalpic and entropic components
the polyelectrolyte effect reflects only the charge spacing
of the transition free energy. averaged over a distance greater than the Debye length,
which can include tens or hundreds of base pairs at low salt.
There is no reason to expect that the charge spacing will be
a simple arithmetic average of the spacings for different
mol K. Thus there is clear evidence from ionic strength types of base pairs, because the secondary structure of
variation experiments of a significant increase in the heat single-stranded DNA, if any, should have some cooperat-
capacity of DNA upon melting. ivity. Attempts to encompass the fine-grained structure of
Using Eqs. 23 and 24, we find that the general polyelec- single-stranded DNA are beyond the scope of mean-field
trolyte contribution DCp is polyelectrolyte theory.
­dHel ­dSel dGel The assumption that DS0 differs significantly among base
DCelp 5 5 5 ~n 2 1!~2 2 n! . (25) pairs also is problematic. The basic physical assumption
­T ­ln T T
that the two kinds of base pairs differ primarily in their
With the parameters for aqueous DNA solutions given binding enthalpy, not their entropy, is supported by a large
above, this yields DCel
p 5 0.55 cal/mol K, which is small body of experimental data obtained primarily on polymeric
compared to DS0 even in low salt, so that the polyelectrolyte DNA (Blake and Delcourt, 1998; Grosberg and Khokhlov,
3248 Biophysical Journal Volume 77 December 1999

1994). On the other hand, some thermodynamic data, ob- Comparing this with Eq. 26 gives T0 5 391.4 K and g0H 5
tained primarily on oligomeric DNA (see Santalucia, 1998, 0.092, while comparison with Eq. 27 gives T0 5 394.8 K
and references therein), showed significant variations in the and g0H 5 0.087. These values for polymeric DNA are
entropies of melting from one nearest-neighbor base pair to strikingly similar to those obtained by analyzing extensive
another. Differences between the oligomeric and polymeric data on oligomers (Santalucia, 1998) (T0 5 397.5 K and
DNA melting parameters will be discussed subsequently. g0H 5 0.123), a strong confirmation of the basic correctness
Here we show that the intrinsic entropies of melting that of the approach. Furthermore, all three estimates of T0 are
enter Eq. 21 are the same for AT and GC pairs, but their very close to the value of 397 K obtained from the fit to the
thermal contributions, DCp(Tm 2 T0), can be different, completely different data analyzed by Petrushka and Good-
because of the strong dependence of Tm on the type of base man (1995).
pair. It is the heat capacity change that is responsible for the The result gH ' 0.1 is important both because it confirms
behavior summarized in Eqs. 26 and 27. the internal consistency of our treatment and because it
Equation 21 for Tm is equivalent to Eq. 11, which ne- enables an estimate of the real “chemical” enthalpy differ-
glects the heat capacity contribution contained in the more ence between GC and AT base pairs: dH 5 DH0g0H 5
complete Eq. 10. This zeroth-order approximation is suffi- 9.93 3 0.1 ' 1 kcal/mol. This value is considerably smaller
cient for the estimate of the average salt dependence, but it than the energy of a single hydrogen bond. It reflects the
is inadequate when finer details, such as the composition fact that the hydrogen bond is not completely lost upon
dependence of the slope, are considered. Taking into ac- melting, but rather is replaced by a weaker bond with water.
count the complete expression, we obtain Combining Eqs. 28, 31, and 33, we predict the slope:

­Tm RT0Dn ­tm ­Tm RT0Dn


5 2.3 , (28) 5 2.3 @1 2 g0H~XGC 2 1!~a 2 1!#, (34)
­ log I DS0 ­gS ­log~I! DS0
where we assumed that the polyelectrolyte perturbation is which can be compared to experiment. Using the already
primarily entropic, i.e., determined values of the parameters g0H, T0, and DS0, we
estimate Dn 5 0.36 and a 5 4.02 from Eq. 26, and Dn 5
RDn 0.37 and a 5 3.81 from Eq. 27. Both estimates for Dn are
gS 5 2.3 log~I!. (29)
DS0 comparable to, but somewhat larger than, the simplest poly-
electrolyte value 0.24. The fitted value of a ' 3– 4 agrees
According to Eqs. 10 and 11, very well with the direct calorimetric measurement a '

S D
­tm ­t0m a 2– 4. This is convincing evidence that the stronger salt
5 12 t0 dependence of the melting temperature of AT sequences
­ gS ­ gS 1 1 gS m (30) relative to GC is due to the heat capacity-induced increase

52
1 1 gH
~1 1 gS! F
2 12a
g H 2 gS
~1 1 gS!2
.G of the transition entropy.
The lesson to be drawn from this result extends beyond
the context in which it was derived. While Tm itself is only
In the case gS ,, 1, appropriate for the whole experimental weakly perturbed by the heat capacity correction, its deriv-
range of interest, this derivative is ative with respect to salt as a function of DNA composition
is determined by a . 1. This should also be the case for any

S D
­tm
­ gS gS30
5 2~1 1 gH!~1 2 agH! < 21 1 gH~a 2 1!,
two parameters that affect Tm, e.g., the salt dependence of
Tm as a function of cosolvent concentration or concentration
(31) of some ligand. The general relationships of this section,
especially Eq. 30, allow one to calculate the derivative of Tm
where we have neglected the term ag2H because of the with respect to any parameter, given the dependence of the
smallness of gH. The right-hand side of Eq. 31 should be perturbation free energy components gH and gS on that
compared to its value 21 2 gH for the case a 5 0. We see parameter.
that, although the magnitude of the slope remains near
unity, the heat capacity effect when a . 1 changes the sign
of the variation of ­tm/­gS with DNA composition. RECONCILING VAN’T HOFF AND
We can cast our results in the form of Eqs. 26 and 27 by CALORIMETRIC ENTHALPIES
assuming that the enthalpic perturbation is linear in XGC:
Enthalpies of reaction may be measured by calorimetry,
gH 5 g ~XGC 2 1!.
0
(32) which monitors the heat of the reaction directly, or by van’t
H
Hoff analysis of the temperature dependence of the free
Substituting this into Eq. 11 and evaluating at I 5 1 M energy of the reaction according to the general expression
where gS ' 0, we obtain
­~DG/T!
Tm~XGC , I 5 1 M! 5 T0@1 1 g0H~XGC 2 1!#. DHvH 5 . (35)
(33) ­~1/T!
Rouzina and Bloomfield Heat Capacity Effects on DNA Melting 1 3249

The van’t Hoff method is convenient and can be imple- Despite this clear evidence of significant heat capacity
mented in many different ways. For DNA melting, any effects, temperature-dependent DH and DS values are con-
experiment that monitors the fraction of melted base pairs as ventionally averaged and compared with the corresponding
a function of temperature is sufficient to obtain DHvH. quantities obtained by data fitting to Eq. 36. These two
However, when DHvH values from such analyses are determinations of average DH and DS are generally in good
compared to the calorimetric values DHcal, they often dis- agreement, providing proof of the ability of the van’t Hoff
agree with each other. This has been conventionally attrib- analysis to capture the thermal variation of DH and DS in a
uted to the different experimental conditions used in such consistent way. Ironically, this is a standard way to produce
studies. The development of highly precise microcalorim- DH and DS values, which are then reported as temperature-
eters has made it possible to perform both calorimetric and independent parameters (Breslauer et al., 1986; Krug et al.,
van’t Hoff measurements within a single experiment (Liu 1976; Owczarzy et al., 1997; Plum et al., 1995; Santalucia,
and Sturtevant, 1995, 1997; Naghibi et al., 1995). When the 1998).
authors of these studies performed direct comparisons of the
DHvH and DHcal values for a large number of different
reactions, they found that the two quantities never agreed DEFINING THE STANDARD STATE FOR
and that the discrepancies between them often approached DNA MELTING
100%. No reasonable interpretation of this observation was
suggested, except for “unaccounted participants in the We have seen that DNA melting enthalpy and entropy are
reaction.” strongly influenced by solvent, DNA composition, and tem-
Our analysis in terms of heat capacity effects, which perature. To be able to compare data from different studies,
contribute substantially to DH and DS individually but and to come to definite conclusions about the influence of a
largely cancel out in DG, gives insight into this discrepancy. particular parameter, we need to define the standard state.
If the precise T dependence of DG is known, then the At first it might seem that it could be chosen arbitrarily, e.g.,
derivative, Eq. 35, should yield a temperature-dependent as some particular set of conditions for which the most data
enthalpy similar to the calorimetric value. But in practice, are available. Then the enthalpy and entropy components
any appreciable experimental noise in DG significantly ob- will be arbitrarily split into the standard parts DH0 and DS0,
scures the weak curvature of its T dependence and leads to and the perturbations dH and dS.
an incorrect apparent DHvH. The relative error in this quan- However, the presence of the heat capacity change DCp
tity should be on the order of atm, which can easily reach introduces a temperature dependence into both components.
;50% because of the large a 5 2– 4. The same conclusion If the T dependencies were linear, then shifting the reference
was reached in a recent Monte Carlo study (Chaires, 1997). temperature T0 would simply result in redefining the stan-
In fact, reasonable temperature-dependent enthalpies can dard values DH0 and DS0. But because of the different
be obtained by van’t Hoff analysis. A good example is a functional dependences of DH and DS on Tm (Eqs. 2 and 3),
study of melting of short DNA oligomers (Petersheim and this is not true. Instead, changing the reference temperature
Turner, 1983), which depends strongly on the concentration by the relative amount Dt results in a relative variation of
of single strands in solution, Ct. An extra 2RT ln(Ct) of DS0 on the order ;atDt, which depends on temperature and
binding entropy per mole of oligomer results in the variation can be significant for a . 1, t ' 1, Dt ' 1.
of melting temperature with Ct: Strictly speaking, thermal contributions to DH and DS are
both zero only at a single temperature T0, which is the true
1 R ln~Ct! DS standard temperature for melting. This hypothetical temper-
52 1 . (36) ature can be determined from experimental data for DH and
Tm DH DH
DS if these quantities are available over a wide range of
The slope of a plot of 1/Tm versus ln(Ct) yields DHvH of stabilities, so that the nonlinearity of DS with Tm becomes
melting per oligomer. apparent. This is the case for the data analyzed by Petrushka
A complementary way to measure the same quantity is to and Goodman (1995), who found strong nonlinear correla-
perform traditional van’t Hoff analysis of the DNA optical tions between experimental values of DH and DS for regular
melting curves. This procedure, performed at several values base-pair doublets and a number of mismatched and mod-
of Ct, yields DH and DS as a function of Ct. When these ified base pairs. Some of these noncanonical base-pair dou-
measured DH and DS values are plotted as a function of Tm blets had very low or even negative stabilities; i.e., they
rather than Ct, the strong temperature dependence of both decreased the stability of the normal oligomer. This pro-
quantities becomes apparent. The slopes ­DH/­Tm and vided the required broad range of stabilities of individual
­DS/­ ln Tm are similar to each other and close to the base pairs.
calorimetrically measured DCp, which ranges between 52 The authors fitted DS versus DH data at I 5 1 M to the
and 95 cal/mol K in different oligomers. This is in close phenomenological expression
agreement with the DCp values obtained in studies of poly-
meric DNA melting as a function of solution ionic strength DH
DS 5 , (37)
(Table 1). T* 1 DH/b9
3250 Biophysical Journal Volume 77 December 1999

with the constants T* 5 273 K and b 5 80 cal/mol K. (We We will not speculate here on the physical origin of T0.
have changed the original notation of Petrushka and Good- We simply note that this is a very high temperature, above
man (1995).) No physical interpretation of this equation was the boiling point of water and thus not normally accessible,
given, but it was noted that Eq. 37 is equivalent to the which is needed as a parameter for the analysis of experi-
temperature dependence of the enthalpy, mental data in the regular temperature range. Its formal
meaning is that temperature at which the thermal contribu-
DH 5 b~Tm 2 T*!. (38) tions to DH and DS vanish.
This is equivalent to the pair of equations Except for the “true” standard temperature T0, the other
conditions for the standard state can be chosen at our
DH 5 DH0 1 ~DS0 1 DCp!~Tm 2 T0!, convenience. Setting the standard Tm 5 T0 gets rid of
(39) thermal components and fixes the ratio T0 5 (DH0 1
DS 5 DS 1 DCp ln~Tm/T !
0 0
dH)/(DS0 1 dS). Then, choosing dS 5 0, for example, for
the case of polymeric DNA in 1 M salt, with DS0 5 25
if
cal/mol K, fixes the value of DH0 5 9927 cal/mol, which is
DS0 1 DCp characteristic of the most stable GC base pair. Therefore the
b 5 DS0 1 DCp and T0 5 T* . (40) standard state we have chosen corresponds to polymeric
DCp
d(GC) in 1 M aqueous salt solution.
Taking the standard value DS0 5 25 cal/mol K, we find

DCp 5 55 cal/mol K, T0 5 397 K,


(41) SUMMARY AND CONCLUSIONS
DH0 5 DS0/T0 5 9927 cal/mol,
Analysis of a large body of experimental data suggests that
in good agreement with other fits of these parameters in there is a significant heat capacity increase, DCp ' 30 –100
previous sections. The fit with parameters of Eq. 41 in Fig. cal/mol K per base pair associated with DNA melting.
5 is at least as good as the original empirical fit (Petrushka Being larger than the standard entropy of DNA melting
and Goodman, 1995), and is obviously different from a simple DS0 5 25 cal/mol K per base pair, DCp dominates changes
linear dependence in the region of maximum stability. in transition enthalpy and entropy induced by any variation
It is worth noting that individual fits of DS versus Tm and in solution conditions. The heat capacity effect on the tran-
DH versus Tm according to Eq. 39 are not nearly as good as sition free energy and melting temperature is less pro-
for the DH versus DS plot. This reflects contributions to DH nounced, but it is responsible for a number of subtle phe-
and DS for different nearest-neighbor base pairs that are not nomena. In particular, the heat capacity change determines
correlated with the nearest-neighbor doublet identity, but the variation of Tm with one parameter (e.g., solution ionic
are induced by random solution variations. Nevertheless, strength) as a function of another (e.g., DNA composition).
because the contributions are always of thermal origin, they This accounts for the stronger stabilization by salt of AT
are always coupled through DCp, as in Eq. 39. relative to GC base pairs.
Any heat capacity effects on DNA melting thermody-
namics can be analyzed with the help of the general expres-
sions in Eqs. 1– 8. The new feature we have introduced is an
explicit account of the perturbation enthalpy dH and entropy
dS, which allows direct linking between Tm, external pa-
rameters, and DCp.
We have discussed practical aspects of the determination
of DCp from experimental data and have shown that the
experimental slopes ­DH/­Tm and ­DS/­ln Tm always con-
tain both entropic and heat capacity contributions and differ
from each other by the value of DS. Appreciation of the
strong temperature dependence of DH and DS helps to
interpret the apparent differences in enthalpy values ob-
tained in van’t Hoff and calorimetric experiments. Even
though DH and DS are often statistically coupled (Krug et
al., 1976; Owczarzy et al., 1997; Plum et al., 1995), because
FIGURE 5 Entropy-enthalpy correlation in DNA melting transition. F, DS is determined as DH/Tm, Tm is measured independently,
Experimental data (Petrushka and Goodman, 1995) for 10 standard and and so the coupling should indeed reflect the coupled ther-
several modified NN doublets. - - - -, Linear approximation from the high- mal contributions to both quantities, which cancel out in Tm.
stability region; ——, Empirical fit by the authors Petrushka and Goodman
(1995); – – –, Our fit, assuming a temperature-independent heat capacity In the particular case of the effect of salt on double-helix
increase DCp 5 55 cal/mol K and standard values DS0 5 25 cal/mol K and stability, we have used polyelectrolyte theory to analytically
T0 5 397 K. calculate the enthalpy, entropy, and heat capacity changes in
Rouzina and Bloomfield Heat Capacity Effects on DNA Melting 1 3251

the course of DNA melting. We have shown that the enthalpic in Na2SO4 solutions of varying ionic strength. Biochim. Biophys. Acta.
340:16 –30.
part of the polyelectrolyte free energy of the transition comes
Karapetian, A. T., P. O. Vardevanian, and M. D. Frank-Kamenetskii. 1990.
from the peculiar temperature behavior of the dielectric con- Enthalpy of helix-coil transition of DNA: dependence on Na1 concen-
stant of water. The calculated polyelectrolyte heat capacity tration and GC content. J. Biomol. Struct. Dyn. 8:131–138.
change is too small to account for the experimental tempera- Klump, H. H., editor. 1988. Conformational transitions in nucleic acids. In
ture dependence of DH and DS of DNA melting. Biochemical Thermodynamics. M. N. Jones, editor. Elsevier, Amster-
dam. 100 –144.
In the accompanying paper (Rouzina and Bloomfield,
Korolev, N. I., A. P. Vlasov, and I. A. Kuznetsov. 1994. Thermal denaturation
1999), we apply this general treatment to a critical analysis of Na- and Li-DNA in salt-free solutions. Biopolymers. 34:1275–1290.
of the melting thermodynamics of oligonucleotides of de- Krakauer, H., and J. Sturtevant. 1968. Heats of the helix-coil transitions of
fined sequence, where we show that it resolves a number of the poly A-poly U complexes. Biopolymers. 6:491–512.
apparent inconsistencies between data from various sources. Krug, R. R., W. G. Hunter, and R. A. Grieger. 1976. Enthalpy-entropy
After this paper was submitted for publication, a paper by compensation. I. Some fundamental statistical problems associated with
the analysis of van’t Hoff and Arrhenius data. J. Phys. Chem. 80:
Breslauer and co-workers (Chalikian et al., 1999) appeared 2335–2341.
that provides remarkable confirmation of the ideas proposed Liu, Y. F., and J. M. Sturtevant. 1995. Significant discrepancies between
here. By direct, high-precision calorimetric measurements van’t Hoff and calorimetric enthalpies. II. Protein Sci. 4:2559 –2561.
of helix-coil transitions of five polymeric duplexes, they Liu, Y. F., and J. M. Sturtevant. 1997. Significant discrepancies between
found positive heat capacity changes with an average value van’t Hoff and calorimetric enthalpies. III. Biophys. Chem. 64:121–126.
DCp 5 64.6 6 21.4 cal/deg-mol. This is in striking agree- Makhatadze, G., and P. Privalov. 1994. Hydration effects in protein un-
folding. Biophys. Chem. 51:291–309.
ment with the values reported in this paper.
Manning, G. S. 1975. On the application of polyelectrolyte limiting laws to
the helix-coil transition of DNA. V. Ionic effects on renaturation kinet-
ics. Biopolymers. 15:1333–1343.
We are grateful to Prof. Kenneth Breslauer for a helpful critique of the
manuscript. Manning, G. S. 1978. The molecular theory of polyelectrolyte solutions
with applications to the electrostatic properties of polynucleotides.
This research was supported in part by National Institutes of Health Q. Rev. Biophys. 11:179 –246.
research grant GM28093. Mrevlishvili, G. M., T. D. Mdzinarashvili, N. O. Merteveli, and G. R.
Kakabadze. 1992. Thermal capacity of DNA in the native and denatured
states. Biofizika. 37:859 – 860.
REFERENCES Naghibi, H., A. Tamura, and J. Sturtevant. 1995. Significant discrepancies
between van’t Hoff and calorimetric enthalpies. Proc. Natl. Acad. Sci.
Blake, R. D., and S. G. Delcourt. 1998. Thermal stability of DNA. Nucleic USA. 92:5597–5599.
Acids Res. 26:3323–3332.
Owczarzy, R., P. M. Vallone, F. J. Gallo, T. M. Paner, M. J. Lane, and
Bond, J., C. Anderson, and M. T. Record, Jr. 1994. Conformational A. S. Benight. 1997. Predicting sequence-dependent melting stability of
transitions of duplex and triplex nucleic acid helices: thermodynamic short duplex DNA oligomers. Biopolymers. 44:217–239.
analysis of effects of salt concentration on stability using preferential
interaction coefficients. Biophys. J. 67:825– 836. Petersheim, M., and D. H. Turner. 1983. Base-stacking and base-pairing
contributions to helix stability: thermodynamics of double-helix forma-
Breslauer, K. J., R. Frank, H. Blocker, and L. A. Marker. 1986. Predicting tion with CCGG, CCGGp, CCGGAp, ACCGGp, CCGGUp, and AC-
DNA duplex stability from the base sequence. Proc. Natl. Acad. Sci. CGGUp. Biochemistry. 22:256 –263.
USA. 83:3746 –3750.
Petrushka, J., and M. F. Goodman. 1995. Enthalpy-entropy compensation
Cantor, C. R., and P. R. Schimmel. 1980. Biophysical Chemistry. Part III.
in DNA melting thermodynamics. J. Biol. Chem. 270:746 –750.
The Behavior of Biological Macromolecules. W. H. Freeman, San Fran-
cisco. Plum, G. E., A. P. Grollman, F. Johnson, and K. J. Breslauer. 1995.
Chaires, J. B. 1997. Possible origin of differences between van’t Hoff and Influence of the oxidatively damaged adduct 8-oxodeoxyguanosine on
calorimetric enthalpy estimates. Biophys. Chem. 64:15–23. the conformation, energetics, and thermodynamic stability of a DNA
duplex. Biochemistry. 34:16148 –16160.
Chalikian, T. V., J. Volker, G. E. Plum, and K. J. Breslauer. 1999. A more
unified picture for the thermodynamics of nucleic acid duplex melting: Privalov, P. L., O. B. Ptitsyn, and T. M. Birshtein. 1969. Determination of
a characterization by calorimetric and volumetric techniques. Proc. Natl. stability of the DNA double helix in an aqueous medium. Biopolymers.
Acad. Sci. USA. 96:7853–7858. 8:559 –571.
DeVoe, H., and I. Tinoco, Jr. 1962. The stability of helical polynucleotides: Rouzina, I., and V. A. Bloomfield. 1996. Competitive electrostatic binding
base contributions. J. Mol. Biol. 4:500 –517. of charged ligands to polyelectrolytes: planar and cylindrical geometries.
Eisenberg, D., and W. Kauzmann. 1964. The Structure and Properties of J. Phys. Chem. 100:4292– 4304.
Water. Clarendon Press, Oxford. Rouzina, I., and V. A. Bloomfield. 1999. Heat capacity effects on the
Frank-Kamenetskii, M. D. 1971. Simplification of the empirical relation- melting of DNA. 2. Analysis of nearest-neighbor base pair effects.
ship between melting temperature of DNA, its GC content and concen- Biophys. J. 77:3252–3255.
tration of sodium ions in solution. Biopolymers. 10:2623–2624. Santalucia, J. 1998. A unified view of polymer, dumbbell, and oligonucle-
Frank-Kamenetskii, M. D., V. V. Anshelevich, and A. V. Lukashin. 1987. otide DNA nearest-neighbor thermodynamics. Proc. Natl. Acad. Sci.
Polyelectrolyte model of DNA. Sov. Phys. Uspekhi. 151:595– 618. USA. 95:1460 –1465.
Franks, F. 1995. Protein destabilization at low temperatures. Adv. Protein Shiao, D. D. F., and J. M. Sturtevant. 1973. Heats of thermally induced
Chem. 46:105–139. helix-coil transitions of DNA in aqueous solutions. Biopolymers. 12:
Grosberg, A. Y., and A. R. Khokhlov. 1994. Statistical Physics of Mac- 1829 –1836.
romolecules. AIP Series in Polymers on Complex Materials. R. Larson Sturtevant, J. 1977. Heat capacity and entropy changes in processes in-
and P. Pincus, editors. AIP Press, New York. volving proteins. Proc. Natl. Acad. Sci. USA. 74:2236 –2240.
Gruenwedel, D. W. 1974. Salt effects on the denaturation of DNA. III. A Weast, R. C., and M. J. Astle, editors. 1979. CRC Handbook of Chemistry
calorimetric investigation of the transition enthalpy of calf thymus DNA and Physics, 60th ed. CRC Press, Boca Raton, FL.

You might also like