Sei sulla pagina 1di 8

Current Pharmaceutical Biotechnology, 2005, 6, 215-222 215

1389-2010/05 $50.00+.00 2005 Bentham Science Publishers Ltd.


Application of Solution Calorimetry in Pharmaceutical and
Biopharmaceutical Research
P.G. Royall
1,
*
and S. Gaisford
2
1
Department of Pharmacy, Kings College London, Franklin-Wilkins Building, 150 Stamford Street, London, SE1 9NH
and
2
Department of Pharmaceutics, School of Pharmacy, University of London, 29/39 Brunswick Square, London,
WC1N 1AX, UK
Abstract: In solution calorimetry the heat of solution (
sol
H) is recorded as a solute (usually a solid) dissolves in an
excess of solvent. Such measurements are valuable during all the phases of pharmaceutical formulation and the number of
applications of the technique is growing. For instance, solution calorimetry is extremely useful during preformulation for
the detection and quantification of polymorphs, degrees of crystallinity and percent amorphous content; knowledge of all
of these parameters is essential in order to exert control over the manufacture and subsequent performance of a solid
pharmaceutical. Careful experimental design and data interpretation also allows the measurement of the enthalpy of
transfer (
trans
H) of a solute between two phases. Because solution calorimetry does not require optically transparent
solutions, and can be used to study cloudy or turbid solutions or suspensions directly, measurement of
trans
H affords the
opportunity to study the partitioning of drugs into, and across, biological membranes. It also allows the in-situ study of
cellular systems. Furthermore, novel experimental methodologies have led to the increasing use of solution calorimetry to
study a wider range of phenomena, such as the precipitation of drugs from supersaturated solutions or the formation of
liposomes from phospholipid films. It is the purpose of this review to discuss some of these applications, in the context of
pharmaceutical formulation and preformulation, and highlight some of the potential future areas where solution
calorimetry might find applications.
Key Words: Solution calorimetry, polymorphism, amorphous content, pharmaceuticals, dissolution, enthalpy of transfer.
1. INTRODUCTION
Many active pharmaceutical ingredients (APIs) exist in
the solid-state or are formulated into solid dosage forms.
This confers several advantages; stability in the solid-state is
usually greater than in the liquid state, solids are easier to
transport, package and store and the majority of medicines,
at least in the United Kingdom, are formulated as tablets as
the oral route is both quick and convenient. However, solid-
state formulations are usually slower-acting than liquid
medicines containing the same active, because the first event
that must occur following administration is that the API must
dissolve in a suitable biological fluid (such as saliva or
gastric juices). On a molecular level, this requires the
intramolecular forces holding the solid lattice together to be
overcome (endothermic) and the formation of new inter-
actions with the solvent (generally exothermic). This event is
(usually) rate-limiting, and thus controls the observed
dissolution rate of the API from the formulation.
The intra-molecular forces in a crystalline material (the
lattice energy) will vary if the material exhibits polymor-
phism, (i.e. if the molecules can pack in more than one
crystalline formation), the most stable polymorph having the
highest lattice energy. It is therefore imperative that the
crystal form of any solid-state API is known, since the
dissolution rates of each polymorph will be different.
*Address correspondence to this author at the Department of Pharmacy,
Kings College London, Franklin-Wilkins Building, 150 Stamford Street,
London, SE1 9NH, UK; Tel: +44 (0)20 7848 4780; Fax: +44 (0)20 7848
4800; E-mail: paul.royall@kcl.ac.uk
Moreover, over time all the metastable polymorphs will
convert to the stable polymorph, which may have a
disastrous effect on the efficacy of the formulated product. A
classic example of this is provided by chloramphenicol
palmitate, which has three polymorphs, termed A, B and C
[1]. The A polymorph is unstable and is not allowed to be
included in any medicine; similarly, the C polymorph has
virtually no appreciable bioavailability and its percentage in
the final product must be strictly controlled. Any formulator
of chloramphenicol palmitate must therefore know, and be
able to quantify, the amount of the B polymorph in any
formulation.
In addition to polymorphism, solid APIs may also exhibit
amorphicity (a lack of long-range molecular order). Because
amorphous materials dont have a lattice energy and are
essentially unstable (over time they will recrystallise to a
crystalline form) they usually have appreciably faster
dissolution rates than their crystalline equivalents, which
makes them especially suited for formulation into fast-acting
medicines

. The accidental inclusion of amorphous content


in what is otherwise presupposed to be a crystalline material
is also a hazard associated with processing of solid-state

It is often stated that amorphous materials have higher solubilities than their
crystalline counterparts; since solubility is an equilibriumstate, this cannot be true.
What may often happen is that thefast dissolution rateof theamorphous formresults
in the production of a super-saturated solution which over time, although not
necessarily in the time frame involved in drug administration and absorption, will
precipitate to forma saturated solution. As such, it should be stated that amorphous
materials may exhibit higher apparent solubilities than their crystallineequivalents.
216 Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 Royall and Gaisford
materials (during, say, milling or compression steps) [2-3].
Although the percentage of amorphous content introduced in
this way is usually low (of the order of 1% w/w) its location
primarily on the surface of what are usually small particles
gives it a disproportionate control on the properties of the
material [4]. It is clear, then, that the detection and quanti-
fication of (often small) amorphous contents is of the utmost
importance during the characterization of a solid pharma-
ceutical.
One technique that offers the sensitivity to differentiate
between polymorphs as well as to detect low (typically
around 1% w/w) amorphous contents is solution calorimetry.
The principle of solution calorimetry is simple; the heat
change when a small quantity of a solid or liquid sample is
dispersed into a (relatively) large volume of solvent is
measured (either directly or by measuring a temperature
change which is subsequently converted to a heat change).
For a pure material dissolving into a solvent this results in
the enthalpy of solution (
sol
H), an enthalpy change that
reflects contributions from the bonds broken when the
crystal lattice breaks apart (
lattice
H) and from the bonds
formed when the molecules are solvated (
solvation
H). This
can be expressed as;
H H H
solvation lattice sol
+ = Equation 1
Depending on the relative magnitudes of
lattice
H and

solvation
H, the heat of solution can have a positive (endo-
thermic) or negative (exothermic) sign. Careful experimental
design (discussed in more detail below) is required to ensure
that the measured enthalpy does not contain erroneous
contributions from other sources, such as dilution (liquid
samples), friction effects (ampoule breaking and sample
stirring) and sample-sample interactions (if an ideal solution
is not formed), or simply as a corollary of poor experimental
design or execution. Since
sol
H differs for each polymorph
of a solid sample, it is clear that solution calorimetry
provides valuable information that allows the detection and
quantification of polymorph content.
Moreover, since
sol
H contains information on the attrac-
tive forces holding a solid together, solution calorimetry can
provide fundamental information on the nature of multi-
component systems. Such systems include solid-dispersions,
polymeric systems and proteins freeze-dried with carbo-
hydrates. As such, solution calorimetry is a powerful tool
that can be used to investigate a wide range of pharma-
ceutical systems, and it is the purpose of this review to
discuss some of these applications in the context of the
preparation and formulation of novel pharmaceuticals and
biopharmaceuticals.
2. THE PRINCIPLES OF SOLUTION CALORIMETRY
A solution calorimeter records the heat change when a
sample (solid or liquid) is dissolved in a large volume of
solvent. There are two types of solution calorimeter design
commercially available; instruments that operate on a semi-
adiabatic principle (i.e. that record a temperature change
upon reaction) and instruments that operate on a heat-
conduction principle (i.e. that record a power change directly
upon reaction). The sensitivities of these instruments, and
hence the quantities of solute and solvent required for
experiment, vary considerably (typically, semi-adiabatic
instruments are less sensitive and require much larger sample
volumes).
In a carefully constructed experiment, the enthalpy of
solution,
sol
H, is measured directly. Care must be taken to
ensure that other contributing heat sources are accounted for.
These may include effects from breaking the ampoule and
sample stirring (most easily compensated for by subtracting
a blank experiment from the sample experiment), changes in
the solvent activity because of solute dissolution, changes in
the rate of evaporation of solvent into the headspace and
changes in volume upon mixing of the two phases. The last
three effects can be assumed to be negligible if an ideal
(dilute) solution is formed; if this is not the case, a fact that
would be confirmed by obtaining different values of
sol
H
with different amounts of sample, then the data must be
extrapolated back to infinite dilution to give
sol
H

.
2.1. Types of Solution Calorimeter
Semi-adiabatic and heat-conduction instruments operate
on different principles and, while these have been discussed
in detail elsewhere (See for example, 5-7) it is worth briefly
considering these differences here because they impact upon
the subsequent discussion of the data.
2.1.1. Semi-Adiabatic Solution Calorimeters
In an ideal adiabatic calorimeter there is no heat ex-
change between the calorimetric vessel and its surroundings.
This is usually attained by placing an adiabatic shield around
the vessel. Thus any change in the heat content of a sample
as it reacts causes either a temperature rise (exothermic
processes) or fall (endothermic processes) in the vessel. The
change in heat is then equal to the product of the temperature
change and an experimentally determined proportionality
constant (or calibration constant, ). The proportionality
constant is usually determined by electrical calibration.
Thus;

q
T =
Equation 2

=
t
T
d
d
Equation 3
where represents power. Ideally, the value of returned
after calibration should equal the heat capacity of the
calorimeter vessel (C
v
, the vessel including the calorimetric
ampoule, block, heaters, thermopiles and the sample) but in
practice losses in heat mean the value may differ slightly.
However, assuming the losses are the same for both sample
and reference, the power value returned will be accurate. It is
also the case that C
v
varies depending on the heat capacity of
the sample being studied, which affects the measuring
sensitivity of the instrument. Thus, a small heat capacity
results in a large rise in temperature for a given quantity of
heat and, consequently, better sensitivity. However, calori-
meters with low heat capacities are more sensitive to
environmental temperature fluctuations and have lower
baseline stabilities. Any calorimeter design therefore results
in a compromise between baseline stability and measurement
sensitivity.
Application of Solution Calorimetry in Pharmaceutical Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 217
In practice, true adiabatic conditions are difficult to
achieve and there is usually some heat-leak to the surround-
ings. If this heat-leak is designed into the calorimeter (such
as is the case with the SolCal, an example of a commercial
semi-adiabatic calorimeter, Thermometric AB, J rflla,
Sweden) the system operates under semi-adiabatic (or
isoperibol) conditions and corrections must be made in order
to return accurate data. These corrections are usually based
on Newtons law of cooling (the most common being the
method of Regnault-Pfaundler, discussed below).
In the case of the SolCal (the principles apply to all
similar solution calorimeters), at the start of an experiment
the instrument is held above or below the temperature of its
thermostatting bath (typically by up to 200 mK). With time
the instrument will approach the temperature of the
thermostatting bath; data capture is initiated when this
approach becomes exponential (this assumption is a
necessary precursor to employing the heat-balance equations
used to calculate the heat evolved or absorbed by the system
contained within the vessel). Thus, the response due to
dissolution, and any electrical calibrations (usually two are
performed; one before and one after the break to ensure the
heat capacity of the system has remained constant), must be
performed before the instrument reaches thermal equilibrium
with the bath. In practice, this limits the technique to
studying events that, ideally, reach completion in less than
30 min.
Upon completion of an event in a solution calorimeter a
quantity of heat will have been recorded. As noted above, the
heat will be given by the product of the temperature change
and the calibration constant (which in the ideal case is the
heat capacity of the vessel). This interpretation assumes that
the measured temperature change arises solely from the
event occurring in the vessel; in practice, other events, such
as ampoule breaking, stirring and heat-leakage, all contribute
to the temperature change of the vessel. For accurate data
analysis these effects must be removed from the observed
temperature change (T
obs
) to give T
corr
, the temperature
change that would have occurred under ideal conditions.
Thus;
adj corr obs
T T T + = Equation 4
where T
adj
is defined as the temperature change arising
from all the other contributing events in the vessel. Usually
the method of Regnault-Pfaundler, which is based on the
dynamics of the break, is used to determine the value of T
adj
[8]. In this case;

T
adj
=
1

(T

T)dt
t
start
t
end

Equation 5
where T is the temperature of the vessel and its contents at
time t, T

is the temperature that the vessel would attain after


an infinitely long time period, t
start
and t
end
are the start and
end times of the experiment respectively and is the time
constant of the instrument. Note that T

is effectively the
value of T at t

and is commonly described as the steady-
state temperature of the vessel. The time constant has units
of seconds and can also be expressed as;
k

=
Equation 6
where k is the heat exchange coefficient of the vessel.
The values of T

and are calculated by analysis of the


baseline regions immediately preceding and following the
break. These baseline sections will be approaching the
temperature of the surrounding heat-sink exponentially and
are described by;

T = T

+(T
0
T

)e
t

Equation 7
The data are then fitted to Equation 7 using a least-
squares minimising routine to return values for T

and .
Once these are known, T
adj
can be calculated. This value is
then used to calculate T
corr
for the sample break and also for
the two electrical calibrations (T
corr, calibration
). There should
be no significant difference in the T
corr, calibration
values
determined for the two calibrations and an averaged value is
used. The calibration constant is then determined from;
n calibratio corr,
n calibratio
T
Q

=
Equation 8
the heat-change for the break is then easily determined;
corr reaction
. T Q = Equation 9
2.1.2. Heat-Conduction Calorimeters
A heat-conduction calorimeter is surrounded by a heat-
sink, which acts to maintain the system at a constant tem-
perature. Between the vessel and the heat-sink is a thermo-
pile wall. Any heat released or absorbed upon reaction is
quantitatively exchanged with the heat-sink. The thermopiles
generate a voltage signal that is proportional to the power
flowing across them; this signal is amplified, multiplied by
the cell constant (determined through electrical calibration)
and recorded as power versus time. An isothermal system is
not limited to reaction processes that reach completion
within 30 min, as semi-adiabatic instruments are, because it
is always (essentially) in equilibrium with its surrounding
heat-sink. Furthermore, the greater measuring sensitivity of
the thermopiles (as opposed to the thermisters used in semi-
adiabatic instruments) means that smaller sample masses can
be used.
A further consideration of the use of isothermal
instruments concerns dynamic correction of the data. The
aim of dynamic correction is to remove the effect of the
thermal inertia inherent in any calorimeter (i.e. the delay
between heat being released by the sample and that heat
causing a measurable voltage to be generated by the
thermopiles) and it is principally used for short-term events,
typically in titration experiments. However, since one
outcome from dynamic correction is an improvement in the
signal to noise (S/N) ratio of the data it offers the potential to
reduce the standard deviation of dissolution experiments,
because peak areas can be determined with greater precision.
In the case of a typical instrument (the 20 ml micro
reaction ampoule, Thermometric AB, Sweden), dynamic
correction is achieved by application of a modified form of
the Tian equation;
218 Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 Royall and Gaisford
2
R
2
2 1
R
2 1 R C
d
d
. .
d
d
). (
t
P
t
P
P P + + + =
Equation 10
where P
R
and P
C
are the raw and corrected powers
respectively and
1
and
2
are termed the first and second
time constants of the instrument (for a further discussion of
the derivation and use of Equation 1 see, for example, [9]).
The time constants,
1
and
2
, are determined by a least
squares analysis of data following an electrical calibration
(and are hence not user defined). It is important to note that a
number of assumptions are made in order to derive Equation
10 (the major assumption being that there are no temperature
gradients within the sample) and that it only approximates
the true dynamic delay inherent to the instrument. The
corrected data so produced, while much more closely
resembling the true response of the sample, therefore often
contain artefacts, such as overshoots where both endo- and
exothermic events are indicated even though it is known that
only one event is occurring in the sample. In principle, these
artefacts could be removed by altering the values of
1
and
2
but this is difficult in practice. It is therefore easier to use
corrected data to determine reaction enthalpies only and to
note that the use of such data to elucidate kinetic information
must be undertaken with caution. It can be shown that the
total net heat change recorded for both dynamically
corrected and raw data, in the ideal case, is the same [9].
2.1.3. Sample Solvent Mixing
There are several methods by which the solute and
solvent can be introduced. The solute and solvent can be
held in separate chambers of the same cell - rotating the
sample cell end-over-end mixes the materials and initiates
the interaction (this is the system employed by, for instance,
the C80, Setaram, France); the solute can be held in a sealed
glass ampoule which is broken into the solvent to initiate
reaction (used in the 2225 precision solution calorimeter,
Thermometric AB, Sweden); the solute can be held in
reusable metal canisters which are broken into the solvent
(used in the 20 ml micro solution ampoule, Thermometric
AB, Sweden); or liquid solutes/solutions can be titrated into
a cell containing solvent (such as the VP-ITC, Microcal Inc.,
USA).
2.2. Calibration
Calibration is vital to ensure that instruments are opera-
ted properly and are functioning correctly, that data from
different instruments or different laboratories are comparable
and, perhaps most importantly when considering pharma-
ceutical samples, that the data obtained are validated and can
be incorporated into regulatory documents. Most calori-
meters are calibrated using an electrical substitution method.
In this case, a resistance heater (usually located under the
sample cell) produces a known amount of heat, which is
measured by the instrument. The heat output recorded is
adjusted until the measured and expected heats are the same
(by multiplying the raw data signal by a constant value; the
cell constant). However, it is debatable whether an electrical
calibration truly represents an accurate method by which to
validate instrument performance, primarily because the
source and rate of heat generation will be different from that
caused by a chemical interaction. In this sense, chemical
standards offer a better alternative and have been the subject
of some discussion in the literature.
There are a number of requirements imposed upon a test
reaction; it should be robust, simple to perform, require
commonly available materials that require no special
preparation prior to use and it should be applicable across a
range of instrumentation [10]. A number of chemical test
reactions for solutions calorimeters have been proposed and
discussed, including the dissolution of Tris in 0.1 M HCl [5,
11], the dissolution of KCl or NaCl in water [12-14] and the
dissolution of propan-1-ol in water [15]. The dissolution of
sucrose in water can also be used [16-17], although this is
not currently recognised as a test reaction.
Of these systems, the dissolution of KCl into water is
usually recommended because it is robust, easy to perform
and a standard reference material is available from NIST (the
National Institute for Standards and Technology, USA); the
enthalpy of solution of the NIST certified KCl into water is
17.584 0.017 kJ mol
-1
[13]. However, the use of KCl is not
without drawbacks; principally, the value of
sol
H varies as a
function of the concentration achieved after dissolution,
because of the effect of the enthalpy of dilution (
L
); thus,
the certified value for the NIST reference material of 17.584
kJ mol
-1
applies only if a final concentration of 0.111 mol kg
-
1
is attained in the calorimetric vessel. This corresponds to a
molar ratio of water to KCl of 500 to 1 and is often written
as
sol
H (500 H
2
O, 298.15 K). If measurements are per-
formed under different conditions, then the value obtained
(nH
2
O, 298.15 K) must be corrected to that which would
have been recorded at 500 H
2
O, in order to draw com-
parison. These corrections are explained in the certification
certificate supplied with the NIST sample [13], although the
data supplied there apply only to experiments performed
where n varies from 100 to 1000.
The effects of KCl concentration on
sol
H have been
studied extensively by Kilday [12], who corrected the
observed enthalpy values over a range of water ratios (n =
500 to 10000) to account for the enthalpy of dilution; this
resulted in a value for
sol
H

, the enthalpy of solution at


infinite dilution (
sol
H

=17.241 0.018 kJ mol


-1
).
Modern solution calorimeters often use microgram
samples and are capable of detecting very small powers; one
consequence of this is that it is not possible to perform the
KCl experiment under the NIST certification conditions
because the heat generated would be of a magnitude
sufficient to saturate the amplifiers. Because of this several
recent studies have been conducted looking at the applicabi-
lity of test reactions to modern solution calorimeters [7, 18].
The outcome from these studies appears to be that if KCl is
to be used then it is a smaller, and easier, correction to the
enthalpy of solution at infinite dilution, but that sucrose may
offer a better, and cheaper, alternative, especially for heat-
conduction instruments that use very small (mg) samples.
3. APPLICATIONS
3.1. Polymorphism
As mentioned above, if a drug is prepared in the solid-
state then it is highly likely it will exhibit polymorphism and
it is imperative that the polymorph(s) present in a sample are
Application of Solution Calorimetry in Pharmaceutical Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 219
known before formulation, such that the resulting medicine
performs within its pharmacopoeial specifications. For new
drug entities, the pressure to bring a product to market
usually means that the developing company doesnt have the
time to characterise the relative bioavailability of each
polymorphic form; usually, the most stable form is selected
for development and the formulation is tailored to ensure its
stability. This formulation strategy ensures compliance with
the guidance provided by the International Commission on
Harmonisation (ICH) on the selection of solid forms of drugs
[19]. Similarly, for Abbreviated New Drug Applications
(ANDA) the sponsor must provide evidence that the pro-
posed generic product and the original product are pharma-
ceutically equivalent and bioequivalent [20]; clearly, not
having any control or knowledge of the polymorphic forms
present would render submission of such data impossible.
Solution calorimetry allows the direct measurement of
the lattice energy of a sample, which alters with each poly-
morphic form of the material and, although other analytical
techniques may be used to discriminate between polymorphs
(such as DSC, XRPD, solid-state NMR, IR and Raman
spectroscopy and microscopy [21]), the technique has three
principle advantages; firstly, the measurement is performed
directly at the storage (or, indeed, any desired) temperature,
in contrast to DSC measurements which usually need to be
extrapolated to give an indication of structure at room
temperature; secondly, the data allow comparison between
samples, giving insights into batch-to-batch or formulation
variability; and thirdly the various contributions to the
measured heat of solution can be dissected out by careful
experimentation, allowing a mechanistic analysis of the
process under investigation. The principal drawback, related
to the latter point above, is that the heat of solution is a
composite of all those processes that are occurring during
dissolution and it may sometimes be impossible to separate
the heat into its component parts.
Solution calorimetry has been used to assess the different
polymorphs of many drugs. For instance, forms I and II of
cyclopenthiazide have been shown to have comparable
enthalpies of solution (~6 kJ mol
-1
) while form III has a
higher enthalpy of solution of 15 kJ mol
-1
[22]. Similarly,
solution calorimetry has been used to characterise the three
polymorphic forms of a pre-clinical drug, Abbott-79175 [23]
and the polymorphs of an angiotensin II antagonist agent
(MK996) [24].
A study of urapidil resulted in the enthalpies of solution
of forms I and II of the drug (21.96 and 23.89 kJ mol
-1
respectively) being determined [25]. However, the authors
note that urapidil exists in an additional form (form III),
which they were unable to obtain in a pure form. By mea-
suring the heat of solution of the mixture (
mix
H), and using
DSC data to determine the percentage of each polymorph in
a sample containing form III (7.4% form I, 2.7% form II and
89.9% form III), they calculated the enthalpy of solution of
form III using the following relationship;
) . ( ) . ( ) . (
III III II II I I mix
H X H X H X H + + = Equation 11
where X
I
, X
II
and X
III
are the fractions of form I, II and III
respectively and H
I
, H
II
and H
III
are the heats of solution
of forms I, II and III respectively. It was determined that
H
III
equaled 22.98 kJ mol
-1
.
Since Hesss law allows the interpretation of
sol
H by
invoking any number of steps, as long as the starting state is
the solid and the final state is the solution, one approach to
the analysis of dissolution data is to consider the heat as
comprising two major contributions; one which represents
the breaking of the solid-state bonds and the other which
represents all the processes through which the molecules
become solvated [26]. If bond breaking is considered to be
similar to a melting process, then it can be represented by

f
H
298
, the enthalpy of formation of a supercooled liquid at
298 K (assuming the dissolution experiment is conducted at
298 K), while the enthalpy changes associated with mixing
and solvation can be represented by
m
H. This can be
represented as shown in Fig. 1.
Fig. (1). A description of dissolution, based on Hesss law [26].
This model has been applied to the study of the dissolu-
tion of various poly(ethylene) glycols, commonly encoun-
tered pharmaceutical excipients, in water [26], where
f
H
298
was measured using DSC data and
sol
H was measured in
the solution calorimeter. Four different molecular weight
PEGs were selected (3, 400, 6000, 10, 000 and 20, 000) and
samples were prepared with different thermal histories;
samples were quench-cooled from the melt, slow-cooled
from the melt (5
o
C min
-1
) or left untreated. It was found that

sol
H varied as a function of both molecular weight and
thermal history, although no discernable relationships could
be found. According to the model above,
m
H should be
constant for each molecular weight PEG, independent of any
thermal history, since this represents the dissolution of
individual molecules, while
f
H
298
should be highly depen-
dent on thermal history. Accordingly, the
so
H
l
values will
vary, since this parameter is the sum of
f
H
298
and
m
H.
Subtraction of the
f
H
298
values, determined from DSC data,
from the measured
sol
H values showed that, within a 5%
confidence level the
m
H values did remain constant for
each molecular weight PEG, and that the measured
variabilities in the measured heats of solution derived from
changes in
f
H
298
.
The correlation of the heats of solution of different
polymorphs of a drug with dissolution rates has also been
attempted. For instance, Terada et al. [27] showed there was
a linear correlation between the heats of solution of different
polymorphs of indomethacin and the logarithms of their
220 Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 Royall and Gaisford
initial dissolution rates (determined by the rotating disk
method). The same authors showed a similar relationship for
a range of samples of terfenadine of varying crystallinity.
In some cases, a drug will be insoluble in water, and in
order to derive some meaningful information on its poly-
morphic forms it is desirable to measure heats of solution in
two non-aqueous solvents. Again, in accordance with the
model discussed above, the value of
sol
H for each poly-
morph will alter in each solvent. However, the difference
between the heats of solution for each polymorph should be
constant irrespective of the solvent used;
Constant
T II form sol, I form sol,
= = H H H Equation 12
where
T
H is the transfer enthalpy, and is the energy
required to transfer between the two forms. Differences in

sol
H determined in each solvent will reflect changes in the
solvation interaction with that solvent.
This approach has been used to investigate the poly-
morphs of enalapril maleate, by recording the heats of
solution of forms I and II of the drug in acetone and metha-
nol [28]. The differences in the heats of solution of forms I
and II in each solvent (2.88 and 2.13 kJ mol
-1
in acetone and
methanol respectively) were the same within experimental
error, and gave
T
H
,
the energy that would be required for
form II to convert to form I. The difference in the heat of
solution for forms I and II in methanol, compared with
acetone, was 23.53 kJ mol
-1
; that is, the heat of solution in
methanol was more exothermic than in acetone. This value
was found to be consistent with the formation of a single
hydrogen bond.
Similar data, recorded using ethanol and methanol, have
been recorded for terfenadine [29], where a difference in
energy of approximately 13 kJ mol
-1
was found between
forms I and II. In this case, it was observed that
sol
H for any
solid form of terfenadine was around 4 kJ mol
-1
lower in
methanol than ethanol, indicating that polar groups on the
drug molecule play an important role in the solute/solvent
interaction.
A different approach to study water insoluble drugs is to
use a concentrated surfactant solution instead of pure water
as the solvent. For instance, the enthalpies of solution of a
number of cimetidine polymorphs have been studied using
concentrated SDS and Tween 20 solutions [30].
3.2. Determination of Degree of Crystallinity/Amorphous
Content
Solid pharmaceuticals may often not be entirely
crystalline and contain regions of amorphous content. An
amorphous material (most simply described as a material
with the structure of a liquid and the viscosity of a solid) is
characterised by a lack of any long-range crystal structure
and, because it therefore has no lattice energy, the lack of a
melting point (upon heating, an amorphous solid will
recrystallise and it is the crystalline form that subsequently
melts). Amorphous solids are thermodynamically unstable
and usually exhibit fast dissolution rates compared with their
crystalline equivalents. Pharmaceutically, this is a big advan-
tage for medicines that are required to be fast acting and
many solid drugs are prepared in the amorphous state
although, as noted earlier, knowledge of the percent amor-
phous content and its stability are vital prerequisites to
ensure the stability and continued efficacy of the medicine
upon storage.
The use of solution calorimetry to quantify the degree of
crystallinity in a solid sample is predicated on the
relationship shown in Equation 11, where the measured heat
of solution is given by the sum of the enthalpies and weight
fractions for the crystalline and amorphous states present.
The usual methodology is to prepare a calibration line of

sol
H against degree of crystallinity using a number of
known standards (usually prepared by blending the
appropriate mass quantities of wholly amorphous and wholly
crystalline material); the calibration plot should be linear and
can thus be used to determine the degree of crystallinity in
an unknown sample.
There are two potential drawbacks to using this approach.
Firstly, the sample may exhibit polymorphism, in which case
there may be more than one crystalline form present; if this
is the case then, for the calibration plot to be linear, it must
be ensured that the crystalline material used to prepare the
standards and the unknown sample contain the same
proportions of the polymorphs. Secondly, it is likely that the
interactions in a particle that has a crystalline core and
amorphous material on its surface (the likely situation for a
processed pharmaceutical) differ from those of wholly
amorphous and wholly crystalline particles, which may
result in the calibration plot producing spurious results.
The first use of solution calorimetry for the quantitative
measurement of the degree of crystallinity of pharma-
ceuticals was by Pikal et al. [31], who measured the heat of
solution of various -lactam antibiotics. Subsequent studies
include the analysis of sulphamethoxazole from different
sources [32], sucrose [16] and clathrate warfarin sodium
[16].
In many of these studies, the aim was to quantify relat-
ively large mass percentages of crystalline material, giving
heat changes easily within the detection limit of the tech-
nique. The issue of detection limits becomes more important
if the objective of the study is to assess the quantity of
amorphous material present in what is a predominately
crystalline sample because, as stated earlier, in a milled
sample amorphous material may typically only be present up
to 1% w/w. An assessment of the applicability of solution
calorimetry to study small amorphous contents in solid
pharmaceuticals was conducted by Hogan and Buckton [4],
who prepared a calibration curve for lactose between 0 and
10% w/w amorphous content in the same way as described
above. They found that the technique could quantify
amorphous content to 0.5% w/w but noted that care needed
to be taken when preparing the ampoules, because ingress of
even small amounts of humidity caused partial recrystalli-
sation of the sample before measurement.
Usually, in experiments designed to measure degrees of
crystallinity of amorphous content a solvent is selected in
which the solute is freely soluble. This ensures complete
dissolution of the sample within the time frame of the
experiment. Harjunen et al. [33] studied the dissolution of
Application of Solution Calorimetry in Pharmaceutical Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 221
lactose into saturated aqueous lactose solutions, a system
where clearly the solute would not completely dissolve.
Interestingly, they observed a linear relationship between the
amorphous content of the lactose solute and the measured
heat of solution in the saturated lactose solution (
sat
H).
Similarly, a linear relationship was found between the amor-
phous content of lactose and
sat
H in methanolic saturated
solutions of lactose [34].
3.3. Characterisation of Interactions
As discussed above, application of Hesss law allows the
construction of solution calorimetry experiments that allow
valuable information on the interaction of a solvent with a
mixed solvent system. An example would be to measure the
dissolution of a solid in a buffer and a buffered solution of
micelles. The difference in the
sol
H values recorded in the
two systems would represent the enthalpy of transfer of the
solute from the buffer to the micelle. Similarly, if a complex
solvent system is employed, such as one that matched a
biological medium for instance, it would be possible to
measure the enthalpy of interaction of a solute with the entire
medium and then each of its constituents in turn.
This approach has been used to investigate the interaction
of two solutes, propranolol HCl and mannitol, with two
simulated intestinal fluids (fasted-state and fed-state) and
Hanks balanced salt solution (HBSS) [35]. The simulated
intestinal fluids contained bile salts and lipids which formed
mixed micelles while no micelles formed in HBSS. It was
found that the two solutes exhibited endothermic heats of
solution in all solvents; however, the values for propranolol
HCl were lower in the two simulated fluids than in HBSS
while the values for mannitol were constant in all media.
Calculation of the enthalpy of transfer (
trans
H) for pro-
pranolol HCl into the micelles in the two simulated fluids
revealed an exothermic (and hence favourable) interaction
(-10.3 kJ mol
-1
for the fed-state and -2.1 kJ mol
-1
for the
fasted-state).
In an earlier study, Beezer et al. [36] calculated values of

trans
H for a series of alkoxyphenols to Escherichia coli cells,
while the combination of Caco-2 cells and simulated
intestinal fluids has been suggested as a model for studying
drug permeability through membranes [37]. It seems as if
solution calorimetry may an ideal technique by which to
monitor these interactions, especially as it is unaffected by
cloudy or turbid solutions or suspensions.
Chada et al. [38] used solution calorimetry to probe the
interactions of diclofenac sodium in cyclodextrin solutions
and water/ethanol mixtures. Tong et al. [39] used solution
calorimetry to evaluate the stability constants and enthalpy
changes associated with the formation of complexes between
2-hydroxypropyl--cyclodextrin and a group of 12 amine
compounds which all had a diphenylmethyl functional
group. They found that only terfenadine HCl formed a 1:2
complex with the -cyclodextrin, the other 11 compounds all
forming 1:1 complexes.
Solution calorimetry has also been used to measure the
enthalpy of solution of diclofenac sodium, paracetamol and
their binary mixtures [40] and to evaluate the in vitro
compatibility of amoxicillin/clavulanic acid and ampicillin/
sulbactam with ciprofloxacin [41].
3.4. Other Applications
Perlovich and Bauer-Brandl [42] have discussed the use
of heat of solution data to predict drug solubility using two
model compounds, benzoic acid and aspirin. Their data
suggest it may be possible to predict the solubility or
solvation of a drug in different media. Similarly, Willson and
Sokoloski [43], as part of a study developing a method to
rank the stability of drug polymorphs, correlated solution
calorimetry measurements to conventionally determined
solubility data. Solution calorimetry has also been employed
to study the dissolution and solvation of the model systems,
flurbiprofen and diflunisal [44] and to compare the solvation
of (+)-naproxen with three model NSAIDs (benzoic acid,
diflunisal and flurbiprofen) [45].
Recent work has shown that solution calorimetry can be
used to investigate the stability of supersaturated systems
[46]. Supersaturated systems are particularly important for
topical and transdermal formulations where the API is may
formulated above its solubility in order to maximise the
diffusional driving force for absorption [47, 48]. Clearly,
supersaturated formulations are inherently thermodynami-
cally unstable and it is hence likely that crystallisation of the
active will occur during storage with a resultant change in
drug bioavailability. The ability to measure the time taken
before recrystallisation occurs is therefore an essential
prerequisite to the development of such transdermal formula-
tions and is difficult because the formulations are often
opaque semi-solids.
Solution calorimetry allows the study of these formula-
tions because it permits the in-situ formation of saturated
solutions. Hadgraft et al. [46] have shown that by adding
ibuprofen in a co-solvent to a saturated solution of ibuprofen
it is possible to form a supersaturated drug solution in the
calorimetric ampoule. In the absence of any stabilising com-
pound the supersaturated system immediately precipitates
and a heat signal is observed. If a small concentration (0.1%
w/v) of hydroxypropyl methyl cellulose (HPMC) is added to
the system the time to recrystallisation increased to
approximately 8 minutes. At a concentration of HPMC of
0.5 % w/v the supersaturated system remained stable for
approximately 30 minutes, while at an HPMC concentration
of 1% w/v the system was stable for longer than 24 hours.
Barriocanal et al. [49] studied the formation of liposome
formation by coating the inside of a glass ampoule with
phospholipids (deposited from a chloroform solution). Upon
breaking, the phospholipids film hydrated and liposomes
formed; the solution calorimeter measured the changes in
heat associated with the processes. They found that the
formation of liposomes from egg phosphatidylcholine was
exothermic while the formation of liposomes from dimyris-
toylphosphatidylcholine was endothermic and suggest that
this difference arose from the influence of the hydrocarbon
chains predominately on the hydration process. They also
noted that the retention of small quantities of chloroform in
the phospholipids film significantly altered the enthalpy
change of liposome formation, an effect that was ascribed to
the effect of chloroform on hydration.
Solution calorimetry can also be used to study the actions
of formulations directly. For instance, Gaisford et al. [50]
222 Current Pharmaceutical Biotechnology, 2005, Vol. 6, No. 3 Royall and Gaisford
used solution calorimetry to follow the acid neutralisation of
magnesium trisilicate mixture BP. They showed that of the
three active components in the mixture (magnesium carbo-
nate, magnesium trisilicate and sodium bicarbonate), magne-
sium carbonate contributed most to the action of the product,
followed by sodium bicarbonate. The results suggested that
magnesium trisilicate did not neutralise a significant quantity
of acid by itself but that it extended the neutralising response
of the mixture, presumably by interacting with the
magnesium carbonate. The results provided an insight into
the mechanism of action of the product and the data would
be useful in any reformulation exercise.
SUMMARY
Solution calorimetry offers the same virtues to the
pharmaceutical scientist as other forms of calorimetry; it is
non-invasive, non-destructive and is not limited to the study
of homogeneous systems. It does not require large quantities
of sample and the data are easy to interpret, so long as care
has been taken in experimental design and execution. Solu-
tion calorimetry has classically been used during preformula-
tion, where it allows the detection and identification of
polymorphs and the quantification of crystalline or amor-
phous contents. However, its ability to study cloudy or turbid
solutions or suspensions means it is finding increasing
applications in the direct study of the interactions of drugs
with complex biological systems, where it is possible to
measure the enthalpy of transfer of a drug between phases.
Novel experimental strategies are also expanding its range of
applications, and it has been used to investigate the stability
of supersaturated systems, antacid formulations and the
formation of liposomes. Its continued use and development
will ensure it retains its place as a vital tool in pharmaceut-
ical formulation and preformulation.
REFERENCES
[1] Augiar, A.J ., Krc J r. J ., Kinkel, A.W. and Samyn, J .C. (1967) J.
Pharm. Sci., 56, 847-853.
[2] Grant, D.J .W. and York, P. (1986) Int. J. Pharm., 30, 161-180.
[3] Giron, D., Remy, P., Thomas, S. and Vilette, E. (1997) J. Thermal
Analysis, 48, 465-472.
[4] Hogan, S. and Buckton, G. (2000) Int. J. Pharm., 207, 57-64.
[5] Hill, J .O.; Ojlund, G. and Wads, I. (1969) J. Chem. Thermodyn.,
1, 111-116.
[6] Bastos, M., Bai, G., Qvarnstrm, E. and Wads, I. (2003)
Thermochimica Acta, 405, 21-30.
[7] Yff, B.T.S.; Royall, P.G.; Brown, M.B. and Martin, G.P. (2004)
Int. J. Pharm., 269, 361-372.
[8] Wads I (1966) The LKB Inst J, 13, 33-39.
[9] Randzio, S.L. and Suurkuusk, J . (1980) In Beezer AE, editor.
Biological microcalorimetry. 1st ed., London:Academic Press
p311-341.
[10] Beezer, A.E., Hills, A.K., ONeill, M.A.A., Morris, A.C., Kierstan,
K.T.E., Deal, R.M., Waters, L.J ., Hadgraft, J ., Mitchell, J .C.,
Connor, J .A., Orchard, J .E., Willson, R.J ., Hofelich, T.C., Beaudin,
J ., Wolf, G., Baitalow, F., Gaisford, S., Lane, R.A., Buckton, G.,
Phipps, M.A., Winneke, R.A., Schmitt, E.A., Hansen, L.D.,
OSullivan, D. and Parmar, M.K. (2001) Thermochimica Acta, 380,
13-17.
[11] Irving, R.J . and Wads, I. (1964) Acta. Chem. Scand., 18, 195-201.
[12] Kilday, M.V. (1980) J. Res. Nat. Bur. Stand., 85, 467-481.
[13] Uriano, G.A. (1981) National Bureau of Standards Certificate.
Standard Reference Material 1655, PotassiumChloride, KCl (cr)
for Solution Calorimetry.
[14] Archer, D.G. and Kirklin, D.R. (2000) Thermochimica Acta, 347,
21-30.
[15] Olofsson, G.; Berling, D.; Markova, N. and Molund, M. (2000)
Thermochimica Acta, 347, 31-36.
[16] Gao, D. and Rytting, J .H. (1997) Int. J. Pharm., 151, 183-192.
[17] Salvetti, G., Tognoni, E., Tombari, E. and J ohari, G.P. (1996)
Thermochimica Acta, 285, 243-252.
[18] Ramos, R., Gaisford, S., Buckton, G., Royall, P.G., Yff, B.T.S. and
ONeill, M.A.A. (2005) Int. J. Pharm., Submitted for publication.
[19] Grant, D.J .W. and Byrn, S.R. (2004) Adv. Drug Del. Rev., 56, 237-
239.
[20] Raw, A.S., Furness, M.S., Gill, D.S., Adams, R.C., Holcombe J r,
F.O. and Yu, L.X. (2004) Adv. Drug Del. Rev., 56, 397-414.
[21] Yu, L., Reutzel, S.M. and Stephenson, G.A. (1998) Pharm. Sci.
Tech. Today, 1, 118-127.
[22] Gerber, J .J .; vanderWatt, J .G. and Ltter, A.P. (1991) Int. J.
Pharm., 73, 137-145.
[23] Li, R.C.Y., Mayer, P.T., Trivedi, J .S. and Fort, J .J . (1996) J.
Pharm. Sci., 85, 773-780.
[24] J ahansouz, H., Thompson, K.C., Brenner, G.S. and Kaufman, M.J .
(1999) Pharm. Dev. Tech., 4, 181-187.
[25] Botha, S.A., Guillory, J .K. and Lotter, A.P. (1986) J. Pharm.
Biomed. Anal., 4, 573-587.
[26] Craig, D.Q.M. and Newton, J .M. (1991) Int. J. Pharm., 74, 43-48.
[27] Terada, K., Kitano, H., Yoshihashi, Y. and Yonemochi, E. (2000)
Pharm. Res., 17, 920-924.
[28] Ip, D.P.; Brenner, G.S.; Stevenson, J .M.; Lindenbaum, S.; Douglas,
A.W.; Klein, S.D. and McCauley, J .A. (1986) Int. J. Pharm., 28,
183-191.
[29] Canotilho, J .; Costa, F.S.; Sousa, A.T.; Redinha, J .S. and Leitao,
M.L.P. (1997) Thermochimica Acta, 299, 1-6.
[30] Souillac, P.O.; Dave, P. and Rytting, J .H. (2002) Int. J. Pharm.,
231, 185-196.
[31] Pikal, M.J ., Lukes, A.L., Lang, J .E. and Gaines, K. (1978) J.
Pharm. Sci., 67, 767-772.
[32] Guillory, J .K. and Erb, D.M. (1985) Pharm. Manuf., Sep, 28-33
[33] Harjunen, P., Lehto, V-P., Koivisto, M., Levonen, E., Paronen, P.
and J rvinen, K. (2004) Drug Dev. Ind. Pharm., 30, 809-815.
[34] Katainen, E., Niemel, P., Pllysaho, M., Harjunen, P., Suhonen,
J . and J rvinen, K. (2003) Eur. J. Pharm. Sci. Abstracts, 19, S36.
[35] Arnot, L.F., Minet, A., Patel, N., Royall, P.G. and Forbes, B.
(2004) Thermochimica Acta, 419, 259-266.
[36] Beezer, A.E., Gooch, C.A., Hunter, W.H. and Volpe, P.L.O. (1987)
J. Pharm. Pharmacol., 39, 774-779.
[37] Patel, N., Murray, J .G. and Forbes, B. (2001) British
Pharmaceutical Conference Abstracts Book, p51.
[38] Chadha, R., Kashid, N., Kumar, A. and J ain, D.V.S. (2002) J.
Pharm. Pharmacol., 54, 481-486.
[39] Tong, W.Q., Lach, J .L., Chin, T.F. and Guillory, J .K. (1991) J.
Pharm. Biomed. Anal., 9, 1139-1146.
[40] Chada, R., Kashid, N. and J ain, D.V.S. (2003) J. Pharm. Biomed.
Anal., 30, 1515-1522.
[41] Chada, R., Kashid, N. and J ain, D.V.S. (2004) J. Pharm. Biomed.
Anal., 36, 295-307.
[42] Perlovich, G.L. and Bauer-Brandl, A. (2003) Pharm. Res., 20, 471-
478.
[43] Willson, R.J . and Sokoloski, T.D. (2004) Thermochimica Acta,
417, 239-243.
[44] Perlovich, G.L., Kurkov, S.V. and Bauer-Brandl, A. (2003) Eur. J.
Pharm. Sci., 19, 423-432.
[45] Perlovich, G.L., Kurkov, S.V., Kinchin, A.N. and Bauer-Brandl, A.
(2004) Eur. J. Pharm. Biopharm., 57, 411-420.
[46] Hadgraft, J ., ONeill, M.A.A., Gaisford, S., Beezer, A.E., Al-
Moshey, L., Farah, F. and Auner, B. (2005) J. Drug Del. Sci. Tech.,
Submitted for publication.
[47] Lervolino, M., Raghavan, S.L. and Hadgraft J . (2000) Int. J.
Pharm., 198, 229-238.
[48] Lervolino, M., Cappello, B., Raghavan, S.L. and Hadgraft J . (2001)
Int. J. Pharm., 212, 131-141.
[49] Barriocanal, L., Taylor, K.M.G. and Buckton, G. (2004) Int. J.
Pharm., 287, 113-121.
[50] Gaisford, S., Royall, P.G. and Greig, D.T.G. (2004)
Thermochimica Acta, 417, 217-221.

Potrebbero piacerti anche