throbber
THERMODYNAMICS OF
`
`PHARMACEUTICAL
`
`SYSTEMS
`
`An Introduction for
`Students of Pharmacy
`
`Kenneth A. Connors
`
`School of Pharmacy
`University of Wisconsin—Madison
`
`OXILEEIEEQENCE
`
`A JOHN WILEY & SONS, INC., PUBLICATION
`
`000001
`
`Exhibit 1 107
`
`ARGENTUM
`
`Exhibit 1107
`ARGENTUM
`IPR2017-01053
`
`IPR2017-01053
`
`000001
`
`

`

`Copyright © 2002 by John Wiley & Sons, Inc. All rights reserved.
`
`Published by John Wiley & Sons, Inc., Hoboken, New Jersey.
`Published simultaneously in Canada.
`
`No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or
`by any means, electronic, mechanical, photocopying, recording, scanning, or otherwise, except as
`permitted under Section 107 or 108 of the 1976 United States Copyright Act, without either the prior
`written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to
`the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400,
`fax 978-750-4470, or on the web at www.copyright.com. Requests to the Publisher for permission
`should be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken,
`NJ 07030, (201) 748-6011, fax (201) 748-6008, e-mail: permcoordinator@wiley.com.
`
`For general information on our other products and services please contact our Customer Care Department
`within the US. at 877-762-2974, outside the US. at 317-572-3993 or fax 317-572-4002.
`
`Wiley also publishes its books in a variety of electronic formats. Some content that appears in print,
`however, may not be available in electronic format.
`
`Library of Congress Cataloging-in-Publication Data:
`
`Connors, Kenneth A. (Kenneth Antonio), 1932-
`Thermodynamics of pharmaceutical systems: an introduction
`for students of pharmacy / Kenneth A. Connors.
`p.
`cm.
`Includes bibliographical references and index.
`ISBN 0-471-20241-X (paper : alk. paper)
`I. Title.
`1. Pharmaceutical chemistry.
`2. Thermodymamics.
`[DNLM:
`1. Thermodymamics.
`2. Chemistry, Pharmaceutical.
`QC 311 C752t 2003]
`RS403.C665 2003
`615’.1941c21
`
`2002011151
`
`Printed in the United States of America.
`
`10987654321
`
`000002
`
`000002
`
`

`

`Thermodynamics of Pharmaceutical Systems: An Introduction for Students of Pharmacy.
`Kenneth A. Connors
`Copyright © 2002 John Wiley & Sons, Inc.
`ISBN: 0—471—20241—X
`
`
`This article is protected bycopyrighrand is provided byrhe Universny ofwisconsinr
`Madison under license from John Wiley &Sons. All riym reserved.
`
`
`
`
`10.1. SOLUBILITY AS AN EQUILIBRIUM CONSTANT
`
`The topic of solubility merits special attention because of its great importance in
`pharmaceutical systems. We can generally anticipate that a drug must be in solution
`if it is to exert its effect. Typically the type of system we encounter is a pure solid
`substance (the solute)
`in contact with a pure liquid (the solvent). We allow
`equilibrium to be achieved at fixed temperature and pressure, such that at equili-
`brium the system consists of (excess) pure solid phase and liquid solution of solute
`dissolved in solvent. According to Gibbs’ phase rule, P = 2 and C = 2, so
`F = C — P —|— 2 = 2 degrees of freedom. These are the temperature and pressure,
`which we have specified as fixed. Thus there remain n0 degrees of freedom;
`the system is invariant. This means that at fixed temperature and pressure, the con-
`centration of dissolved solute is fixed. We call this invariant dissolved concentration
`
`the equilibrium solubility of the solute at this pressure and temperature. (We say
`that the solution is saturated.) Our present concern is with how the equilibrium
`solubility depends on the temperature and on the chemical natures of the solute
`and the solvent.
`
`Expressed as a reaction, the dissolution process is
`
`Pure solute : solute in solution
`
`At equilibrium the chemical potentials of the solute in the two phases are equal, or,
`letting component 1 be the solvent and component 2 the solute
`
`1 1 6
`
`I12 (solid) : I12 (soln)
`
`000003
`
`000003
`
`

`

`THE IDEAL SOLUBILITY
`
`117
`
`Writing out the chemical potentials gives
`
`“‘2’ (solid) + RT ln a2 (solid) : u; (soln) + RT In 612 (soln)
`
`(10.1)
`
`where the standard state of the solid is the pure solid, and we will adopt as the
`standard state of the solute in solution the Henry’s law definition on the molar
`concentration scale. Rearranging Eq. (10.1) leads to
`
`Au° : —RT In
`
`a2 (soln)
`
`oz (solid)
`
`(10.2)
`
`where Aug 2 pg (soln) — u; (solid). But the solid is in its standard state,
`(12 (solid) : 1.0 by definition, and we obtain
`
`so
`
`A110 2 —RTlna2 (soln)
`
`(10.3)
`
`is the activity corresponding to
`is invariant—it
`We have seen that a2 (soln)
`the equilibrium solubility—so comparison of Eq. (10.3) with the fundamental
`thermodynamic result
`
`AG" 2 —RTan
`
`(10.4)
`
`leads to the conclusion that a2 (soln), the activity of the solute in a saturated
`solution, must have the character of an equilibrium constant. As a consequence,
`we can evaluate standard free energy, enthalpy, and entropy changes for the solution
`process in the usual manner (Chapter 4). These quantities are respectively called the
`free energy, heat, and entropy of solution.
`For nonelectrolyte solutes, particularly those of limited solubility, so that the
`saturated solution is fairly dilute, it will be acceptable to approximate the activity
`a2 (soln) by the equilibrium solubility concentration. This is usually in molar
`concentration units, and is often symbolized s.
`
`10.2. THE IDEAL SOLUBILITY
`
`A thermodynamic argument can predict the equilibrium solubility of a nonelectro-
`lyte, provided it dissolves to form an ideal solution. Ideal behavior does not mean
`that intermolecular interactions are absent. On the contrary, solids and liquids
`would not exist without the intermolecular forces of interaction. In the present
`context, ideal behavior means that the energy of interaction between two solvent
`molecules is identical to that between one solvent and one solute molecule, so
`that a solvent molecule may be replaced with a solute molecule without altering
`the intermolecular energies. (This requires that the solvent and solute molecules
`have the same size, shape, and chemical nature, a demanding set of limitations.)
`
`000004
`
`000004
`
`

`

`1 18
`
`SOLUBILITY
`
`Quantitatively, an ideal solution can be defined as one having the following proper-
`ties (Chapter 7):
`
`AHmix = 0
`
`AVmix = 0
`
`ASmix = —R(X11I1X1 +X21I1X2)
`
`(10.5)
`
`(10.6)
`
`(10.7)
`
`According to Eqs. (10.5) and (10.6), there is no heat or volume change on mixing
`the solute and solvent in an ideal solution, and the entropy change is given by
`Eq. (10.7). Since an +x2 = 1, the logarithmic terms are necessarily negative, so
`ASmix is positive, and this constitutes the “driving force” for dissolution, because
`of the relationship AG 2 AH — TAS.
`If the entropy of mixing is the driving force for dissolution, what is the “resis-
`tance”? It is the solute—solute interaction forces, which, for solids, lead to the
`“crystal
`lattice energy.” These must be overcome for the solute to dissolve.
`Now, the free-energy change for the dissolution process is the same no matter
`what reversible mechanism (path) is taken to pass from the initial state (pure solute)
`to the final state (saturated solution), so we can divide the process as follows (for a
`solid solute):
`
`Crystalline solute \— supercooled liquid solute
`
`Pure liquid solvent
`
`\— solvent containing cavity
`
`Supercooled liquid solute F saturated solution
`
`+ solvent-containing cavity
`
`Crystalline solute + pure liquid solvent ;‘ saturated solution
`
`Since in an ideal solution the solvent—solvent interactions match the solvent—solute
`
`interactions, the energy required to create molecule-sized cavities in the solvent is
`offset by the energy recovered when the solute molecules are inserted into these
`cavities. The energetic cost of the dissolution process then appears in the first
`step, the melting of the solid. An equivalent viewpoint (Grant and Higuchi 1990,
`p. 16) is that the enthalpy of solution is given by
`
`AI'Isoln : AIifusion + AI'Imix
`
`But AHmix : 0 for an ideal solution, so AHSOIn : AHfusion.
`The saturation solubility, we have seen, is an equilibrium constant, so the van’t
`Hoff equation [Eq. (4.29)] is applicable
`
`
`(11an _ AHf
`dT _ RT2
`
`(10.8)
`
`where the solubility is expressed as the mole fraction simply to maintain con-
`sistency with Eq. (10.7), and where AHf is the heat of fusion and T is the absolute
`
`000005
`
`000005
`
`

`

`THE IDEAL SOLUBILITY
`
`119
`
`temperature. We have seen above why the heat of fusion appears in a solubility
`expression. (Incidentally, a dissolved solid should be viewed as possessing some
`of the properties of the liquid state, consistent with the above view that fusion is
`the first step in the dissolution process.) Now suppose that AHf is independent
`of temperature, which is equivalent to writing, for the solute from Eq. (1.23):
`
`ACp = €1qu — c;°“d = 0
`
`(109)
`
`Then integrating Eq. (10.8) from Tm to T gives
`
`
`
`1an = —% (Tm _ T)
`
`R
`
`Tr,"
`
`(1010)
`
`where Tm is the melting temperature and T is the experimental
`Equation (10.10) allows us to calculate the ideal solubility.
`
`temperature.
`
`Example 10.1. The melting point of naphthalene is 802°C, and its heat of fusion at
`the melting point is 4.54 kcal mol‘l. What is the ideal solubility of naphthalene
`at 20°C?
`
`Logx2 2
`
`—4540calmol‘l
`
`(2.303)(1.987 cal mol’lK’l)
`= —0.577
`
`<
`
`60.2K
`
`353.35 K x 293.15 K
`
`>
`
`x2 = 0.265
`
`Deviations from ideality will be manifested by discrepancies from the ideal solubi-
`lity as calculated with Eq. (10.10). Table 10.1 lists equilibrium solubilities for
`
`Table 10.1. Naphthalene solubility at 20°C
`
`Solvent
`
`(Ideal)
`Chlorobenzene
`Benzene
`Toluene
`Carbon tetrachloride
`Hexane
`Aniline
`Nitrobenzene
`Acetone
`n-Butanol
`Methanol
`Acetic acid
`
`Water (25°C)
`
`x2
`
`0.265
`0.256
`0.241
`0.224
`0.205
`0.090
`0. 130
`0.243
`0. 183
`0.0495
`0.0180
`0.0456
`
`0.0000039
`
`000006
`
`000006
`
`

`

`120
`
`SOLUBILITY
`
`naphthalene in many solvents. Observe that those solvents most chemically like
`naphthalene, that is, aromatic and nonpolar solvents, show behavior most closely
`approximating ideal behavior.
`At the melting temperature Tm the solid and liquid forms of the solute are
`in equilibrium, so AG,« 2 0 and we get AHf = TmASf, giving Eq. (10.11) as an
`alternative form of Eq. (10.10):
`
`lIle = —
`
`mm, — T)
`RT
`
`1.11
`(0 )
`
`10.3. TEMPERATURE DEPENDENCE OF THE SOLUBILITY
`
`Since AHf is always a positive quantity, Eq. (10.10) predicts that the solubility of
`a solid will increase with temperature. Moreover, Eq. (10.10) shows that if two solid
`substances have the same heat of fusion, the one with the higher melting point will
`have the lower solubility. Conversely, if they have the same melting point, the one
`with the lower heat of fusion will have the higher solubility. All of these inferences
`from Eq. (10.10) refer to systems forming ideal solutions, so deviations from the
`predictions can occur for real systems. Nevertheless,
`the increase of solubility
`with temperature is very widely observed for solids. Even the relationship of solu-
`bility to melting point can be a useful guide, though confounding phenomena can
`introduce complications; for example, hydrogen-bonding or other polar interactions
`may raise both the melting point and the aqueous solubility. The comparison of the
`temperature dependence of solubility of solids and gases is instructive; see
`Table 10.2.
`
`Equation (10.10) can be rearranged to Eq. (10.12):
`
`lnxz = —
`
`AHf AHf
`RT +RTm
`
`10.12
`
`)
`
`(
`
`Table 10.2. The contrary effects of temperature on the solubilities of solids and gases
`
`01 : 1111 w:
`ShdAH‘l.
`.dAHv
`Amg
`q
`
`as
`
`Solids
`
`Gases
`
`Solution is the process of
`passing from solid to liquid
`(fusion, AHf)
`
`Solution is the process of
`passing from gas to
`liquid (condensation, AHC), which
`is the reverse of vaporization (AHU)
`
`AHf is positive, so x2 increases
`as Tincreases
`
`AH“ is positive, so AHc is negative; thus x2
`decreases as T increases
`
`000007
`
`000007
`
`

`

`TEMPERATURE DEPENDENCE OF THE SOLUBILITY
`
`121
`
`Inx2
`
`1/T
`
`Figure 10.1. Hypothetical solubility van’t Hoff plots for polymorphs.
`
`If AHf is essentially constant over the experimental temperature range, Eq. (10.12)
`predicts that a plot of In x2 against
`1 / T will be linear with a slope equal to
`—AHf/R. The line should terminate at the melting point, where 1 / T = 1/ Tm. Often
`such lines are straight, probably because the usual range of temperatures is small.
`The slope gives AHf in principle, but in actuality the quantity evaluated from the
`slope is not precisely AHf because the solution is seldom ideal, and instead the
`quantity found in this way is termed the heat of solution.
`Throughout this discussion we have been assuming that the solid phase consists
`of the pure solid and not a solid solution. Another possible complication arises if
`the solid substance can exist in two crystalline forms (polymorphs; Chapter 6),
`which interconvert at transition temperature T[. The van’t Hoff plot can resemble
`Fig. 10.1a or Fig. 10.1b depending primarily on the kinetics of the transformation.
`In Fig. 10.1a, the two forms are sufficiently stable that their solubilities can be
`separately measured at the same temperatures, which are below the transition
`temperature. Nevertheless, the crystal form having the higher solubility (at a given
`temperature) is thermodynamically unstable (it is said to be metastable, since its
`kinetics of transformation permit it to exist for some period during which it acts
`as if it were stable), and will ultimately be converted to the stable form. Extrapola-
`tion of the lines to the transition temperature may be possible. Sulfathiazole in 95%
`ethanol shows the Fig. 10.1a behavior (Milosovich 1964; Carstensen 1977, p. 7).
`In Fig. 10.1b, one form exists in one temperature range, the other form in a
`temperature range on the other side of TI. The melting point observed will be that
`of the higher-melting polymorph. Carbon tetrabromide exemplifies this behavior
`(Hildebrand et a1. 1970, p. 23).
`
`000008
`
`000008
`
`

`

`122
`
`SOLUBILITY
`
`Let us return to the assumption that the change in heat capacities, AC , is zero,
`for all the subsequent discussion was based on this assumption. If AHf in fact is a
`function of temperature, then ACI, is not zero. Suppose we make the more reason-
`able assumption that ACI, is a nonzero constant, and write AHf as
`
`AHf 2 AH?“ — AC,,(Tm — T)
`
`(1013)
`
`where AH;n is the heat of fusion at Tm. Equation (10.13) is inserted into Eq. (10.8),
`which can be rearranged and integrated to give Eq. (10.14):
`
`
`. —AHm Tm — T
`AC Tm — T
`111x; =
`f
`"
`+
`R
`TTm
`R
`T
`
`
`Tm
`AC
`”111—
`R
`T
`
`—
`
`(10.14)
`
`This equation is useful for assessing the error that may be introduced by making the
`
`simple assumption ACI, = 0. Suppose, for example, that the experimental tempera-
`ture is 25°C and the melting point is 100°C. Then the last two terms in Eq. (10.14)
`become equal
`to 0.25ACp/R — 0.22 ACp/R : 0.03ACp/R. Thus considerable
`compensation can take place, making the approximation ACI, : 0 more acceptable
`than it might have seemed.
`
`Example 10.2. These are solubility data for nitrofurantoin in water (Chen et al.
`1976). Analyze the data to obtain the heat of solution.
`
`I (0C)
`
`24
`30
`37
`45
`
`105;;2
`
`6.01
`8.57
`13.16
`18.99
`
`The data are manipulated as required to make the van’t Hoff plot according to
`Eq. (10.12):
`
`T (103K)
`
`3.37
`3.30
`3.23
`3.14
`
`log x2
`
`—5.22
`—5.06
`—4.88
`—4.72
`
`The plot is shown in Fig. 10.2. It is possible that the points describe a curve, but this
`is uncertain with the data as given, for conceivably the scatter is a consequence of
`experimental random error. A straight line has therefore been drawn. Its slope is
`
`000009
`
`000009
`
`

`

`SOLUBILITY OF SLIGHTLY SOLUBLE SALTS
`
`1 23
`
`logX2
`
`3.1
`
`3.2
`
`3.3
`
`3.4
`
`Figure 10.2. van’t Hoff plot for nitrofurantoin solubility.
`
`103/ T
`
`—2300 K, so we calculate
`
`AHsoln = (2300 K)(1.987 cal mol’l K”)
`= 4570 cal mol‘l
`
`= 4.57 kcalmol‘l
`
`= 19.1 kJ mol’1
`
`Note that the enthalpy change is labeled AHsoln to indicate explicitly that this is a
`heat of solution.
`
`10.4. SOLUBILITY OF SLIGHTLY SOLUBLE SALTS
`
`Many salts exhibit very low solubilities in water. Silver chloride is an example;
`if aqueous solutions of silver nitrate and sodium chloride are mixed, solid silver
`chloride precipitates. It is conventional to describe this process as the reverse of
`the precipitation reaction, namely, as the dissolution of the salt. Let us begin
`with the simplest case of a 1 2 1 sparingly soluble salt MX. The solid crystalline
`form is ionic. When it dissolves in water the ions dissociate, and no ion pairs are
`detectable. We therefore write the equilibrium as
`
`MX(s) : M+ + x—
`
`(10.15)
`
`Proceeding as we have done for several earlier processes, we equate the chemical
`potentials of the solid and the dissolved solute at equilibrium:
`
`u<s> = u<soln>
`
`000010
`
`000010
`
`

`

`124
`
`SOLUBILITY
`
`Expanding these gives
`
`u°(s) + RTlna(s) = u: +RTlna+
`+ u: +RTlna_
`
`and collecting terms (and noting that a(s) = 1 by our standard state definition),
`
`A}? 2 —RT lna+a,
`
`(10.16)
`
`the
`where Au" : 111+ 11‘: — u°(s). Evidently then [compare with Eq. (10.4)],
`
`product a+a_ is an equilibrium constant. By Eq. (8.23a) we see that a+ a_ : c122,
`
`
`
`
`where a: is the mean ionic activity, and since 61?, = 721:1, Eq. (10.16) can b
`written
`
`
`
`Aw = —RT1nyEc?
`
`(10.17)
`
`If no extraneous ions are present, so that the ionic strength is due solely to the ions
`from the sparingly soluble salt (and hence is very low), the activity coefficient term
`is essentially unity. Moreover, the molar concentrations of the cation M+ and the
`anion X‘ are equal, and each is numerically equal
`to the equilibrium molar
`solubility of the salt, which is commonly denoted s. Thus Eq. (10.17) becomes
`
`Aw = —RTln 52
`
`(10.18)
`
`Equation (10.17) is exact; Eq. (10.18) is usually a reasonable approximation, and
`both implicitly define the equilibrium constant for Eq. (10.15). This constant is
`symbolized Ksp and is called the solubility product. Since solubility products are
`very small numbers, it is common to state them as szp, where szp = —log Ksp.
`Table 10.3 lists some szp values.
`
`Table 10.3. Solubility products for slightly soluble salts“
`
`Salt
`
`BaSO4
`CaCO3
`Ca(OH);
`Ca3(PO4)2
`CuI
`AuCl
`AuCl3
`Fe(OH)2
`Fe(OH)3
`
`pK,1,
`
`9.96
`8.54
`5 .26
`28.7
`1 1.96
`12.7
`24.5
`15.1
`37.4
`
`“In the temperature range 18725°C; water is the solvent.
`
`Salt
`
`PbCO3
`PbS
`MgCO3
`HgZS
`HgS (red)
`HgS (black)
`AgBr
`AgCl
`Agl
`
`pKg1,
`
`13.13
`27.9
`7.46
`47.0
`52.4
`51.8
`12.30
`9.75
`16.08
`
`000011
`
`000011
`
`

`

`SOLUBILITY OF SLIGHTLY SOLUBLE SALTS
`
`1 25
`
`Example 10.3. What is the solubility of silver chloride in water? From Table 10.3,
`szp = 9.75 for AgCl, so K5,, = 1.78 x 10*". From Eq. (10.18), K3,, = s2, so
`5 = K5,, = 1.33 x 10-5 M.
`
`In the general case of the salt whose formula is Mqu the solubility product is
`defined, in accordance with the usual formulation of equilibrium constants:
`
`KSp = 65:46;
`
`(10.19)
`
`The quantity that we label s then depends on the stoichiometry.
`
`Example 10.4. What is the molar solubility of ferrous hydroxide?
`From Table 10.3, szp = 15.1, or Ksp = 7.9 X 10716. The dissolution reaction is
`
`Fe(OH)2 : Fe2+ + 20H—
`
`so Ksp = cFecéH. (The charges on the subscripts are omitted for clarity.) Since each
`molecule of Fe (OH)2 that dissolves yields one Fe2+ ion, we define the solubility as
`the concentration of ferrous ion, or CFe = s. The stoichiometry yields 601.1 2 2 cFe,
`so the result is1
`
`Ksp = s X (25)2 = 453
`
`Therefore s = 5.8 X 10’6 M.
`
`Example 10.5. What is the solubility of silver chloride in 0.02MKCl? Assume
`activity coefficients are unity.
`is defined
`the solubility. The solubility product
`Again we set
`cAg = s,
`Ksp = cAgcCl; however,
`the chloride concentration has been augmented by the
`addition of potassium chloride, so we write cc] = 0.02 + s; that is, the chloride con-
`centration is the sum from two sources, the KCl and the AgCl. We therefore have
`Ksp : s(0.02 —|— s), which is a quadratic equation that can be solved for 5. Before
`doing that, however, it is worth trying the approximation CC] = 0.02, which involves
`neglecting the relatively small contribution from dissolution of the AgCl. Thus
`
`K,,, = 0.02s = 1.78 x 10—10
`s = 8.9 x 10‘9M
`
`First note that the approximation seems well justified. More interestingly, observe
`that the solubility of silver chloride has been reduced from about l x 10’5 M in
`water (Example 10.3) to about l x 10‘8 M in 0.02 M KCl. This is an example of
`the common ion efi‘ect. The solubility of any slightly soluble salt can be reduced by
`adding an excess of one of its constituent ions.
`
`The accuracy of such calculations can be improved by making use of the Debye—
`Hiickel equation to estimate the values of mean ionic activity coefficients.
`
`000012
`
`000012
`
`

`

`126
`
`SOLUBILITY
`
`10.5. SOLUBILITIES OF NONELECTROLYTES: FURTHER ISSUES
`
`Salt Effects. In Example 10.5 we encountered one type of salt effect. There is
`another type of salt effect that is observed when the solubility of a nonelectrolyte
`is studied as a function of ionic strength (or of the concentration of an added
`electrolyte). Compare the nonelectrolyte solubility in the absence and presence
`of added salt. Since the solid solute is present in both cases
`
`u(solid) 2 ”(CS 2 0) = “(09
`
`where cs is the concentration of added salt. Therefore a(cS : 0) : a(cs), or
`
`so'yo = .97
`
`(10.20)
`
`where so and s are the solubilities in the two cases. Thus 7/70 = 50/5; and since
`70 = l is a reasonable assumption, 7 = so/s, and we have a method for measuring
`nonelectrolyte activity coefficients. Moreover, it is found experimentally that the
`quantity log (so/s) often varies linearly with cs, or
`
`Logs—0 = kscs
`S
`
`(10.21)
`
`If so/s > 1, then kS is positive, and the nonelectrolyte is said to be “salted out”; if
`so/s < 1, then kS is negative, and the solute is “salted in.” These are called the
`“salting-out and salting-in effects,” and the constant kS is known as the Setschenow
`constant.
`
`Regular Solution Theory. We have seen that an ideal solution has thermody-
`namic mixing quantities AHmix = 0 and ASmix = —R(x1 In x1 + x2 ln x2). A regular
`solution is defined to be one having an ideal entropy of mixing but a nonideal
`enthalpy of mixing. Recall also that the ideal solubility of a nonelectrolyte (i.e.,
`the solubility when a nonelectrolyte forms an ideal solution) is given by
`
`
`lnxz = _AHf T'“ _ T
`R
`TTm
`
`(1022)
`
`where ACI, is assumed to be zero or negligible. The molecular interpretation of an
`ideal solution is that the energy of interaction of a solute molecule with a solvent
`molecule is identical with the energy of interaction of two solvent molecules.
`in
`The molecular interpretation of regular solution theory is quite different;
`regular solution theory the energy of 1—2 interactions (where l is the solvent, 2
`is the solute) is approximated as the geometric mean of l—l and 2—2 interaction
`energies, or2
`
`U12 = (U11U22)1/2
`
`(10.23)
`
`000013
`
`000013
`
`

`

`SOLUBILITIES OF NONELECTROLYTES: FURTHER ISSUES
`
`127
`
`This approximation results in regular solution theory being applicable mainly to
`fairly nonpolar systems, that is, nonpolar nonelectrolytes dissolved in nonpolar
`solvents. For our present interest, the essential result (Hildebrand and Scott 1964,
`p. 271) of regular solution theory is embodied in Eq. (10.24), which may be
`compared with Eq. (10.22):
`
`lIle = ——
`R
`
`AHf (Tm — T)
`
`TTIn
`
`
`—
`
`V20?
`
`(231 — 52)2
`
`(10.24)
`
`where V2 is the molar volume of solute and (p1 is the volume fraction concentration
`of solvent in the solution. The quantities 51 and 52 are the solubility parameters of the
`solvent and solute. These are physical properties with the following significance.
`
`The term AHvap, the molar heat of vaporization, is the enthalpy required to effect
`the transformation of one mole of liquid to its vapor state. During this process all
`the solvent—solvent interactions (which are responsible for the existence of the
`liquid phase) are overcome. A quantity called the cohesive energy density (CED)
`is defined
`
`AHvap — RT
`
`CED =
`
`(10.25)
`
`where V is the molar volume of the liquid. We anticipate, and we find, that liquids
`with strong intermolecular interactions (especially polar “associated” liquids
`having the potential for strong dipole—dipole and hydrogen-bonding interactions)
`have larger ced values than do nonpolar liquids. Table 10.4 lists some CED values.
`Because of the manner in which CED appears in regular solution theory equa-
`tions, Hildebrand (Hildebrand et a1. 1970; Hildebrand and Scott 1964, p. 271)
`defined the solubility parameter 5 by Eq. (10.26). Table 10.4 also gives 5 values.
`
`5 = (CED)‘/2
`
`(1026)
`
`Referring now to Eq. (10.24), note that if 81 = 52, we recover Eq. (10.22) for the
`ideal solution; in other words, the condition 51 = 52 is equivalent to the condition
`AHmix = 0. The greater the difference 51 — 52 (or of 52 — 51, because the differ-
`ence is squared), the greater the deviation from ideality, and, as Eq. (10.24) shows,
`the lower the solubility that is predicted. This provides a guide for experimental
`design; to achieve maximal solubility according to regular solution theory, strive
`to equate the solubility parameters of solvent and solute. Since the solute identity
`is usually established by the nature of the problem, the experimental variable is the
`solvent identity. Sometimes mixed solvent systems function better than do pure sol-
`vents for this reason. For example, a mixture of ether (5 = 7.4) and ethanol
`(5 = 12.7) dissolves nitrocellulose (5 = 11.2), although neither pure liquid serves
`as a good solvent for this solute.3
`Although the cohesive energy density, and therefore the solubility parameter, is a
`well-defined physical property for any solvent, regular solution theory is limited
`
`000014
`
`000014
`
`

`

`128
`
`SOLUBILITY
`
`Table 10.4. Cohesive energy densities and solubility parameters
`
`Solvent
`n—Pentane
`
`Cyclohexane
`1,4-Dioxane
`Benzene
`
`Diethyl ether
`Ethyl acetate
`Acetic acid
`
`n—Butyl alcohol
`n Propyl alcohol
`Acetone
`Ethanol
`Methanol
`Acetonitrile
`
`Dimethylformamide
`Ethylene glycol
`Glycerol
`Dimethylsulfoxide
`Water
`
`CED (cal cm’3)
`50.2
`
`«3(ca11/2 cm‘3/2)
`7.0
`
`67.2
`96
`84.6
`
`59.9
`83.0
`102
`
`130.0
`141.6
`95
`168
`212
`141.6
`
`146.4
`212
`272
`144
`547.6
`
`8.2
`10.0
`9.2
`
`7.4
`9.1
`10.1
`
`11.4
`11.9
`9.9
`12.7
`14.5
`1 1.9
`
`12. 1
`14.6
`16.5
`12.0
`23.4
`
`(e.g., by the geometric mean approximation) to solutions of nonpolar substances. It
`should therefore not be expected to apply quantitatively to polar systems such as
`aqueous solutions.
`
`Example 10.6. Predict the solubility of naphthalene in n-hexane at 20°C. The solu-
`bility parameters are 51 = 7.3 and 52 = 9.9 (both in call/2 cm_3/2), and the molar
`volumes are V1 = 132 cm3 mol’1 and V2 = 123 cm3 mol’l. See Example 10.1 for
`additional data.
`
`We use Eq. (10.24), which in Example 10.1 was expressed in terms of log x2. In
`that form the first term on the right had the value —0.577, which we need not recal-
`culate. Now we consider the second term. We lack only the quantity (p1, the volume
`fraction of solvent. This appears to be a dilemma, because we cannot estimate (p1
`until we know x2, which is what we want to calculate.
`If we anticipate that the solute has a low solubility, it may be acceptable to make
`the approximation (p1 = 1. An alternative is to take the result for an ideal solution
`(Example 10.1, which gave x2 = 0.265) as a basis for estimating (p1. We will do the
`problem in both ways.
`(a) Let (p1 = 1. Then from eq. (10.24),
`
`Log x2 = —0.577 —
`
`(123 cm3)(1)2(7.3 — 9.9 call/2 mr3/2)2
`(2.303)(1.987 cal mol’l K-1)(293.15 K)
`= —0.577 — 0.620 = —1.197
`
`x2 = 0.064
`
`000015
`
`000015
`
`

`

`SOLUBILITIES OF NONELECTROLYTES: FURTHER ISSUES
`
`129
`
`(b) The volume fraction is defined as follows:
`
`(P1
`
`MW
`2—
`n1V1 + 11sz
`
`10.27
`
`)
`
`(
`
`Suppose n1 +n2 = 1; from Example 10.1, x2 = 0.265, or n2 2 0.265 and n1 2
`0.735. Using these numbers in Eq. (10.27) gives (p1 = 0.748. (Note how close
`(p1 is to am, because V1 and V2 are similar.) Repeating the calculation gives
`
`Log x2 = —0.577 — 0.335 = —0.912
`
`x2 = 0.122
`
`We therefore predict that x2 is between 0.064 and 0.122, and we might take the
`average as our best estimate. The experimental result (Table 10.1) is x2 = 0.090.
`
`Prediction Of Aqueous SOIubiIities. Water is the preferred solvent for liquid
`dosage forms because of its biological compatibility, but unfortunately many drugs
`are poorly soluble in water. To be able to predict the aqueous solubility of com-
`pounds, even if only approximately, is a valuable capability because it can guide
`or reduce experimental effort. Water is a highly polar and structured medium in
`which nonideal behavior is commonly observed, so we must abandon hope that
`the ideal solubility prediction of Eq. (10.10) will be useful, and even the regular
`solution theory [Eq.
`(10.24)] is ineffectual
`in solving this problem. Effective
`approaches may be guided by thermodynamic concepts, but they incorporate
`much empirical (i.e., experimental) content.
`Although the ideal solubility equation will not suffice to predict nonelectrolyte
`solubility in water,
`the solute—solute interactions responsible for maintaining
`the crystal lattice must nevertheless be overcome, so Eq. (10.10) will still be applic-
`able as a means of estimating the solute—solute interaction. What must be done in
`addition is to take account of the solvent—solvent and solvent—solute interactions,
`for these will in general not offset each other. In a paper that includes a valuable
`collection of solubility data, Yalkowsky and Valvani (1980) have developed a very
`useful method based on this approach. They start with Eq. (10.10), which they
`transform to Eq. (10.11), repeated here:
`
`lIlX2=—
`
`Asfm, — T)
`RT
`
`1.2
`(0 8)
`
`They then carry out an analysis of experimental entropies of fusion, reaching these
`conclusions:
`
`For spherical (or nearly so) molecules: ASf = 3.5 cal mol’1 K—1
`
`For rigid molecules: ASf = 13.5 cal mol’1 K’1
`
`000016
`
`000016
`
`

`

`130
`
`SOLUBILITY
`
`having n > 5
`For molecules
`(n — 5) calmol’l K—1
`
`flexible
`
`chain
`
`atoms: ASf = 13.5 + 2.5
`
`In the following we will use only the result for rigid molecules.
`Yalkowsky and Valvani then take the log P value of the solute (where P is the
`1-octanol/water partition coefficient) as an empirical measure of the solution phase
`nonidealities. They combine this with Eq. (10.28), convert to molar concentration,
`and apply a small statistical adjustment, finally getting Eq. (10.29) for the calcula-
`tion of rigid nonelectrolyte molar solubility in water at 25°C:
`
`Log (:2 = —o.o11(:m — 25) — log P + 0.54
`
`(10.29)
`
`where tm is the solute melting point in centigrade degrees. For liquid nonelectro-
`lytes tIn is set to 25, so the first term vanishes. Log P may be available from
`experimental studies, but it may have to be estimated by methods cited in Chapter 7.
`Yalkowsky and Valvani applied Eq. (10.29) to solubility data on 167 compounds
`whose solubilities ranged over nine orders of magnitude, finding that the estimated
`solubilities agreed with the observed solubilities to within 0.5 log unit for all but
`eight compounds, and in no case was the error greater than a factor of 10. Equation
`(10.29) is a very practical solution to the problem of predicting aqueous solubilities.
`Amidon and Williams (1982) refined the approach of Yalkowsky and Valvani,
`achieving better accuracy but at the cost of increased complexity in the equation.
`Grant and Higuchi (1990) describe alternative methods of calculation that are based
`on different pathways from the initial to the final state.
`Equation (10.10) and equations derived from it, such as Eqs. (10.28) and (10.29),
`contain the difference (Tm — T), showing that a higher melting temperature is
`reflecting stronger solute—solute interactions in the solid state. As a general but
`not precise rule, we may anticipate that very polar molecules (or functional
`groups) will conduce to strong intermolecular interactions by means of electrostatic
`forces, which for certain groups may include hydrogen bonding. Thus high
`molecular polarity tends to be associated with high melting temperature, and higher
`melting temperatures lead to lower solubilities, at least as they are described by
`Eq. (10.10).
`Now consider the special case of water as a solvent. Water is a very polar solvent
`and is capable of functioning as a hydrogen bond donor and acceptor. Very polar
`solute molecules will tend to interact strongly with the solvent water; these are the
`solvent—solute or solvation interactions that increase solubility. But we have seen
`that highly polar substances tend to have high melting temperatures, so we are
`led to the tentative conclusion that melting temperature may be an approximate
`indicator of the extent of solvent—solute interaction. It follows (still arguing in
`this approximate mode) that the opposing factors of solute—solute (crystal lattice)
`and solvent—solute (solvation)
`interactions are both measured by, or a

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket