throbber
P1: FQF/SJS
`
`P2: FQF
`
`SVNY358-Augustijns
`
`April 15, 2007
`
`8:16
`
`Solvent Systems and Their
`Selection in Pharmaceutics
`and Biopharmaceutics
`
`Patrick Augustijns
`Catholic University of Leuven, Belgium
`
`Marcus E. Brewster
`Janssen Pharmaceutica N.V., Beerse, Belgium
`
`MYLAN EXHIBIT 1026
`
`iii
`
`

`

`P1: FQF/SJS
`
`P2: FQF
`
`SVNY358-Augustijns
`
`April 12, 2007
`
`9:4
`
`Patrick Augustijns
`Laboratory for Pharmacotechnology
`and Biopharmacy
`Catholic University of Leuven
`Belgium
`
`Marcus E. Brewster
`Janssen Pharmaceutica N.V.
`Beerse, Belgium
`
`Library of Congress Control Number: 2007924356
`
`ISBN-13: 978-0-387-69149-7
`
`e-ISBN-13: 978-0-387-69154-1
`
`Printed on acid-free paper.
`C(cid:2) 2007 American Association of Pharmaceutical Scientists.
`All rights reserved. This work may not be translated or copied in whole or in part without the written
`permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York,
`NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in
`connection with any form of information storage and retrieval, electronic adaptation, computer
`software, or by similar or dissimilar methodology now known or hereafter developed is forbidden.
`The use in this publication of trade names, trademarks, service marks, and similar terms, even if they are
`not identified as such, is not to be taken as an expression of opinion as to whether or not they are subject
`to proprietary rights.
`
`While the advice and information in this book are believed to be true and accuate at the date of going
`to press, neither the authors nor the editors nor the publisher can accept any legal responsibility for any
`errors or omissions that may be made. The publisher makes no warranty, express or implied, with respect
`to the material contained herein.
`
`9 8 7 6 5 4 3 2 1
`
`springer.com
`
`iv
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`1
`
`Principles of Solubility
`
`YUCHUAN GONG AND DAVID J.W. GRANT
`Department of Pharmaceutics, College of Pharmacy,
`University of Minnesota, Minneapolis, MN
`
`HARRY G. BRITTAIN
`Center for Pharmaceutical Physics, Milford, NJ
`
`Introduction
`Solubility is defined as the maximum quantity of a substance that can be com-
`pletely dissolved in a given amount of solvent, and represents a fundamental
`concept in fields of research such as chemistry, physics, food science, pharma-
`ceutical, and biological sciences. The solubility of a substance becomes especially
`important in the pharmaceutical field because it often represents a major factor
`that controls the bioavailability of a drug substance. Moreover, solubility and
`solubility-related properties can also provide important information regarding
`the structure of drug substances, and in their range of possible intermolecular
`interactions. For these reasons, a comprehensive knowledge of solubility phe-
`nomena permits pharmaceutical scientists to develop an optimal understanding
`of a drug substance, to determine the ultimate form of the drug substance, and
`to yield information essential to the development and processing of its dosage
`forms.
`the solubility phenomenon will be developed using
`In this chapter,
`fundamental theories. The basic thermodynamics of solubility reveals the re-
`lation between solubility, and the nature of the solute and the solvent, which
`facilitates an estimation of solubility using a limited amount of information.
`Solubility-related issues, such as the solubility of polymorphs, hydrates, solvates,
`and amorphous materials, are included in this chapter. In addition, dissolution
`rate phenomena will also be discussed, as these relate to the kinetics of solubility.
`A discussion of empirical methods for the measurement of solubility is outside
`the scope of this chapter, but is reviewed elsewhere (Grant and Higuchi, 1990;
`Grant and Brittain, 1995).
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`2
`
`Chapter 1: Principles of Solubility
`
`Units for the Expression of Solubility
`A discussion of the thermodynamics and kinetics of solubility first requires a
`discussion of the method by which solubility is reported. The solubility of a sub-
`stance may be defined in many different types of units, each of which represents
`an expression of the quantity of solute dissolved in a solution at a given temper-
`ature. Solutions are said to be saturated if the solvent has dissolved the maximal
`amount of solute permissible at a particular temperature, and clearly an un-
`saturated solution is one for which the concentration is less than the saturated
`concentration. Under certain conditions, metastable solutions that are supersatu-
`rated can be prepared, where the concentration exceeds that of a saturated solu-
`tion. The most commonly encountered units in pharmaceutical applications are
`molarity, normality, molality, mole fraction, and weight or volume percentages.
`The molarity (abbreviated by the symbol M) of a solution is defined as the
`number of moles of solute dissolved per liter of solution (often written as mol/L
`or mol/dm3), where the number of moles equals the number of grams divided
`by its molecular weight. A fixed volume of solutions having the same molarity
`will contain the same number of moles of solute molecules. The use of molarity
`bypasses issues associated with the molecular weight and size of the solute, and
`facilitates the comparison of different solutions. However, one must exercise
`caution when using molarity to describe the concentrations of ionic substances
`in solution, because the stoichiometry of the solute may cause the solution to
`contain more moles of ions relative to the number of moles of dissolved solute.
`For example, a 1.0 M solution of sodium sulfate (Na2SO4) would be 1.0 M in
`sulfate ions and 2.0 M in sodium ions.
`The normality (abbreviated by the symbol N) of a solution is defined as the
`number of equivalents of solute dissolved per liter of solution, and can be written
`as eq/L or eq/dm3. Normality has the advantage of describing the solubility of
`the ionic compounds since it takes into account the number of moles of each
`ion in the solution liberated upon dissolution of a given number of moles of
`solute. The number of equivalents will equal the number of grams divided by
`the equivalent weight. For ionic substances, the equivalent weight equals the
`molecular weight divided by the number of ions in the compound. Equivalent
`weight of an ion is the ratio of its molecular (atomic) weight and its charge.
`Therefore, a molar solution of Na2SO4 is 2 N with respect to both the sodium
`and the sulfate ion. Since the volume of solution is temperature dependent,
`molarity and normality can not be used when the properties of solution, such as
`solubility, is to be studied over a wide range of temperature.
`Molality is expressed as the number of moles of solute dissolved per kilogram
`of solvent, and is therefore independent of temperature since all of the quanti-
`ties are expressed on a temperature-independent weight basis. The molality of a
`solution is useful in describing solubility-related phenomena at various temper-
`atures, and as the concentration unit of colligative property studies. When the
`density of the solvent equals unity, or in the case of dilute aqueous solutions, the
`molarity and the molality of the solution would be equivalent.
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`3
`
`Expressing solution concentrations in terms of the mole fraction provides the
`ratio of the number of moles of the component of interest to the total number
`of moles of solute and solvent in the solution. In a solution consisting of a single
`solute and a single solvent, the mole fraction of solvent, XA, and solute, XB, is
`expressed as:
`
`(1)
`
`(2)
`
`(3)
`
`XA = nA
`nA + nB
`XB = nB
`nA + nB
`where nA and nB are the number of moles of solvent and solute, respectively.
`Obviously the sum of the mole fraction of the two components must equal one:
`XA + XB = 1
`Since mole fractions provide quantitative information of a mixture that can be
`readily translated down to the molecular level, this unit is most commonly used
`in thermodynamic studies of solubility behavior.
`Volume fraction is frequently used to define the composition of mixed solvent
`systems, or to express the solubility of one solvent in another. However, since the
`volumes of solutions exhibit a dependence on temperature, the expression of
`concentrations in terms of volume fraction requires a simultaneous specification
`of the temperature. In addition, since volume defects may occur during the mix-
`ing of the solvents, and since these will alter the final obtained volume, defining
`the solubility of a solution in terms of volume fraction can lead to inaccuracies
`that can be avoided through the use of other concentration parameters.
`The concept of percentage is widely used as a concentration parameter in
`pharmaceutical applications, and is expressed as the quantity of solute dissolved
`in 100 equivalent units of solution. The weight percentage (typically abbreviated as
`% w/w) is defined as the number of grams of solute dissolved in 100 grams of
`solution, while the volume percentage (typically abbreviated as % v/v) is defined as
`the number of milliliters of solute dissolved in 100 mL of solution. A frequently
`encountered unit, the weight-volume percentage (typically abbreviated as % w/v)
`expresses the number of grams of solute dissolved in 100 mL of solution. The
`choice of unit to be used depends strongly on the nature of solute and solvent,
`so the solubility of one liquid in another is most typically expressed in terms
`of the volume percentage. The use of weight or weight-volume percentages is
`certainly more appropriate to describe the concentration or solubility of a solid
`in its solution.
`For very dilute solutions, solubility is often expressed in units of parts per
`million (ppm), which is defined as the quantity of solute dissolved in 1,000,000
`equivalent units of solution. As long as the same unit is used for both solute and
`solvent, the concentration in parts per million is equivalent to the weight, vol-
`ume, or weight-volume percentages multiplied by 10,000. The descriptive terms
`of solubility that is expressed in units of parts of solvent required for each part of
`solute can be found in each edition of the United States Pharmacopeia (Table 1).
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`4
`
`Chapter 1: Principles of Solubility
`
`Descriptive term
`
`Very soluble
`
`Freely soluble
`
`Soluble
`
`Sparingly soluble
`
`Slightly soluble
`
`Parts of solvent required
`for 1 part of solute
`
`Solubility < 1
`
`1 < Solubility < 10
`
`10 < Solubility < 30
`
`30 < Solubility < 100
`
`100 < Solubility < 1,000
`
`Very slightly soluble
`
`1,000 < Solubility < 10,000
`
`Practically insoluble, or Insoluble
`Table 1. Descriptive terms of solubility.
`Reproduced from:
`United States Pharmacopeia, 25th edition. United States Pharmacopeial Convention;
`Rockville, MD; 2002, p. 2363.
`
`Solubility > 10,000
`
`Thermodynamics of Solubility
`The equilibrium solubility of a substance is defined as the concentration of solute in
`its saturated solution, where the saturated solution exists in a state of equilibrium
`with pure solid solute. As solutes and solvents can be gaseous, liquid, or solid,
`there are nine possibilities for solutions, although liquid-gas, liquid-liquid, and
`liquid-solid are of particular interest for pharmaceutical applications. Among
`these, the most frequently encountered solubility behavior involves solid solutes
`dissolved in liquid solvent, so systems of this type will constitute the examples of
`the following discussions.
`For the particular system of a saturated solution, the dissolved solute in the
`solution and the undissolved solute of the solid phase are in a state of dynamic
`equilibrium. Under those conditions, the rate of dissolution must equal the
`rate of precipitation and hence the concentration of the solute in the solution
`remains constant (as long as the same temperature is maintained).
`For two phases in equilibrium, the chemical potential, μi, of the component
`in the two phases must be equal:
`
`μsolute = μsolid
`The chemical potential, also known as the molar free energy, can be represented
`by:
`
`(4)
`
`μ = μ◦ + RT ln a
`(5)
`◦ is the chemical potential of the solute molecule in its reference state,
`where μ
`and a is the activity of the solute in the solution. Since both the dissolved solute
`and the undissolved solid must refer back to the same standard state, it follows
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`5
`
`that the activities of the dissolved solute and that of the undissolved solid must
`be identical.
`The activity of a component in a solution is defined as the product of its
`activity coefficient, γ , and its mole fraction, X:
`a = γX
`For the solute B in a saturated solution:
`asolid = asolute = γB XB
`
`(7)
`
`(6)
`
`or
`
`XB = asolid
`γB
`
`(8)
`
`According to equation (8), the solubility of a substance would be propor-
`tional to the activity of the undissolved solid, and inversely proportional to its
`activity coefficient. Although the activity of a substance in its standard state is
`defined as unity, the activity of the undissolved solid must depend on reference
`state. A hypothetical, supercooled liquid state of solute at the temperature of
`interest is commonly taken as the standard state, making the activity coefficient
`a more complicated term. The activity coefficient will depend on the nature of
`both the solute and solvent, as well as on the temperature of the solution.
`
`Solubility in Ideal Solutions
`
`In order to understand the thermodynamics of solubility, it is appropriate to
`begin with a simplified model of solution, namely that of an ideal solution. An
`ideal solution is defined as one where the activity coefficient of all components
`in the solution equals one. Under these stipulations, the activity of the dissolved
`solute, the activity of the solid, and the molar solubility of the dissolved solute
`would be equal.
`
`(9)
`
`asolute = asolid = XB
`As discussed above, the absolute activity of the solid depends on the chosen
`reference or standard state, and the usual practice is to take the supercooled
`liquid state of the pure solute at the temperature of solution as the standard
`state of unit activity. At temperatures lower than the melting point, the liquid
`state of the solute is less stable than its solid state, making the activity of the
`corresponding solid less than one.
`An ideal solution requires that the scope of solute-solute, solvent-solvent,
`and solute-solvent intermolecular forces be all the same. Thus, the net energy
`change associated with breaking bonds between two solute molecules and two
`solvent molecules, and then forming new bonds between solute and solvent
`molecules must be zero. Moreover, the mixing process is ideal as well, so that the
`total volume of the solute/solvent system does not change during the mixing
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`6 p
`
`Chapter 1: Principles of Solubility
`
`rocess.
`
` Umix = 0
` Hmix = 0
` Vmix = 0
`
`(10)
`
`(11)
`
`(12)
`
`where Umix is the energy of mixing, Hmix is the enthalpy of mixing, and Vmix
`is the volume change of mixing. The ideal entropy of mixing, Smix, can be
`derived from pure statistical substitution
` Smix = −R(nA ln XA + nB ln XB)
`
`(13)
`
`where nA and nB are the number of moles of the solvent (A) and the solute (B),
`respectively. Because the mole fractions of the solvent and the solute, XA and XB,
`are less than unity, it follows that Smix is always positive. From this analysis, one
`can conclude that the mixing processes associated with an ideal solution would
`be thermodynamically favored.
`The dissolution of a solid in a solvent can be considered as consisting of two
`steps. The first step would be, in effect, a melting of the solid at the absolute
`temperature (T) of the solution, and the second step would entail mixing of the
`liquidized solute with the solvent. The enthalpy of solution ( Hs) is therefore
`equal to the sum of the enthalpy of fusion ( HT
`f ) and the enthalpy of mixing
`( Hmix). However, since the enthalpy of mixing must equal zero for an ideal
`solution, it follows that the enthalpy of solution must equal the enthalpy of
`fusion of the solid at the given temperature, T:
` Hs = HT
`
`(14)
`
`f
`
`f
`
`f
`
`(15)
`
`For those situations where the temperature of study is not the same as the melt-
`(cid:3)= Hm
`f , where now Hmf
`ing point, then HT
`is the enthalpy of fusion at the
`f
`
`melting point( Tm). If one makes the approximation that the enthalpy of fusion
`is constant over the temperature range in the vicinity of the melting point, then:
` Hs = HT
`≈ Hm
`(cid:3)
`(cid:2)
`
`Applying the Clausius-Clapeyron equation to the solubility calculation yields:
`= Hs
`RT2
`
`P
`
`(16)
`
`∂ ln a
`∂T
`
`Integration of equation (16) provides the relationship known as the van’t Hoff
`equation, which expresses the temperature dependence of the solubility of a
`solid solute (identified as species B) in an ideal solution:
`− 1
`ln XB = ln aB = − Hs
`R
`Tm
`
`(cid:3)
`
`1 T
`
`(cid:2)
`
`(17)
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`7
`
`(18)
`
`(19)
`
`(cid:3)
`
`1 T
`
`(cid:2)
`
`f
`
` G m
`f
`
`By combining equations (15) and (17), one finds that the molar solubility of the
`solute in an ideal solution (expressed in natural logarithmic form) is given by:
`ln XB = ln aB = − Hm
`− 1
`R
`Tm
`Since the solid solute and its corresponding molten solid must be in a state of
`equilibrium at the melting point, it follows that:
`= 0
`= Hm
`− Tm Sm
`f
`f
`where the enthalpy of fusion ( Hm
`
`f ) is equal to Tm Smf , where Smf
`is the entropy
`
`of fusion at the melting temperature. Under these circumstances, equation (18)
`(cid:2)
`(cid:3)
`may also be written as:
`
`f
`
`− 1
`
`(20)
`
`ln XB = ln aB = − Sm
`Tm
`T
`R
`The enthalpy and entropy of fusion, and the melting temperature may all be mea-
`sured through the use of differential scanning calorimetry (DSC), and therefore
`equations (18) and (20) provide a simple way to predict the solubility of a solute
`in an ideal solution.
`To achieve a better prediction of the solubility of a solute, one must consider
`the temperature dependence of the enthalpy of fusion, which is described by
`the Kirchoff equation
`
`(cid:2)
`
`(cid:3)
`
`= C p
`
`(21)
`
`(22)
`
`∂ Hf
`∂T
`P
`where Cp is the difference between the heat capacities of the supercooled liquid
`and that of the corresponding solid. Therefore:
`= Hm
`− C p (Tm − T)
` HT
`f
`f
`With the assumption that Cp is independent of temperature, integration of
`equation (16) and the replacement of Hs by HT
`f , yields the Hildebrand equa-
`(cid:2)
`(cid:3)
`tion
`Tm − T
`− C p
`− 1
`ln XB = ln aB = − Hm
`+ C p
`Tm
`T
`R
`T
`R
`Tm
`R
`Equation (23) provides a better prediction of the solubility of a solute in an ideal
`solution.
`Prediction of solubility in an ideal solution can also be performed using the
`entropy approach developed by Hildebrand and Scott (Hildebrand and Scott,
`1962). Assuming that Hs ≈ T Sm
`≈ T C p, they found that:
`ln XB = ln aB = − Sm
`R
`m
`Equation (24) is similar to equation (20), except that ln(XB) is correlated to
`ln(T) instead of 1/T. The solubility prediction using equation (24) was found
`
`ln
`
`(23)
`
`1 T
`
`f
`
`(24)
`
`T T
`
`f
`
`ln
`
`f
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`8 t
`
`o have a better tolerance for the non-ideality of the solution than that obtained
`using equation (20).
`Several approaches have been used to predict the entropy of fusion required
`for the prediction of solubility. According to Walden’s rule, the entropy of fusion
`f ) is approximately equal to 13 cal/K·mol for most organic compounds
`( Sm
`(Walden, 1908). Use of this approximation reduces equation (20) to:
`
`ln XB = ln aB = − θm − 25
`
`298.15
`
`(25)
`
`where θm is the melting point of the solute in degrees centigrade.
`Yalkowsky proposed that the entropy of fusion of an organic compound is
`the sum of translational, rotational, and internal entropy changes when it is
`released from the crystal lattice (Yalkowsky, 1979):
` Sf = Strans + Srot + Sint
`
`(26)
`
`while, the translational entropy change consists of the components associated
`with the expansion and change of position as the solid melts.
` Strans = Sexp + Spos
`
`(27)
`
`Yalkowsky also proposed empirical values and limits for these components. Both
`the Walden and Yalkowsky models provide ways by which one can predict the
`entropy of fusion, and therefore predict the solubility of the solute in an ideal
`solution.
`Over a small temperature range, the enthalpy of solution of a solid can be
`assumed to be independent of temperature. The van’t Hoff equation shows that
`ln(XB) increases with temperature, until the solid melts at T = Tm. At this con-
`dition, the solid forms a liquid in the absence of solvent, and since XB = 1, the
`slope of the van’t Hoff plot is equal to ( HS/R). The degree of ideality associ-
`ated with a given solution may therefore be tested by evaluating the degree of
`linear correlation between ln(XB) and 1/T. Figure 1 shows the ideal behavior
`of naphthalene dissolved in benzene and xylene, which is due to the similar
`nature of the molecules involved, and the strength of intermolecular interac-
`tions such as polarity, polarizability, molecular volume, and hydrogen-bonding
`characteristics (Grant and Higuchi, 1990). On the other hand, the molecular
`properties of ethanol are very different from those of naphthalene. Thus one
`finds that for solutions of naphthalene in ethanol, ln(XB) does not exhibit a
`linear dependence on 1/T, which is taken as an indication of the non-ideal
`character of the solution.
`Typically, one finds that the solubility that would be predicted assuming the
`model of an ideal solution is normally much higher than the solubility that is
`actually measured for a non-ideal solution.
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`9
`
`Figure 1. Van’t Hoff plot of the molar solubility of naphthalene in benzene, xylene,
`and ethanol as a function of the reciprocal of the absolute temperature. The solid
`line corresponds to equation (17) for the ideal solubility of solid. Reproduced from
`DJW Grant, and T Higuchi, Solubility Behavior of Organic Compounds, John Wiley &
`Sons, New York, NY, 1990, p. 17.
`
`Solubility in Regular Solutions
`
`One rarely encounters ideal solutions in practice, and practically all solutions of
`pharmaceutical interest are non-ideal in character. For such non-ideal solutions,
`the activity coefficient (γ B) of the solute does not equal one because the range
`of solute-solute, solvent-solvent, and solute-solvent interactions are significant.
`Therefore, one must consider the effect of the activity coefficient in order to
`predict the properties of non-ideal solutions:
`XB = aB
`γB
`ln XB = ln aB − ln γB
`
`(28)
`
`(29)
`
`In equations (28) and (29), aB is the activity of the dissolved solute and the
`undissolved solid, which may be evaluated using the hypothetical supercooled
`liquid as the standard state of unit activity. ln(aB) may be expressed by equation
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`10
`
`Chapter 1: Principles of Solubility
`
`1 T
`
`(cid:2)
`
`(17), as was the case for ideal solutions. Therefore:
`ln XB = − Hs
`− 1
`− ln γB
`R
`Tm
`
`(cid:3)
`
`(30)
`
`The value of the activity coefficient depends on many factors, and for non-ideal
`solutions the activity coefficient may be predicted from knowledge of the nature
`of the solute and the solvent.
`For the sake of simplicity, the prediction of activity coefficients in regular so-
`lutions, the simplest non-ideal solution, will be discussed. For a regular solution,
`the energy of mixing and the enthalpy of mixing are not negligible because the
`intermolecular solute-solute, solvent-solvent, and solute-solvent interactions are
`different. However, the total volume is still assumed to be unchanged during
`mixing.
`The activity coefficient in a regular solution can be estimated by considering
`the changes in intermolecular interaction energies that accompany the mixing
`of solute and solvent. For this purpose, the solution process may be divided
`into the three steps illustrated in Figure 2. The first step would consist of the
`removal of a solute molecule from its pure solute phase into the vapor phase,
`the second step would be the creation of a hole in the solvent for incorporation
`of the solute molecule, and the third step is the process where the free solute
`molecule fills the hole created in the solvent (Higuchi, 1949; Hildebrand and
`Scott, 1950; Martin, 1993).
`To begin the analysis, the potential energy of solute-solute, solvent-solvent,
`and solute-solvent pairs is identified as w BB, w AA, and w AB. In the first step, an
`energy equal to 2w BB must be absorbed to break the solute-solute interaction
`between two adjacent solute molecules in the solid. After the solute molecule
`
`Figure 2. Hypothetical steps in solution process.
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Chapter 1: Principles of Solubility
`
`11
`
`is removed to the vapor phase, the hole created in the solute closes, which
`releases an energy equal to w BB, making the net energy change associated with
`liberation of a solute molecule equal to w BB. In the second step, energy equal to
`w AA is absorbed to separate a pair of solvent molecules, and to produce a hole in
`the solvent which the solute molecule may occupy. Finally, the solute molecule
`liberated from its solid phase is inserted in the hole in the solvent, forming two
`solute-solvent interactions and releasing an energy equal to 2w AB. The overall
`potential energy change, u, is therefore:
` u = w AA + w BB − 2w AB
`Using this simplified model, Hildebrand and Wood (1933) proposed
`ln γB = (w AA + w BB − 2w AB)
`VB 2
`A
`RT
`where VB is the molar volume of the solute in the supercooled state, A is the
`volume fraction of the solvent in solution, R is the gas constant, and T is the
`absolute temperature of the solution.
`The attractive interactions between pairs of solute and solvent molecules are
`assumed to be derived from van der Waals forces, so the solute-solvent interaction
`energy (w AB) may be represented by the geometric mean of the solute-solute
`(w BB) and the solvent-solvent (w AA) interaction energies:
`
`(31)
`
`(32)
`
`w AB = √
`
`w AA w BB
`
`(33)
`
`Therefore, equation (32) becomes:
`ln γB = ((w AA) 12 − (w BB) 1
`
`2 )2 VB 2
`RT
`The square root of the interaction energy is defined as the solubility parameter,
`δ, and so equation (34) can be rewritten as:
`ln γB = (δA − δB)2 VB 2
`RT
`where δA and δB are the solubility parameters of the solvent and solute, respec-
`tively. In the case of a mixed solvent system, the total solubility parameter of the
`solvent mixture is given by:
`
`A
`
`A
`
`(34)
`
`(35)
`
`δA = φ1δ1 + φ2δ2 + ···
`where δ1 and δ2 refer to the respective solvent parameters of pure solvents 1 and
`2, and φ1 and φ2 are the respective volume fractions in the solvent mixture.
`Introducing equation (35) into equation (30) yields the Hildebrand solubil-
`(cid:3)
`(cid:2)
`ity equation describing regular solution behavior:
`− (δA − δB)2 VB 2
`ln XB = − Hs
`− 1
`R
`Tm
`RT
`According to equation (37), if the difference between δA and δB is very small,
`then the second term approaches zero. The implication of this is that a regular
`
`(36)
`
`A
`
`(37)
`
`1 T
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`12
`
`Chapter 1: Principles of Solubility
`
`2
`
`(38)
`
`and/or hydrogen-bonding interactions exist in the solution, w AB (cid:3)= √
`
`solution would behave in an ideal manner when the solute and solvent have
`similar chemical properties. It may be seen that the Hildebrand solubility equa-
`tion enables the prediction of solubility in regular solutions, as long as one has
`knowledge of the solubility parameters of both components in the solution.
`Following the introduction of the Hildebrand model, the topic of solu-
`bility parameters has been extensively discussed (Hildebrand and Scott, 1962;
`Hildebrand et al., 1970; Kumar and Prausnitz, 1975; Barton, 1983), and values
`of δ can be found in these reference works. As a general rule, compounds hav-
`ing stronger London forces will be characterized by larger solubility parameters
`values.
`Hildebrand and Scott (1950) proposed that the solubility parameters of sim-
`ilar molecules could be calculated using the enthalpy of vaporization ( Hv) and
`(cid:2)
`(cid:3) 1
`the molar volume of the liquid component (Vl) at the temperature of interest:
` Hv − RT
`δ =
`Vl
`Predictions of the solubility of non-polar solutes in non-polar solvents have
`been successfully achieved using the Hildebrand solubility equation (Davis et al.,
`1972). These solutions may be classified as regular solutions since the primary
`intermolecular interactions are London dispersion forces. However, the equa-
`tion does not provide a good prediction of solubility for solutions involving
`polar components. When dipole-dipole, dipole-induced-dipole, charge transfer,
`w AAw BB,
`and with the presence of hydrogen bonding the entropy of mixing is no longer
`ideal. In addition, Vmix will not equal zero if the dimensions of the solute and
`solvent molecules are very different.
`Modifications to the Hildebrand solubility parameter model have been ad-
`vanced in attempts to achieve better degrees of solubility prediction (Taft et al.,
`1969; Rohrschneider, 1973). Among these, the three-dimensional solubility pa-
`rameter introduced by Hansen and Beerbower (1971) showed the most practical
`application. These workers calculated the total solubility parameter (δtotal) using
`three partial parameters, δD, δP, and δH:
`= δ2
`δ2
`D
`P
`H
`total
`where the parameters δD, δP, and δH account for dispersion, polar, and hydrogen-
`bonding interactions, respectively. Some of the values deduced for δD, δP, δH,
`and δtotal are listed in Table 2. Another modification of Hildebrand solubility
`parameter considered the effects of polar interaction and hydrogen bonding,
`and was found to yield good solubility predictions in many cases (Kumar and
`Prausnitz, 1975). However, the modified Hildebrand solubility equation can only
`be used empirically in predicting solubility in polar solvents, since the original
`assumptions associated with regular solutions do not apply in polar solvents
`(Grant and Higuchi, 1990).
`In solvent systems where polar interactions exert a major role, the molecular
`and group-surface-area (MGSA) approach provides a better quality solubility
`prediction (Yalkowsky et al., 1972, 1976; Amidon et al., 1974, 1975). Instead of
`
`+ δ2
`
`+ δ2
`
`(39)
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`Solvents
`
`n-Butane
`
`n-Hexane
`
`n-Octane
`
`Diethyl ether
`
`Cyclohexane
`
`n-Butyl acetate
`
`Carbon tetrachloride
`
`Toluene
`
`Ethyl acetate
`
`Benzene
`
`Chloroform
`
`Acetone
`
`Acetaldehyde
`
`Solubility parameter (cal/cm3)1/2
`
`δD
`
`6.9
`
`7.3
`
`7.6
`
`7.1
`
`8.2
`
`7.7
`
`8.7
`
`8.8
`
`7.7
`
`9.0
`
`8.7
`
`7.6
`
`7.2
`
`δP
`
`0
`
`0
`
`0
`
`1.4
`
`0
`
`1.8
`
`0
`
`0.7
`
`2.6
`
`0
`
`1.5
`
`5.1
`
`3.9
`
`δH
`
`0
`
`0
`
`0
`
`2.5
`
`0.1
`
`3.1
`
`0.3
`
`1.0
`
`3.5
`
`1.0
`
`2.8
`
`3.4
`
`5.5
`
`δtotal
`
`6.9
`
`7.3
`
`7.6
`
`7.7
`
`8.2
`
`8.5
`
`8.7
`
`8.9
`
`8.9
`
`9.1
`
`9.3
`
`9.8
`
`9.9
`
`Carbon disulfide
`
`10.0
`
`Dioxane
`
`1-Octanol
`
`Nitrobenzene
`
`1-Butanol
`
`1-Propanol
`
`Dimethylformamide
`
`Ethanol
`
`Dimethyl sulfoxide
`
`Methanol
`
`Propylene glycol
`
`Ethylene glycol
`
`Glycerin
`
`Formamide
`
`9.3
`
`8.3
`
`9.8
`
`7.8
`
`7.8
`
`8.5
`
`7.7
`
`9.0
`
`7.4
`
`8.2
`
`8.3
`
`8.5
`
`8.4
`
`0
`
`0.9
`
`1.6
`
`4.2
`
`2.8
`
`3.3
`
`6.7
`
`4.3
`
`8.0
`
`6.0
`
`4.6
`
`5.4
`
`5.9
`
`12.8
`
`0.3
`
`3.6
`
`5.8
`
`2.0
`
`7.7
`
`8.5
`
`5.5
`
`9.5
`
`5.0
`
`10.9
`
`11.4
`
`12.7
`
`14.3
`
`9.3
`
`10.0
`
`10.0
`
`10.3
`
`10.9
`
`11.3
`
`12.0
`
`12.1
`
`13.0
`
`13.0
`
`14.5
`
`14.8
`
`16.1
`
`17.7
`
`17.9
`
`20.7
`7.8
`7.6
`Water
`Table 2. Solubility parameters for some common solvents.
`Reproduced from:
`Hansen C, and Beerbower A. Solubility Parameters. In: Standen A. Kirk-Othmer
`Encyclopedia of Chemical Technology, 2nd ed. Supplement Volume. New York, NY: Wiley;
`1971. 889–910.
`
`23.4
`
`13
`
`

`

`P1: GFZ
`SVNY358-Augustijns
`
`March 15, 2007
`
`16:46
`
`14
`
`Chapter 1: Principles of Solubility
`
`the potential energy term that was used in equation (32), a free energy model
`was used in the MGSA approach to represent the change of the interactions at
`mixing. The power of this approach is that changes in enthalpy and entropy are
`included:
`
`(40)
`
`ln γB = (WAA + WBB − 2WAB)
`VBφ2
`A
`RT
`In equation (40), W is reversible work which represents the internal free en-
`ergy. Yalkowsky et al. (1976) used the molar surface area (A) and the surface
`tension (σ ) to replace the molar volume (V ) and reversible work. Under those
`circumstances, equation (40) becomes:
`ln γB = σAB AB
`kT
`where σ A and σ B are the surface energies of the pure liquids A and B, while
`σ AB is the interfacial energy between the two liquids. The interfacial tension can
`be experimentally measured for substances of different polarity, and therefore
`equation (41) better predicts solubility in polar solvents.
`
`(41)
`
`Intermolecular Interactions in Non-Ideal Solutions
`
`Pr

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