throbber
_g___=_u_s___§_._E53222:
`
`”as“
`
`53.5.oea:38:8
`
`
`
`65watchman:ALE—bums
`
`
`
`
`
`>630...392.$333.
`
`
`
`
`
`Swim.v35?BmZ.ozH.mmmumfl$332EAmUfl<E
`
`mwmxmn—
`
`Lupin Ex. 1030 (Page 1 of 19)
`Lupin Ex. 1030 (Page 1 of 19)
`
`

`

`ISBN: 0-8247-0237-9
`
`This book is printed on acid-free paper.
`
`Headquarters
`Marcel Delcker, Inc.
`270 Madison Avenue, New York, NY 10016
`tel: 212-696-9000; fax: 212-685-4540
`
`Eastern Hemisphere Distribution
`Marcel Dekker AG
`Hutgasse 4, Posffach 812, CH-4001 Basel, Switzerland
`tel: 44-61-261-8482; fax: 44-61-261-8896
`
`World Wide Web
`htlp:t/www.dek.ker.com
`
`The publisher offers discounts on tlxis book when ordered in bulk quantifies. For more
`information, write to Special Sales/Professional Marketing at the headquarters address
`above.
`
`Copyright © 1999 by Marcel Dekker, Inc. All Rights Reserved.
`
`m
`
`Neithe~ this book nor any part may be reproduced or transmitted in any form or by
`any means, electronic or mechanical, including photocopying, microfilming, and re-
`cording, or by any information storage and retrieval system, without permission in
`writiug from the publisher.
`
`Current printing (last digit):
`10 9 8 7 6 5 4 3 2 I
`
`PRINTED IN THE UNITED STATES OF AMERICA
`
`Preface
`
`Since the middle of the last century, it has been noted that organic
`molecules can be obtained in more than one distinct crystal form, a
`property that became known as polymorphism. Once experimental
`methods based on the diffraction of x-rays were developed to determine
`the structures of crystalline substances, it was quickly learned that an
`extremely large number of molecuies were capable of exhibiting the
`phenomenon. In addition, numerous compotmds were shown to form
`other nonequivalent crystzlline structures through the inclusion of sol-
`vent molecuIes in the lattice.
`It was also established that the structure adopted by a given com-
`pound upon crystallization would exert a profound effect on the solid-
`state properties of that system. For a given material, the heat capacity,
`conductivity, volume, density, viscosity, surface tension, diffttsivity;
`crystal hardness, crystal shape and color, refractive index, electrolytic
`conductivity, melting or sublimation properties, latent heat of fusion,
`heat of solution, solubility, dissolution rate, enthalpy of transitions,
`phase diagrams, stability, hygroscopicity, and rates of reactions are all
`determined primarily by the nature of the crystal structure.
`
`iii
`
`

`

`1
`
`Theory and Origin. of Polymorphism
`
`David J. W. Grant
`
`University of Minnesota
`Minneapolis, Minnesota
`
`I.
`
`INTRODUCTION
`
`II. THERMODYNAMICS OF POLYMORPHS
`
`l-fL ENANTIOTROPY AND MONOTROPY,
`
`IV, KINETICS OF CRYSTALLIZATION
`
`V. NUCLEATION O1~ POLYMORPHS
`
`VI. NEW OR DISAPPEARING POLYMORPHS
`
`REFERENCES
`
`t
`
`10
`
`18
`
`19
`
`25
`
`31
`
`I.
`
`INTRODUCTION
`
`Many pharmaceutical solids exhibit polymorphism, which is frequently
`defined as the ability of a substance to exist as two or more crystalline
`phases that have different arrangements and/or conformations of the tool-
`
`m
`
`

`

`Grant
`
`(a)
`
`2
`
`(b)
`
`(b)
`
`Fig. 1 Molecular structure of (a) acetarninophen and (b) spiperone.
`
`ecutes in the crystal lattice [ 1-3]. Thus, in the strictest sense, polymorphs
`are different crystalline forms of the same pure substance in which the
`molecules have different arrangements and/or different conformations
`of the molecules. As a result, the polymorphic solids have different unit
`cells and hence display different physicalproperties, including those due
`to packing, and various thermodynamic, speclroscopic, interracial, and
`mechanical properties, as discussed below [1-3].
`For example, acetaminophen (paracetamol, 4-acetamidophenol,
`4-hydroxyacetanilide, shown in Fig. la) can exist as a monoclinic form,
`of space group P21/n [4], which is thermodynamically stable under
`ambient conditions. The compound can also be obtained as a less stable
`o~daorhombic form, of space group Pbca, and which has a higher den-
`sity indicative of closer packing [5-7]. The unit celts of these two forms
`are compared in Fig. 2 and Table 1. The molecule of acetaminophen
`is rigid on account of resonance due to conjugation involving the by-
`
`Fig. 2 View of the unit cell contents for two polymorphs of acetaminophen:
`(a) orthorhombic form (b) monoclinic form [4,5,7]. (Reproduced w~th permis-
`sion of the copyright owner, the American Crystallographic Association,
`Washington, DC.)
`
`m
`x
`
`

`

`4
`
`Grant
`
`Theory and Origin of Polymorphism
`
`5
`
`Table 1 Crystal Data for Two Polymorphs of Acetaminophen
`
`Form I
`
`Crystal data and
`structure refinement Orthorhombic phase Monoclhaic phase
`
`Empirical formula
`Formula weight
`Crystal system
`Space group
`Unit cell dimensions
`
`Volume
`Z
`Density (calculated)
`Crystal s~ze
`Refinement method
`
`Hydrogen bond
`length§ and angles
`
`C~HgNO~
`151.16
`Orthorhombic
`Pbca
`a = 17.1657(12) ~
`b = 11.7773(11) ,~
`c = 7,212(2) 2~
`ot = 90.000°
`,6 = 90,000°
`7/ = 90.000°
`1458.1(4) ,~3
`8
`1.377 g/cm3
`0.28 × 0.25 × 0.15 mm
`Pull-matrix least-squares
`on F2
`
`CsH~NO~
`151.16
`Monoclinic
`P21/n
`a = 7.0941(12) ~
`b = 9.2322(1i) ~
`c = 11.6196(10)
`~ ~- 90,000°
`,6--- 97.821(10)°
`7’ = 90.000°
`753.9(2) 2~~
`4
`1.332 g/cm3
`0.30 × 0.30 × 0.15 mm
`Full-matrix least-squares
`on F~
`
`H(5)O(2)
`H(6)O(1)
`O(1)--H(5)O(2)
`N(1)--H(6)O(1)
`
`1.852(26) ~ 1.772(20) ~
`2.072(28) ~
`2.007(18) ~
`170.80(2.35)°
`166.15(1.75)°
`163.93(1.51)°
`163.52(2.19)°
`
`Source: Refs. 4, 5, and 7. Reproduced with permission of the copyright owner, the
`American Crystallographic Association, Washington, DC.
`
`droxyl group, the benzene ring, and the amido group. Therefore the
`conformation of the molecule is virtually identical in the two poly-
`morphs of acetaminophen. On the other hand, the spiperone molecule
`(8-[3-(p-fluorobenzoyl)-propyl]-l-phenyl-l,3,8-triazaspiro[4,5]decan-
`4-one, shown in Fig. lb) contains a flexibIe -CH2-CH2-CH2- chain and
`is therefore capable of existing in different molecular conformations
`[8]. Two such conformations, shown in Fig. 3, give rise to two different
`conformafional polymorphs (denoted Forms I and EO, which have dif-
`ferent unit cells (one of which is shown in Fig. 4) and densities, even
`
`m
`x
`
`Form II
`
`Fig, 3 The molecular conformations of the spiperone molecule in potymor-
`phic forms I and II [8], (Reproduced with permission of the copyright owner,
`the American Pharmaceutical Association, Washington, DC.)
`
`though their space groups are the same, both being P21/n, monoclirdc,
`as shown in Table 2 [8].
`As mentioned above, the various poIymorphs of a substance can
`exhibit a variety of different physical properties. Table 3 lists some of
`the many properties that differ among different polymorphs [1-3,9].
`Because of differences in the dimensions, shape, symmetry, capacity
`
`

`

`6
`
`Grant
`
`Theory and Origin of Polymorphism
`
`7
`
`Table 3 List of Physical Properties that Differ Among VazSous
`Polymorphs
`
`1. Packing properties
`a. Molar volume and density
`b. Refractive index
`c. Conductivity, electrical and thermal
`d. Hygroscopicity
`2. Thermodynamic properties
`a. Melting and sublimation temperatures
`Internal energy (i.e., Structural energy)
`b.
`c. Enthalpy (i.e., Heat content)
`d. Heat capacity
`e. Entropy
`f. Free energy and chemical potential
`g. Thermodynamic activity
`lx Vapor presstlre
`i.
`Solubility
`3. Spectroscopic properties
`Electronic transitions (i.e., ultraviolet-visible absorption spectra)
`Vibrational transitions (i.e., infrared absorption spectra and Raman
`spectra)
`Rotational transitions (i.e., far infrared or microwave absorption
`spectra)
`Nuclear spin transitions (i.e., nuclear magnetic resonance spectra)
`d.
`4. Kinetic properties
`a. Dissolution rate
`b. Rates of solid state reactions
`c. Stability
`5. Surface p~operties
`a. Surface free energy
`Interracial tensions
`b.
`c. Habit (i.e., shape)
`6. Mechanical properties
`a. Hardness
`b. Tensile strength
`c. Compactibility, tableting
`d. Handling, flow, and blending
`
`b.
`
`Fig. 4 View of the unit celJ contents for the form I polymorph of spiperone
`[8]. (Reproduced with permission of the copyright owner, the American Phar-
`maceutical Association, Washington, DC.)
`
`Table 2 Crystal Data for Two Polymorphs of Spiperone
`
`Empirical formula
`Molecular weight
`.Crystal system
`Space group
`Unit cell dimensions
`
`Unit cell volume
`Z
`
`Form i
`
`C:3H:6FN~O2
`395.46
`Monoclinic
`P21/a
`a = 12.722 ~
`b = 7.510 ~
`c = 21.910 ~
`a = 90.00°
`fl = 95.08°
`7’= 90.00°
`2085.1 ~
`4
`
`Form
`
`C23H26FN302
`395.46
`Monoclinie
`P21/c
`a = 18.571 ,~
`b = 6.072 ~
`c = 20.681 ,~
`~ = 90.00°
`/3 = 118.69°
`~/= 90.00°
`2045.7 ~3
`4
`
`Source: Ref. 8. Reprodnced with permission of the copyright owner, the American
`Pharmaceutical Association, Washington, DC.
`
`m
`x
`
`o
`
`

`

`8
`
`Grant
`
`Theory and Origin of Polymorphism
`
`9
`
`(number of molecules), and void volumes of their unit cells, the differ-
`ent polymorphs of a given substance have different physical properties
`arising from differences in molecular packing. Such properties include
`molecular volume, molar volume (which equals the molecular volume
`multiplied by Avogadro’s number), density (which equals the molar
`mass divided by the molar volume), refractive index in a given direc-
`tion (as a result of the interactions of light quanta with the vibrations
`of the electrons in that direction), thermal conductivity (as a result of
`the interaction of infrared quanta with the intramolecular and intermo-
`lecular vibrations and rotations of the molecules), electrical conductiv-
`ity (as a result of movement of the electrons in an electric field), and
`hygroscopicity (as a result of access of water molecules into the crystal
`and their interactions with the molecules of the substance). Differences
`in melting point of the various polymorphs arise from differences of
`the cooperative interactions of the molecules in the solid state as com-
`pared with the liquid state. Differences in the other thermodynamic
`properties among the various polymorphs of a given substance are dis-
`cussed below. Also involved are differences in spectroscopic proper-
`ties, kinetic properties, and some surface properties. Differences in
`packing properties and in the energetics of the intermolecular interac-
`tions (thermodynamic properties) among polymorphs give rise to dif-
`ferences in mechanical properties.
`Many pharmaceutical solids can exist in an amorphous form,
`which, because of its distinctive properties, is sometimes regarded as
`a polymorph. However, unlike true polymorphs, amorphous forms are
`not crystalline [ 1,2,10]. In fact, amorphous solids consist of disordered
`arrangements of molecules and therefore possess no distinguishable
`crystal lattice nor unit cell and consequently have zero crystallinity. In
`amorphous forms, the molecules display no long-range order, although
`the-short-range intermolecular forces give rise to the short-range 6~Ier
`typical of that between nearest neighbors (see Fig. 5). Thermodynami-
`cally, the absence of stabilizing lattice energy causes the molar internal
`energy or molar enthalpy of the amorphous form to exceed that of the
`crystalline state. The absence of long-range order causes the molar en-
`tropy of the amorphous form to exceed that of the crystalline state.
`Furthermore, the lower stability and greater reactivity of the amorphous
`form indicates that its molar Gibbs free energy exceeds that of the crys-
`
`m
`x
`
`(a)
`
`Fig. 5 Schematic diagram showing the difference in long-range order of
`silicon dioxide in (a) the crystnlline state (crystobalite) and (b) the amorphous
`state (silica glass) [2]. The two forms have the same short-range order. (Repro-
`duced with permission of the copyright owner, the American Pharmaceutical
`Association, Washington, DC.)
`
`

`

`10
`
`Grant
`
`Theory and Origin of Polymorphism
`
`11
`
`talline state. This observation implies that the increased molar enthalpy
`of the amorphous form outweighs the TAS term that arises from its
`increased molar entropy.
`
`I!. THERMODYNAMICS OF POLYMORPHS
`
`The energy of interaction between a pair of molecules in a solid, liquid,
`or real gas depends on the mean intermolecular distance of separation
`according to the Morse potential energy curve shown in Fig. 6 [11,12].
`For a given pair of molecules, each polymorph, liquid or real gas has
`its own characteristic interaction energies and Morse curve. These in-
`termolecular Morse curves are similar in shape but have smaller ener-
`gies and greater distances than the Morse potential energy curve for the
`interaction between two atoms linked by a covalent bond in a diatomic
`molecule or within a functional group of a polyatomic molecule. The
`Morse potential energy curve in Fig. 6 is itself the algebraic sum of a
`curve for intermolecular attraction due to van der Waals forces or hy-
`drogen bonding and a curve for intermotecular electron-electron and
`nucleus-nucleus repulsion at closer approach. The convention em-
`ployed is that attraction causes a decrease in potential energy, whereas
`repulsion causes an increase in potential energy. At the absolute zero
`of temperature, the pair of molecules would occupy the lowest or zero
`point energy level. The Heisenberg uncertainty principle requires that
`the molecules have an indeterminate position at a defined momentum
`or energy. This indeterminate position corresponds to the familiar vi-
`bration of the molecules about the mean positions that define the mean
`intermolecular distance. At a temperature T above the absolute zero,
`a proportion of the molecules will occupy higher energy levets ac-
`cording to the Boltzmann equation:
`
`where N~ is the number of molecules occupying energy level 1 (for
`which the potential energy exceeds the zero point level by the energy
`difference A~I), No is the number of molecules occupying the zero point
`
`(1)
`
`rn
`x
`
`0.0
`0
`
`Fig, 6 Morse potential energy curve of a given condensed phase, solid or
`liquid [11]. The potential energy of interaction V is plotted against the mean
`intermolecular distance d. (Reproduced with permission of the copyright
`owner, Oxford University Press, Oxford, UK.)
`
`

`

`12
`
`Grant
`
`Theory and Origin of Polymorphism
`
`13
`
`level, and k is the Boltzmann constant (1.381 × 10-a3 J/K, or 3.300
`× 10-2~ calIK, i.e. the gas constant per molecule).
`With increasing temperature, increasing numbers of molecules
`occupy the higher energy levels so that the distribution of the molecules
`among the various energy levels (lcnown as the Boltzmanu distribution)
`becomes broader, as shown in Fig. 7. At any given temperature, the
`number of distinguishable arrangements of the molecules of the system
`among the various energy levels (and positions in space) available to
`them is termed the .th_e.rm0.dynamic pr.obability_~.l-2~. With increasing tem-
`perature, ffZ increases astron0mically. According to the Boltzmann
`equation,
`
`S = k. In f2
`
`(2)
`
`where the entropy S is a logarithmic function of ~2, so increasing tem-
`perature causes a steady rise, though not an astronomical rise, in the
`entropy. In a macroscopic system, such as a given polymorph, the prod-
`uct T ¯ S represents the energy of the system that is associated with
`
`the disorder of the molecules. This _energy_is ~e bound energy_.gf...th_e_ .....
`sy~s_t~em that is unavailable for doing work.
`The sum of ~£Jindividual ~ii~rgi~~m~eracfion between nearest
`neighbors, next nearest neighbors, and so on, throughout the entire
`crystal lattice, Liquid, or real gas can be used to define the internal
`energy E (i.e., the intermolecular structural energy) of tli~ phase. Nor-
`mally the interactions beyond next nearest neighbors are weak enough
`to be approximated or even ignored. For quantitative convenience one
`mole of substance is considered, corresponding to molar thermody-
`namic quantities. At constant pressure P (usually equal to atmospheric
`pressure), the total energy of a phase is represented by the enthalpy H:
`H = E + P. V
`(3)
`where V is the volume of the phase (the other quantifies have already
`been defined). With increasing temperature, E, V, and H tend to in-
`crease.
`Figure 8 shows that the enthalpy H and the entropy S of a phase
`
`rn
`x
`
`L
`|
`


`

`
`I
`

`
`i
`~
`
`Go;Ho
`
`-
`
`H
`
`TS
`
`j G-H-~-$
`
`0 AbsoluLe Temperature, T
`
`Fig. 7 Populations of molecular states at various temperatures [11]. The
`temperature is increasing from left to right. (Reproduced with permission of
`the copyright owner, Oxford University Press, Oxford, UK.)
`
`Fig. 8 Plots of various thermodynamic quantities against the absolute tem-
`perature T of a given solid phase (polymorph) or liquid phase at constant
`pressure. H = enthalpy, S = entropy, and G = Gibbs free energy.
`
`

`

`14
`
`Grant
`
`Theory and Origin of Polymorphism
`
`15
`
`tend to increase with increasing absolute temperature T. According to
`the third law of thermodynamics, the entropy of a perfect, pure crys-
`talline solid is zero at the absolute zero of temperature. The product
`T. S increases more rapidly with increasing temperature than does H.
`Hence the Gibbs free energy G, which is defined by~
`
`G =H- T’S
`
`:
`
`(4)
`
`tends to decrease with increasing temperature (Fig. 8)’. This decrease
`also corresponds to the fact that the slope (6G/ST) ofthe plot of G
`against T is negative according to the equation
`
`P
`
`@
`I1
`t
`!
`
`I
`
`E
`n
`e
`r
`g
`y
`
`Liquid
`
`A
`
`B
`Potymorphs ,
`
`C
`
`(5)
`
`blean Intermolecular Distance
`
`As already stated, the entropy of a perfect, pure crystalline solid is zero
`at the absolute zero of temperature. Hence the value of G at T = 0
`(termed Go) is equal to the value of H at T = 0, termed H0 (Fig. 8).
`Each polymorph yieIds an energy diagram similar to that of Fig. 6,
`although the values of G, H, and the slopes of the curves at a given
`temperature are expected to differ between different polymorphs.
`~’Becanse each polymorph has its own distinctive crystal lattice, it
`has its own distinctive Morse potential energy curve for the dependence
`of the intermolecular interaction energies with intermolecular distance.
`The liquid state has a Morse curve with greater intermolecular energies
`and distances, because the liquid state has a higher energy and molar
`volume (lower density) than does the solid state. Figure 9 presents a
`series of Morse curves, one for each polymorph (A, B, and C) and for
`the liquid state of a typical substance of pharmaceutical interest. The
`composite curve in Fig. 9 is the algebraic sum of the Morse curves for
`each phase (polymorph or liquid). The dashed line corresponds to the
`potential energy of the separated, non_interacting molecules in the gas-
`eous state. The increase in potential energy from the zero point value
`of a given polymorph to the dashed line corresponds to the lattice en-
`ergy of that polymorph or energy of sublimation (if at constant pres-
`sure, the enthalpy of vaporization). For the liquid state the increase in
`potential energy from the average value in the liquid state to the dashed
`line for the gaseous molecules corresponds to the energy of vaporiza-
`
`m
`x
`
`Fig. 9 Composite Morse potential energy curve of a series of polymorphs,
`A, B, and C, and of the corresponding liquid phase.
`
`tion (if at constant pressure, the enthalpy of vaporization). The increase
`in potential energy from the zero point value of a given polymorph to
`the average value for the liquid state corresponds to the energy of fusion
`(if at constant pressure, the enthalpy of fusion).
`When comparing the thermodynamic properties of polymorph 1
`and polymorph 2 (or of one polymorph 1 and the liquid state 2) the
`difference notation is used:
`
`AG= G2- G1
`
`AS= $2-S~
`
`AV = Vz - Vl
`
`(6)
`
`(7)
`
`(9)
`
`¯ In discussions of the relative stability of polymorphs and the driving
`force for polymorphic transformation at constant temperature and pres-
`sure (usually ambient conditions), the difference in Gibbs free energy
`is the decisive factor and is given by
`
`AG = AH - T AS
`
`(10)
`
`

`

`16
`
`Grant
`
`Theory and Origin of Polymorphism
`
`17
`
`of temperature increase must be slow enough to allow polymorph 1 to
`change completely to polymorph 2 over a few degrees. Because in Fig.
`10, Ha > H~, M!is positive and the transition is endothermic in nature.
`Figure 10 shows that, below T~, polymorph 1 (or the solid) has
`the Iower Gibbs free energy and is therefore more stable (i.e., G2 >
`G1). On the other hand, above Tt, polymorph 2 (or the Hquid) has the
`lower Gibbs free energy and is therefore more stable (i.e., G2 < G1).
`Under defined conditions of temperature and pressure, only one poly-
`morph can be stable, and the other polymorph(s) are unstable. If a phase
`is unstable but transforms at an imperceptibly low rate, then it is some-
`times said to be metastable.
`The Gibbs free energy difference AG between two phases reflects
`the ratio of "e.scaping tendencies" of the two phases. The escaping
`tendency is termed the fugacity f-fffid is approximated by the saturated
`vapor pressure, p. Therefore
`
`where the subscripts 1 and 2 refer to the respective phases, R is the
`universnl gas constant, and T is the absolute temperature. The fugacity
`is proportionai to the thermodynamic activity a (where the constant of
`proportionality is defined by the standard state), whiIe thermodynamic
`activity is approximately proportional to the solubility s (in any given
`solvent) provided the laws of dilute solution apply. Therefore
`
`~G= RTln(~)
`
`(14)
`
`i
`
`!
`
`61
`
`~bsoluLe Temper’aLu~’e. T
`
`Fig, 10 Plots of the Gibbs free energy G and the enthalpy H at constant
`pressure against the absolute temperature T for a system consisting of two
`polymorphs, i and 2 (or a solid, 1, and a liquid, 2). Tz is the transition tempera-
`ture (or melting temperature) and S is the entropy.
`
`Figure 10 shows the temperature dependence of G and H for two
`different polymorphs 1 and 2 (or for a solid 1, corresponding to any
`polymorph, and a liquid 2) [13]. In Fig. 10 the free energy curves cross.
`At the point of intersection, known as the transition temperature Tt (or
`the melting point for a solid and a liquid), the Gibbs free energies of
`the two phases are equnl, meaning that the phases 1 and 2 are in equilib-
`rium (i.e., AG = 0). However, at Tt Fig. 10 shows that polymorph 2
`(or the liquid) has an enthalpy H2 that is higher than that of polymorph
`1 (or the solid), so that H2 > H1. Equations 10 and 6 show that, if AG
`= 0, polymorph 2 (or the liquid) atso has a higher entropy S: than does
`polymorph 1 (or the solid), so that S~ > $1. Therefore according to
`Equation 10, at T~,
`
`~,/-/t = Tt ASt
`
`(11)
`
`i’n
`x
`
`where i~/t = H~ - Ht and ASt = Sz - $l at Tt. By means of differential
`scanning calorimetry, the enthalpy transition AHt (or the enthalpy of
`fusion M-!f) may be determined. For a polymorphic transition, the rate
`
`in which the symbols have been defined above. Hence, because the
`most stabIe polymorph under defined conditions of temperature and
`pressure has the lowest Gibbs free energy, it also has the Iowest values.
`
`

`

`18
`
`Grant
`
`Theory and Origin of Polymorphism
`
`19
`
`of fagacity, vapor pressure, thermodynamic activity, and solubility in
`
`trolled under sink conditions and under constant conditions of hydrody-
`namic flow, the dissolution rate per unit surface area J is proportional to
`the solubility according to the Noyes-Whitney [14] equation; therefore
`
`AG= RTln(~) (16)
`
`According to the law of mass action, the rate r of a chemical reaction
`(including the decomposition rate) is proportional to the thermody-
`namic activity of the reacting substance. Therefore
`
`AG= RTln(r~) (17)
`
`To summarize, the most stable polymorph has the lowest Gibbs free
`energy, fugacity, vapor pressure, thermodynamic activity, solubility,
`and dissolution rate per unit surface area in any solvent, and rate of
`reaction, including decomposition rate.
`
`I!1. ENANTIOTROPY AND MONOTROPY
`
`If as shown in Fig. 10 one polymorph is stable (i.e., has the lower
`free energy content and solubility over a certain temperature range and
`pressure), while another polymorph is stable (has a lower free energy
`and solubility over a different temperature range and pressure), the two
`polymorphs are said to be enantiotrope_~, and the system of the two
`solid phases is said to be ena~.tiotropico For an enantiotropic system a
`reversible transition can be observed at a definite transition tempera-
`ture, at which the free energy curves cross before the melting point
`is reached. Examples showing such behavior include acetazolamide,
`carbamazepine, metochlopramide, and tolbutamide [9,14,15].
`Sometimes only one polymo~rph is stable at all temperatures below
`the melting point, with alt other polymorphs being, therefore-marble.
`These polymorphs are said to be monotropes, and the system of the
`two solid phases is said to be monotropic. For a monotropic system
`
`the free energy curves do not cross, so no reversible transition can be
`observed below the meIting point. The polymorph with the higher free
`energy curve and solubility at any given temperature is, of course, al-
`ways the unstable polymorph. Examples of this type of system include
`chloramphenicol palrnitate and metolazone [9,14,15].
`To help decide whether two polymorphs are enhntiotropes or
`monotropes, Burger and Ramberger developed four thermodynamic
`rules [14]. The application of these rules was extended by Yn [15].
`The most useful and applicable of the thermodynamic rules of Burger
`and Ramberger are the heat of transition rule and the heat of fusion
`rule. Figure 11, which includes the liquid phase as well as the two
`polymorphs, illustrates the use of these rules. The heat of fusion rule
`states that, if an endothermic polymorphic transition is observed, the
`two forms are enantiotropes. Conversely, if an exothermic polymorphic
`transition is observed, the two forms are monotropes.
`The heat of fusion rule states that, if the higher melting polymorph
`has the lower heat of fusion, the two forms are enantiotropes. Con-
`versely, if the higher melting polymorph has the higher heat of fusion,
`the two forms are monotropes. Figure 11, which inciudes the liquid
`phase as we11 as the two polymorphs, is necessary to illustrate the heat
`of fusion rule.
`The above conditions, that are implicit in the thermodynamic
`rules, are summarized in Table 4. The last two rules in Table 4, the
`infrared rule and the density rule, were found by Burger and Ramberger
`[14] to be significantly Iess reliable than the heat of transition rule and
`the heat of fusion rule and are therefore not discussed here.
`
`IV. KINETICS OF CRYSTALLIZATION
`
`Among the various methods for preparihg different polymorphs are
`sublimation, crystallization from the melt, crystallization from super-
`critical fluids, and crystallization from liqNd solutions. In the pharma-
`ceuticai sciences, different polymorphs are usually prepared by crys-
`tallization from soIution employing various solvents and various
`temperature regimes, such as initial supersaturation, rate of de-super-
`saturation, or final supersaturation. The supersaturation of the solution
`
`rn
`x
`
`

`

`2O
`
`(a)
`
`(b)
`
`m
`x
`
`Grant
`
`Theory and Or|g!n of Polymorphism
`
`21
`
`Table 4 Thermodynamic Rules for PoIymorphic Transitions According to
`Burger and Ramberger [14],AVhere Form I is the Higher-Melting Form
`
`En~nfiotropy-
`
`Transition, < melting I
`I Stable > transition
`II Stable .< transition
`Transition reversible
`Solubility i higher < transition
`,Solubility 1 lower > transition
`Transition II -~ I is endothermic
`
`Monotropy
`
`Transition > melting I
`I always stable ..
`
`Transition irreversible
`Solubikity I always lower than II
`
`Transition II --~ I is exothermic
`
`IR peak I before II
`Density I < density II
`
`IR peak I after II
`Density I > density Ii
`
`Source: Reproduced from Refer. 9 with permission.of the copyright owner, Elsevier,
`Amsterdam~ The Netherlands.
`
`that is necessary for crystalhzation may be achieved by evaporation of
`the solvent (although any impurities will be concentrated), coohng the
`solution from a known initial supersaturation (or heating the solution if
`the-heat of solution is exothermic), addition of a Poor solvent (sometimes
`termed a precipitant), chemical reaction between two or more solubte
`species, or variation of pH to produce a less solubIe acid or base from a
`salt or vice versa (while minimizing other changes in composition).
`During the 19~th century, Gay Lussac observed that, during crys-
`tallization, an unstabIe form is frequently obtained first that subse-
`quently Iransforms into a stable form [13]. This observation was later
`explained thermodynamically by Ostwald [13,16-19], who formulated
`the law of successive reactions, also known as Ostwald’s step rule. This
`
`Fig. 11 Plots of the Gibbs free energy G and the enthalpy H at constant
`pressure against the absolute temperature T for a system consisting of two
`polymorphs, A and B, and a liquid phase, i [14]. T~ is the transition tempera-
`ture, Tf is the melting temperature, and S is the entropy for (a) an enanfiotropic
`system and (b) a monotropic system. (Reproduced with permission of the
`copyright owner, Springer Verlag, Vienna~ Austria.)
`
`

`

`22
`
`Grant
`
`Theory and Origin of Polymorphism
`
`23
`
`rule may be stated as, ’ ’In all processes, it is not the most stable state with
`the lowest amount of free energy that is initially formed, but the least
`stable state lying nearest in free energy to the original state [ 13]."
`Ostwald’s step rule [13,16-19] is illustrated by Fig. 12. Let an
`enantiotropic system (Fig. 12a) be initially in a state represented by
`point X, corresponding to an unstable vapor or liquid or to a supersatu-
`rated solution. If this system is cooled, the Gibbs free energy will de-
`
`(a)
`
`TF~PERATORE
`
`rn
`x
`
`TEMPERATURE
`
`Fig. 12 Relationship between the Gibbs f~ee energy G and the temperature
`T for two polymorphs for (a) an enantiotropic system and (b) a monolropic
`system in which the system is cooled from point X [9]. The arrows indicate
`the direction of change. (Reproduced with permission of the copyright owner,
`Elsevier, Amsterdam, The NetherIands.)
`
`crease as the temperature decreases. When the state of the system
`reaches point Y, form 13 will tend to be formed instead of form A,
`because according to Ostwald’s step ruIe Y (not Z) is the least stable
`state lying nearest in free energy to the original state. Similarly, let a
`monotropic system (Fig. 12b) be initially in a state represented by point
`X, corresponding to an unstable vapor or liquid or to a~supersaturated
`solution. If this system is cooled, the Gibbs free energy will decrease
`as the temperature decreases. When the state of the system reaches
`point Z, form A will tend to be formed instead of form B, because
`according to Ostwald’s step rule Z (not Y) is now the 1east stable state
`Iying nearest in free energy to the original state. This rule is not an
`invariable thermodynamic law but a useful practical rule that is based
`on kinetics, and it is not always obeyed.
`An understanding of the kinetics of the crystallization process
`involves consideration of the various steps involved. In the first step
`(termed nucleation) tiny crystallites of the smallest size capable of inde-
`pendent existence (termed nuclei) are formed in the supersaturated
`phase. Molecules of the crystallizing phase then progressively attach
`themseIves to the nuclei, which then grow to form macroscopic crystals
`in the process known as crystaI growth, until the crystallization medium
`is no longer supersaturated because saturation equilibrium has now
`been achieved. If the crystals are now allowed to remain in the saturated
`medium, the smaller crystals, which have a slightly greater solubility
`according to the Thomson (Kelvin) equation [11,20], tend to dissolve.
`At the same time, the larger crystals, which consequently have a lower
`solubility, tend to grow. This process of the growth of larger crystN_~..at
`the expense of smaller crystals is sometimes termed Ostwald ripening.
`The nucleation step is the most critical for the production of dif-
`ferent polymorphs and is therefore discussed in some detail below. Nu-
`cleation may be primary (which does not require preexisting crystals
`of the substance that crystallizes) or secondary (in which nucleation is
`induced by preexisting crystals of the substance). Primary nucleation
`may be homogeneous, whereby the nuclei of the crystallizing substance
`arise spontaneously in the medium in which crystal

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