throbber
Pharmaceutics: The Science of
`Dosage Form Design
`
`EDITED BY
`Michael E. Aulton BPharm PhD MPS
`
`Reader in Pharmacy, Leicester Polytechnic, Leicester, UK
`
`CHURCHILL LIVINGSTONE
`EDINBURGH LONDON MELBOURNE AND NEW YORK 1988
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 1 of 72
`
`

`
`PRy OF CON%
`trcp\
`5 (cid:9)
`NOV 2 5 1987
`coPY
`—.
`C1P
`
`CHURCHILL LIVINGSTONE
`Medical Division of Longman Group UK Limited
`Distributed in the United States of America by
`Churchill Livingstone Inc., 1560 Broadway, New
`York, N.Y. 10036, and by associated companies,
`branches and representatives throughout the world.
`
`C) Michael Aulton 1988
`
`All rights reserved. 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 or otherwise,
`without the prior permission of the publishers
`(Churchill Livingstone, Robert Stevenson House, 1-3
`Baxter's Place, Leith Walk, Edinburgh EH1 3AF).
`
`First published 1988
`
`ISBN 0-443-03643-8
`
`British Library Cataloguing in Publication Data
`Pharmaceutics: the science of dosage form
`design.
`1. Pharmaceutics
`I. Aulton, Michael E.
`615'.19 (cid:9)
`RS403
`
`Library of Congress Cataloging in Publication Data
`Pharmaceutics: the science of dosage form design.
`Replaces: Cooper and Gunn's tutorial pharmacy.
`6th ed. 1972.
`Includes bibliographies and index.
`1. Drugs — Design of delivery systems. 2. Drugs
`— Dosage forms. 3. Biopharmaceutics.
`4. Pharmaceutical technology. 5. Chemistry,
`Pharmaceutical. 6. Microbiology, Pharmaceutical.
`I. Aulton, Michael E.
`[DNLM: 1. Biopharmaceutics. 2. Chemistry,
`Pharmaceutical. 3. Dosage Forms. 4. Technology,
`Pharmaceutical. 5. Microbiology, Pharmaceutical.
`QV 785 P5366]
``1'46--1.9rr
`
`615.5'8 86-25888
`
`Produced by Longman Group (FE) Ltd
`Printed in Hong Kong
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 2 of 72
`
`

`
`Preface
`
`The, first edition of Pharmaceutics has replaced the
`6th edition of Cooper and Gunn's Tutorial Phar-
`macy published by Pitman in 1972. Since then,
`there has been a change in editorship, a change
`in the title of the book, a change in some of the
`authors and a completely redesigned content. But
`all is not new and disjointed, the editorial link
`with Leicester School of Pharmacy continues.
`Sidney Carter recently retired as Deputy Head of
`Leicester School of Pharmacy and passed on the
`book to me. He in turn had inherited the book
`from one of its founders, the late Colin Gunn who
`was formerly Head of Leicester School of Phar-
`macy but sadly died on 25 February 1983.
`There are a greater number and a wider range
`of authors in this edition, each an accepted expert
`in the field on which they have written and, just
`as important, each has experience and ability in
`imparting that information to undergraduate phar-
`macy students.
`
`The philosophy of the subject matter which the
`book covers has changed because pharmaceutics
`has changed. Since the last edition of Tutorial
`Pharmacy there have been very marked changes
`in the concept and content of pharmaceutics.
`Those changes are reflected in this edition. The
`era of biopharmaceutics was in its infancy at the
`time of the previous edition. Since then we have
`become increasingly concerned with not merely
`producing elegant and accurate dosage forms but
`also ensuring that the optimum amount of drug
`reaches the required place in the body and stays
`there for the optimum amount of time, Now we
`are concerned much more with designing dosage
`forms and with all aspects of drug delivery. This
`book reflects that concern.
`
`Dr M E Aulton
`School of Pharmacy
`Leicester Polytechnic
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 3 of 72
`
`

`
`3 (cid:9)
`
`H Richards
`
`Solutions and their properties
`
`DEFINITION OF TERMS
`Methods of expressing concentration
`Quantity per quantity
`Percentage
`Parts
`Molarity
`Molality
`Mole fraction
`Milliequivalents and normal solutions
`
`TYPES OF SOLUTION
`Vapour pressures of solids, liquids and
`solutions
`Ideal solutions; Raoult's law
`Real or non-ideal solutions
`
`IONIZATION OF SOLUTES
`Hydrogen ion concentration and pH
`Dissociation (or ionization) constants and,pKa
`Buffer solutions and buffer capacity
`
`COLLIGATIVE PROPERTIES
`Osmotic pressure
`Osmolality and osmolarity
`I so-osmotic solutions
`Isotonic solutions
`
`DIFFUSION IN SOLUTION
`
`38
`
`The main aim of this chapter and Chapter 5 is to
`provide information on certain physicochemical
`principles that relate to the applications and
`implications of solutions in pharmacy. For the
`purposes of the present discussion these principles
`have . been classified in a somewhat arbitrary
`manner. Thus, this chapter deals mainly with the
`physicochemical properties of solutions that are
`important with respect to systems and processes
`described in other parts of this book or in the
`companion volume, Dispensing 'for. Pharmaceutical
`Students by Carter (1975) (new edition in prep-
`aration). Chapter 5 on the other hand is concerned
`with the principles underlying the formation of
`solutions and the factors that affect the rate and
`extent of this dissolution process. Because of
`limitations of space and the number of principles
`and properties that need to be considered the
`contents of each of these chapters should only be
`regarded as introductions to the various topics.
`The student is encouraged, therefore, to refer to
`the bibliography cited at the end of each chapter
`in order to augment the present contents. The
`textbook written by Florence and Attwood (1981)
`is recommended particularly, because of the large
`number of pharmaceutical examples that are used
`to aid understanding of physicochemical
`principles.
`
`DEFINITION OF TERMS
`
`A solution may be defined as a mixture of two or
`more components that form a single phase, which
`is homogeneous down to the molecular level. The
`component that determines the phase of the
`solution is termed the solvent and usually consti-
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 4 of 72
`
`

`
`tutes the largest proportion of the system. The
`other components are termed solutes and these are
`dispersed as molecules or ions throughout the
`solvent; i.e. they are said to be dissolved in the
`solvent. The extent to which the dissolution
`proceeds under a given set of experilnental
`conditions is referred to as the solubility of the
`solute in the solvent. Thus, the solubility of a
`substance is the amount of it that passes into
`solution when an equilibrium is established
`between the solution and excess, i.e. undissolved,
`substance. The solution that is obtained under
`these conditions is said to be saturated. Since the
`above definitions are general ones they may be
`applied to all types of solution. However, when
`the two components forming a solution are either
`both gases or both liquids then it is more usual
`to talk in terms of miscibility rather than
`solubility.
`
`Methods of expressing concentration
`Quantity per quantity
`
`Concentrations are often expressed simply as the
`weight or volume of solute that is contained in a
`given weight or volume of the solution. The
`majority of solutions encountered in pharmaceuti-
`cal practice consist of solids dissolved in liquids.
`Consequently, concentration is expressed most
`commonly by the weight of solute contained in a
`given volume of solution. Although the SI unit is
`kg rn-3 the terms that are used in practice are
`based on more convenient or appropriate weights
`and volumes. For example, in the case of a
`solution with a concentration of 1 kg m-3 the
`strength may be denoted by any one of the
`following concentration terms depending on the
`circumstances:
`1 g 1-1, 0.1 g per 100 ml, 1 mg ml-', 5 mg in 5 ml
`or 1 µg (cid:9)
`_
`
`Percentage
`
`The British and European Pharmacopoeias use the
`same method as a basis for their percentage
`expressions of the strengths of solutions. For
`example, the concentration of a solution of a solid
`in a liquid would be given by
`
`SOLUTIONS AND THEIR PROPERTIES 39
`concentration -= weight of solute
`in % w/v (cid:9)
`volume of solution X 100
`
`Per cent v/w, % v/v and % w/w expressions are
`also referred to in the General Notices of the British
`Pharmacopoeia (1980) together with the statement
`that the latter two expressions are used for
`solutions of liquids in liquids and solutions of
`gases in liquids, respectively.
`It should be realized that if concentration is
`expressed in terms of weight of solute in a given
`volume of solution then changes in volume caused
`by temperature fluctuations will alter the
`concentration.
`
`Parts
`
`The Pharmacopoeia also expresses some concen-
`trations in terms of the number of 'parts' of solute
`dissolved in a stated number of 'parts' of solution.
`Use of this method to describe the strength of a
`solution of a solid in a liquid infers that a given
`number of parts by volume (ml) of solution
`contain a certain number of parts by weight (g)
`of solid. In the case of solutions of liquids in
`liquids parts by volume of solute in parts by
`volume of solution are intended whereas with
`solutions of gases in liquids parts by weight of gas
`in parts by weight of solution are inferred.
`
`Molarity
`
`This is the number of moles of solute contained
`in 1 dm3 (or, more commonly in pharmacy, 1
`litre) of solution. Thus, solutions of equal molarity
`contain the same number of solute molecules in
`a given volume of solution. The unit of molarity
`is mol 1-1 (equivalent to 103 mol m-3 if converted
`to the SI unit). Although use of the term molar
`concentration and its symbol M to describe the
`molarity of a solution has been discouraged since
`the introduction of SI units the symbol M is still
`used in the current British and European
`Pharmacopoeias.
`
`Molality
`
`This is the number of moles of solute divided
`by the mass of the solvent, i.e. its SI unit is
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 5 of 72
`
`

`
`40
`
`PHYSICOCHEMICAL PRINCIPLES OF PHARMACEUTICS
`
`mol kg-1. Although it is less likely to be encoun-
`tered in pharmaceutical practice than the other
`terms it does offer a more precise description
`of concentration because it is unaffected by
`temperature.
`
`Mole fraction
`This is often used in theoretical considerations and
`is defined as the number of moles of solute divided
`by the total number of moles of solute and
`solvent, i.e.
`
`mole fraction of solute (x1) =
`
`n1
`n1 + n2
`
`where n1 and n2 are the numbers of moles of solute
`and solvent, respectively.
`
`Milliequivalents and normal solutions
`
`The concentrations of solutes in body fluids and
`in solutions used as replacements for these fluids
`are usually expressed in terms of the number of
`millimoles (1 millimole = one thousandth of a
`mole) in a litre of solution. In the case of electro-
`lytes, however, these concentrations may still be
`expressed in terms of milliequivalents per litre. A
`milliequivalent (mEq) of an ion is, in fact, one
`thousandth of the gram equivalent of the ion,
`which is, in turn, the ionic weight expressed in
`grams divided by the valency of the ion.
`Alternatively,
`
`1 mEq —
`
`ionic weight in mg
`valency
`
`A knowledge of the concept of chemical equiv-
`alents is also required in order to understand the
`use of 'normality' as a means of expressing the
`concentration of solutions, because a normal
`solution, i.e. concentration = 1 N, is one that
`contains the equivalent weight of the solute,
`expressed in grams, in 1 litre of solution. It was
`thought that this term would disappear on the
`introduction of SI units but it is still encountered
`in some volumetric assay procedures, e.g. in
`British Pharmacopoeias preceding the 1980 edition
`and in the current European Pharmacopoeia. The
`student is referred to Beckett and Stenlake (1975)
`for an explanation of chemical equivalents.
`
`TYPES OF SOLUTION
`
`Solutions may be classified on the basis of the
`physical states, i.e. gas, solid or liquid, of the
`solute(s) and solvent. Although a variety of
`different types can exist, solutions of pharmaceuti-
`cal interest virtually all possess liquid solvents. In
`addition, the solutes are predominantly solid
`substances. Consequently, most of the comments
`given in this chapter and in Chapter 5 are made
`with solutions of solids in liquids in mind.
`However, appropriate comments on other types,
`e.g. gases in liquids, liquids in liquids and solids
`in solids are included.
`
`Vapour pressures of solids, liquids and
`solutions
`An understanding of many of the properties of
`solutions requires an appreciation of the concept
`of an ideal solution and its use as a reference
`system, to which the behaviours of real (non-ideal)
`solutions can be compared. This concept is itself
`based on a consideration of vapour pressure. The
`present section is included, therefore, as an intro-
`duction to the later discussions on ideal and non-
`ideal solutions.
`The kinetic theory of matter indicates that the
`thermal motions of molecules of a substance in its
`gaseous state are more than adequate to overcome
`the attractive forces that exist between the mol-
`ecules, so that the molecules undergo a completely
`random movement within the confines of the
`container. The situation is reversed, however,
`when the temperature is lowered sufficiently so
`that a condensed phase is formed. Thus, the
`thermal motions of the molecules are now insuf-
`ficient to overcome completely the intermolecular
`attractive forces and some degree of order in the
`relative arrangement of molecules occurs. If the
`intermolecular forces are so strong that a high
`degree of order, which is hardly influenced by
`thermal motions, is brought about then the
`substance is usually in the solid state.
`In the liquid condensed state the relative influ-
`ences of thermal motion and intermolecular
`attractive forces are intermediate between those in
`the gaseous and solid states. Thus, the effects of
`interactions between the permanent and induced
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 6 of 72
`
`

`
`r
`
`dipoles, i.e. the so-called van der Waals forces of
`attraction, lead to some degree of coherence
`between the molecules of liquids. Consequently,
`liquids occupy a definite volume, unlike gases,
`and whilst there is evidence of structure within
`liquids such structure is much less apparent than
`in solids.
`Although solids and liquids are condensed
`systems with cohering molecules some of the
`surface molecules in these systems will occasion-
`ally acquire sufficient energy to overcome the
`attractive forces exerted by adjacent molecules and
`so escape from the surface to form a vaporous
`phase. If temperature is maintained constant an
`equilibrium will be established eventually between
`the vaporous and condensed phases and the
`pressure exerted by the vapour at equilibrium is
`referred to as the vapour pressure of the
`substance.
`All condensed systems have the inherent ability
`to give rise to a vapour pressure. However, the
`vapour pressures exerted by solids are usually
`much lower than those exerted by liquids, because
`the intermolecular forces in solids are stronger
`than those in liquids so that the escaping tendency
`for surface molecules is higher in liquids.
`Consequently, surface loss of vapour from liquids
`by the process of evaporation is more common
`than surface loss of vapour from solids via
`sublimation.
`In the case of a liquid solvent containing a
`dissolved solute then molecules of both solvent
`and solute may show a tendency to escape from
`the surface and so contribute to the vapour
`pressure. The relative tendencies to escape will
`depend not only on the relative numbers of the
`different molecules in the surface of the solution
`but also on the relative strengths of the attractive
`forces between adjacent solvent molecules on the
`one hand and between solute and solvent mol-
`ecules on the other hand. Thus, since the inter-
`molecular forces between solid solutes and liquid
`solvents tend to be relatively strong such solute
`molecules do not generally escape from the surface
`of a solution and contribute to the vapour
`pressure. In other words the solute is non-volatile
`and the vapour pressure arises solely from the
`dynamic equilibrium that is set up between the
`rates of evaporation and condensation of solvent
`
`SOLUTIONS AND THEIR PROPERTIES 41
`
`molecules contained in the solution. In a mixture
`of miscible liquids, i.e. a liquid in liquid solution,
`the molecules of both components are likely to
`evaporate and contribute to the overall vapour
`pressure exerted by the solution.
`
`Ideal solutions; Raoult's Law
`
`The concept of an ideal solution has been intro-
`duced in order to provide a model system that can
`be used as a standard, to which real or non-ideal
`solutions can be compared. In the model it is
`assumed that the strengths of all intermolecular
`forces are identical, i.e. solvent—solvent,
`solute—solvent and solute—solute interactions are
`the same and are equal, in fact, to the strength of
`the intermolecular interactions in either the pure
`solvent or pure solute. Because of this equality the
`relative tendencies of solute and solvent molecules
`to escape from the surface of the solution will be
`determined only by their relative numbers in the
`surface. Since a solution is homogeneous by
`definition then the relative numbers of these
`surface molecules will be reflected by the relative
`numbers in the whole of the solution. The latter
`can be expressed conveniently by the mole frac-
`tions of the components because, for a binary
`solution, i.e. one with two components, x1 + x2
`= 1, where xi and x2 are the mole fractions of the
`solute and solvent, respectively. Thus, the total
`vapour pressure (p) exerted by such a binary
`solution is given by Eqn 3.1:
`
`P (cid:9)
`Pi + P2 = PYX1 (cid:9)
`(3.1)
`P2X2 (cid:9)
`where pi and p2 are the partial vapour pressures
`exerted above the solution by solute and solvent,
`respectively, and pi and p`l are the vapour press-
`ures exerted by pure solute and pure solvent,
`respectively.
`If the total vapour pressure of the solution is
`described by Eqn 3.1 it follows that Raoult's law
`is obeyed by the system because this law states
`that the partial vapour pressure exerted by a
`volatile component in a solution at a given
`temperature is equal to the vapour pressure of the
`pure component at the same temperature, multi-
`plied by its mole fraction in the solution, i.e.
`
`Pi = P7x1
`
`(3.2)
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 7 of 72
`
`(cid:9)
`

`
`42 PHYSICOCHEMICAL PRINCIPLES OF PHARMACEUTICS
`
`On'e of the consequences of the preceding
`comments is that an ideal solution may be defined
`as one which obeys Raoult's law. In addition, ideal
`behaviour should only be expected to be exhibited
`by real systems comprised of chemically similar
`components, because it is only in such systems
`that the condition of equal intermolecular forces
`between components, that is assumed in the ideal
`model, is likely to be satisfied. Consequently,
`Raoult's law is obeyed over an appreciable concen-
`tration range by relatively few systems in reality.
`Mixtures of benzene + toluene, n-hexane + n-
`heptane and ethyl bromide + ethyl iodide are
`commonly mentioned systems that exhibit ideal
`behaviour, whilst a more pharmaceutically inter-
`esting example is provided by binary mixtures of
`fluorinated hydrocarbons. These latter mixtures
`are used as propellants in therapeutic aerosols and
`their approximation to ideal behaviour allows Eqn
`3.1 to be used to calculate the total pressure
`exerted by a given mixture.
`
`Real or non-ideal solutions
`
`The majority of real solutions do not exhibit ideal
`behaviour because solute—solute, solute—solvent
`and solvent—solvent forces of interaction are
`unequal. These inequalities alter the effective
`concentration of each component so that it cannot
`be represented by a normal expression of concen-
`tration, such as the mole fraction term x that is
`used in Eqns 3.1 and 3.2. Consequently, devi-
`ations from Raoult's law are often exhibited by
`real solutions and the previous equations are not
`obeyed in such cases. The equations can be modi-
`fied, however, by substituting each concentration
`term (x) by a measure of the effective concen-
`tration, which is provided by the so-called activity
`(or thermodynamic activity), a. Thus, Eqn 3.2 is
`converted into Eqn 3.3,
`
`(3.3)
`
`Pi = (cid:9)
`which is applicable to all systems whether they be
`ideal or non-ideal. It should be noted that if a
`solution exhibits ideal behaviour then a = x,
`whereas a (cid:9)
`x if deviations from such behaviour
`are apparent. The ratio of activity/concentration
`is termed the activity coefficient (f) and it provides
`a measure of the deviation from ideality. (The
`
`student is encouraged to study relevant parts of
`the bibliography for further information on
`thermodynamic terms such as activity, activity
`coefficient, free energy and chemical potential.)
`If the attractive forces between solute and
`solvent molecules are weaker than those exerted
`between the solute molecules themselves or the
`solvent molecules themselves then the components
`will have little affinity for each other. The
`escaping tendency of the surface molecules in such
`a system is increased when compared with an ideal
`solution. In other words pi , p2 and p are greater
`than expected from Raoult's law and the thermo-
`dynamic activities of the components are greater
`than their mole fractions, i.e. al > x1 and a2 > x2.
`This type of system is said to show a positive
`deviation from Raoult's law and the extent of the
`deviation increases as the miscibility of the
`components decreases. For example, a mixture of
`alcohol and benzene shows a smaller deviation
`than the less miscible mixture of water + diethyl
`ether whilst the virtually immiscible mixture of
`benzene + water exhibits a very large positive
`deviation.
`Conversely, if the solute and solvent have a
`strong mutual affinity that results in the formation
`of a complex or compound then a negative devi-
`ation from Raoult's law occurs. Thus, pi, p2 and
`p are lower than expected and a l < x1 and
`a2 < x2. Examples of systems that show this type
`of behaviour include chloroform + acetone,
`pyridine + acetic acid and water + nitric acid.
`Even though most systems are non-ideal and
`deviate either positively or negatively from
`Raoult's law, such deviations are small when a
`solution is dilute because the effects of a small
`amount of solute on interactions between solvent
`molelcules are minimal. Thus, dilute solutions
`tend to exhibit ideal behaviour and the activities
`of their components approximate to their mole
`fractions, i.e. a l (cid:9)
`x1 and a2 = x2. Conversely,
`large deviations may be observed when the
`concentration of a solution is high. Knowledge of
`the consequences of such marked deviations is
`particularly important in relation to the distillation
`of liquid mixtures. For example, the complete
`separation of the components of a mixture by frac-
`tional distillation may not be achievable if large
`positive or negative deviations from Raoult's law
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 8 of 72
`
`

`
`give rise to the formation of so-called azeotropic
`mixtures with minimum and maximum boiling
`points, respectively. Such knowledge is obviously
`important to the pharmaceutical chemist but is
`beyond the scope of the present chapter.
`
`IONIZATION OF SOLUTES
`
`Many solutes dissociate into .ions if the dielectric
`constant of the solvent is high enough to cause
`sufficient separation of the attractive forces
`between the oppositely charged ions. Such solutes
`are termed electrolytes and their ionization (or
`dissociation) has several consequences that are
`often important in pharmaceutical practice. Some
`of these consequences are indicated below whilst
`others that relate to solubilities and dissolution
`rates are referred to in Chapter 5.
`
`Hydrogen ion concentration and pH
`
`The dissociation of water can be represented by
`Eqn 3.4:
`
`H2O (cid:9) H + OH- (cid:9)
`
`(3.4)
`
`although it should be realized that this is a
`simplified representation because the hydrogen
`and hydroxyl ions do not exist in a free state but
`combine with undissociated water molecules to
`yield more complex ions such as H30+ and H7M
`In pure water the concentrations of H+ and
`OH- ions are equal and at 25 °C both have the
`values of 1 x 10-7 mol 1-1 . Since the
`Lowry—Bronsted theory of acids and bases defines
`an acid as a substance which donates a proton (or
`hydrogen ion) it follows that the addition of an
`acidic solute to water will result in a hydrogen ion
`concentration that exceeds this value. Conversely,
`the addition of a base, which is defined as a
`substance that accepts protons, will decrease the
`concentration of hydrogen ions.
`The hydrogen ion concentration range that can
`be obtained decreases from 1 mol 1-1 for a strong
`acid down to 1 x 10-14 mol 1-1 for a strong base.
`In order to avoid the frequent use of low values
`that arise from this range the concept of pH has
`been introduced as a more convenient measure of
`hydrogen ion concentration. pH is defined as the
`
`SOLUTIONS AND THEIR PROPERTIES 43
`
`negative logarithm of the hydrogen ion concen-
`tration [H-E] as shown by Eqn 3.5:
`
`pH = —logio[H+] (cid:9)
`
`(3.5)
`
`so that the pH of a neutral solution like pure water
`is 7, because the concentration of H+ ions (and
`OH-) ions. = 1 x 10-7 mol 1-1, and the pHs of
`acidic and alkaline solutions will be les's or greater
`than 7, respectively.
`pH has several important implications in phar-
`maceutical practice. For example, in addition to
`its effects on the solubilities of drugs that are weak
`acids or bases, as indicated in Chapter 5, pH may
`have a considerable effect on the stabilities of
`many drugs, be injurious to body tissues and affect
`the ease of absorption of drugs from the gastro-
`intestinal tract into the blood (see Chapter 9).
`
`Dissociation (or ionization) constants and pKa
`
`Many drugs may be classified as weak acids or
`weak bases which means that in solutions of these
`drugs equilibria exist between undissociated
`molecules and their ions. Thus, in a solution of
`a weakly acidic drug HA the equilibrium may be
`represented by Eqn 3.6:
`
`HA -,='• F11- + A-
`
`(3.6)
`
`although the proton H+ would be better
`represented by H30+ because it is always strongly
`solvated by a water molecule. Similarly, the
`protonation of a weakly basic drug B can be
`represented by Eqn 3.7:
`
`B + H+ ± BH+ (cid:9)
`
`(3.7)
`
`Such equilibria are unlikely to occur in solutions
`of most salts of strong acids or bases in water
`because these compounds are completely ionized.
`The ionization constant (or dissociation constant)
`Ka of a weak acid can be obtained by applying the
`Law of Mass Action to Eqn 3.6 to yield Eqn 3.8:
`
`Ka (cid:9)
`
`= (cid:9)
`
`[1-1-1[A-]
`[HA] (cid:9)
`
`(3.8)
`
`Taking logarithms of both sides of Eqn 3.8 yields
`
`log Ka = log [H+] + log [Al — log [HA]
`
`and the signs in this equation may be reversed to
`give Eqn 3.9:
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 9 of 72
`
`(cid:9)
`(cid:9)
`(cid:9)
`(cid:9)
`

`
`44 PHYSICOCHEMICAL PRINCIPLES OF PHARMACEUTICS
`
`-log Ka = -log [H+] - log [A- ] + log [HA]
`(3.9)
`The symbol pKa is used to represent the negative
`logarithm of the acid dissociation constant Ka in
`the same way that pH is used to represent the
`negative logarithm of the hydrogen ion concen-
`tration [H+] and Eqn 3.9 may therefore be
`rewritten as Eqn 3.10:
`pKa = pH + log [HA] - log [A ] (3.10)
`
`or
`
`(3.11)
`
`pKa = pH + log [HA]
`[A-]
`Thus, a general equation, Eqn 3.12, that is appli-
`cable to any acidic drug with one ionizable group
`may be written, where cu and ci represent the
`concentrations of the unionized and ionized
`species, respectively. This equation is known as
`the Henderson-Hasselbalch equation.
`
`pKa = pH + log —ccu (cid:9)
`
`(3.12)
`
`From Eqn 3.7 it can be seen that the acid
`dissociation constant (Ka) of a protonated weak
`base is given by Eqn 3.13:
`
`Ka -
`
`[111[B]
`[BH+]
`Taking negative logarithms yields Eqn 3.14:
`-log Ka = -log [Hr] - log [B] + log [Bill
`(3.14)
`
`(3.13)
`
`or
`
`pKa = pH + log (cid:9)
`
`
`
`[B]
`The Henderson-Hasselbalch equation for any
`weak base with one ionizable group may therefore
`be written as shown by Eqn 3.15:
`
`pKa = pH + log —c (cid:9)
`i (cid:9)
`Cu
`
`(3.15
`)
`
`temperature at which the determination is
`performed should be specified because the values
`of the constants vary with temperature.
`Ionization constants are usually expressed in
`terms of pKa values for both acidic and basic drugs
`and a list of pKa values for a series of important
`drugs is given in the Pharmaceutical Handbook
`(1980).
`The degree of ionization of a drug in a solution
`can be calculated from the Henderson-Hasselbalch
`equations for weak acids and bases (Eqns 3.12 and
`3.15, respectively) if the pKa value of the drug and
`the pH of the solution are known. Such calcu-
`lations are particularly useful in determining the
`degree of ionization of drugs in various parts of
`the gastrointestinal tract and in the plasma (see
`Chapter 9). The following examples are therefore
`related to this type of situation.
`1 The pKa value of aspirin, which is a weak acid,
`is about 3.5, and if the pH of the gastric
`contents is 2.0 then from Eqn 3.12
`
`log —cu = pKa - pH = 3.5 - 2.0 = 1.5
`c,
`
`so that the ratio of the concentration of un-
`ionized acetylsalicyclic acid to acetylsalicylate
`anion is given by
`cu:ci = antilog 1.5 = 31.62:1
`2 The pH of plasma is 7.4 so that the ratio of
`unionized:ionized aspirin in this medium is
`given by
`
`log
`
`CL1
`T, (cid:9)
`
`pKa - pH = 3.5 - 7.4 = -3.9
`
`and
`cu:ci = antilog -3.9 = antilog 4.1
`= 1.259 x 10 -4:1
`3 The pKa of the weakly acidic drug sulphapyri-
`dine is about 8.0 and if the pH of the intestinal
`contents is 5.0 then the ratio of unionized:ionized
`drug is given by
`
`where ci and cu refer to the concentrations of the
`protonated and unionized species, respectively.
`Various analytical techniques, e.g. spectro-
`photometric and potentiometric methods, may be
`used to determine ionization constants but the
`
`log °
`c;
`
`and
`
`= pKa - pH = 8.0 - 5.0 = 3.0
`
`cu:ci = antilog 3.0 = 103:1
`
`NOVARTIS EXHIBIT 2032
`Par v Novartis, IPR 2016-00084
`Page 10 of 72
`
`(cid:9)
`(cid:9)
`(cid:9)
`

`
`Y.
`
`4 The pKa of the basic drug amidopyrine is 5.0,
`and in the stomach the ratio of ionized:unionized
`drug is shown from Eqn 3.15 to be given by
`ci
`log-- = pKa — pH = 5.0 — 2.0 = 3.0
`Cu
`
`and
`
`ci:cu = antilog 3.0 = 103:1
`while in the intestine the ratio is given by
`
`log— =- 5.0 — 5.0 0
`cu
`
`and
`
`ci:cu = antilog 0 = 1:1
`
`Buffer solutions and buffer capacity
`
`These solutions will maintain a constant pH even
`when small amounts of acid or alkali are added to
`the solution. They usually contain mixtures of a
`weak acid and its salt (i.e. its conjugate base)
`although mixtures of weak bases and their salts
`(i.e. their conjugate acids) may be used but suffer
`from the disadvantage that arises from the
`volatility of many of the bases.
`The action of a buffer solution can be appreci-
`ated by considering a simple system such as a
`solution of acetic acid and sodium acetate in
`water. The acetic acid, being a weak acid, will be
`confined virtually to its undissociated form
`because its ionization will be suppressed by the
`presence of common acetate ions produced by
`complete dissociation of the sodium salt. The p1-1
`of this solution can be described by Eqn 3.16,
`which is a rearranged form of Eqn 3.12:
`
`pH = pKa + log L. (cid:9)
`Cu
`
`(3.16)
`
`It can be seen from Eqn 3.16 that the pH will
`remain constant as long as the logarithm of the
`ratio ci/cu does not change. When a small amount
`of acid is added to the solution it will convert
`some of the salt into acetic acid but if the concen-
`trations of both acetate ion and acetic acid are
`reasonably large then the effect of the change will
`be negligible and the pH will remain constant.
`Similarly, the addition of a small amount of base
`will convert some of the acetic acid into its salt
`
`SOLUTIONS AND THEIR PROPERTIES 45
`
`form but the pH will be unaltered if the overall
`changes in concentrations of the two species are
`relatively small.
`If large amounts of acid or base are added to a
`buffer then changes in the log ci/cu term become
`appreciable and the pH alters. The ability of a
`buffer to withstand the effects of acids and bases
`is an im

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