throbber
Pharmaceutical
`Dissolution Testing
`
`
`
`© 2005 by Taylor & Francis Group, LLC© 2005 by Taylor & Francis Group, LLC
`
`MYLAN EXHIBIT 1028
`
`

`

`Pharmaceutical
`Dissolution Testing
`
`Edited by
`
`Jennifer Dressman
`Johann Wolfang Goethe University
`Frankfurt, Germany
`
`Johannes Krämer
`Phast GmbH
`Homburg/Saar, Germany
`
`
`
`© 2005 by Taylor & Francis Group, LLC© 2005 by Taylor & Francis Group, LLC
`
`

`

`Published in 2005 by
`Taylor & Francis Group
`6000 Broken Sound Parkway NW, Suite 300
`Boca Raton, FL 33487-2742
`
`© 2005 by Taylor & Francis Group, LLC
`
`No claim to original U.S. Government works
`Printed in the United States of America on acid-free paper
`10 9 8 7 6 5 4 3 2 1
`
`International Standard Book Number-10: 0-8247-5467-0 (Hardcover)
`International Standard Book Number-13: 978-0-8247-5467-9 (Hardcover)
`
`This book contains information obtained from authentic and highly regarded sources. Reprinted material is
`quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable efforts
`have been made to publish reliable data and information, but the author and the publisher cannot assume
`responsibility for the validity of all materials or for the consequences of their use.
`
`No part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic,
`mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and
`recording, or in any information storage or retrieval system, without written permission from the publishers.
`
`For permission to photocopy or use material electronically from this work, please access www.copyright.com
`(http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC) 222 Rosewood Drive,
`Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration
`for a variety of users. For organizations that have been granted a photocopy license by the CCC, a separate
`system of payment has been arranged.
`
`Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only
`for identification and explanation without intent to infringe.
`
`Library of Congress Cataloging-in-Publication Data
`
`Catalog record is available from the Library of Congress
`
`Taylor & Francis Group
`is the Academic Division of T&F Informa plc.
`
`Visit the Taylor & Francis Web site at
`http://www.taylorandfrancis.com
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`6
`
`Physiological Parameters Relevant to
`Dissolution Testing: Hydrodynamic
`Considerations
`
`STEFFEN M. DIEBOLD
`
`Leitstelle Arzneimittelu¨ berwachung Baden–
`Wu¨ rttemberg, Regierungspra¨sidium Tu¨ bingen,
`Tu¨ bingen, Germany
`
`HYDRODYNAMICS AND DISSOLUTION
`
`Dissolution
`
`Why Is Hydrodynamics Relevant to Dissolution
`Testing?
`
`Release-related bioavailability problems have been encoun-
`tered in the pharmaceutical development of formulations for
`a number of quite different chemical entities, including ciclos-
`porin, digoxin, griseofulvin, and itraconazole, to name but a
`few. A thorough knowledge of hydrodynamics is useful in
`
`127
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`128
`
`Diebold
`
`the course of dissolution method development and formula-
`tion development, as well as for the pharmaceutical industry’s
`quality needs, e.g., batch-to-batch control. Occasionally, qual-
`ity control specifications are not met due to ‘‘minor’’ variations
`involving hydrodynamics, such as the use of different
`volumes, or modified stirring devices or sampling procedures.
`The development of drug formulations is facilitated by the
`choice of an appropriate dissolution apparatus based on
`insight into its specific hydrodynamic performance. Using
`the right test might make it easier, for instance, to isolate
`the impact of different excipients and process parameters on
`drug release at an early stage of pharmaceutical formulation
`development. Furthermore, a sound knowledge of in vivo
`hydrodynamics may help to better understand and possibly
`to improve forecasting of in vivo dissolution and absorption
`of
`biopharmaceutical
`classification
`system (BCS)
`II
`compounds. Although gastrointestinal (GI) fluids are well-
`characterized and biorelevant dissolution media [e.g., Fasted
`State Simulated Intestinal Fluid (FaSSIF) and Fed State
`Simulated Intestinal Fluid (FeSSIF)] have been developed
`to simulate various states in the GI tract, knowledge of hydro-
`dynamics appears to be relatively scant both in vitro and in
`vivo. This chapter gives a brief introduction of the basic
`hydrodynamics relevant to in vitro dissolution testing, includ-
`ing the convective diffusion theory. This section is followed by
`hydrodynamic considerations of in vitro dissolution testing
`and hydrodynamic problems inherent to in vivo bioavailabil-
`ity of solid oral dosage forms.
`
`The Dissolution Process
`
`Dissolution can be described as a mass transfer process.
`Although mass transfer processes commonly are under the
`combined influence of both thermodynamics and hydrody-
`namics, usually one of these prevails in terms of the overall
`dissolution process (1–3). Hydrodynamics is predominant for
`the overall dissolution rate if the mass transfer process is
`mainly controlled by convection and/or diffusion, as is usually
`the case for poorly soluble substances. This is of great
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`129
`
`practical relevance for pharmaceutical development, since
`new drug compounds often exhibit poor solubility in aqueous
`media.
`
`The Dissolution Rate
`
`The dissolution rate (dC/dt) of a pure drug compound is repre-
`sented by an equation based on the work of Noyes, Whitney,
`Nernst, and Brunner (4–6), which is in turn based on earlier
`observations made by Schu¨ karew in 1891 (7):
`
`dC
`dt
`

`
`A D
`dHL V
`The proportionality constant k
`
`ðCs CtÞ
`
`k ¼
`
`A D
`dHL V
`is addressed as the ‘‘apparent dissolution rate constant.’’ Cs
`represents the saturation solubility, Ct describes the bulk con-
`centration of the dissolved drug at time t, D is the effective
`diffusion coefficient of the drug molecule, A stands for the sur-
`face area available for dissolution, and V represents the
`media volume employed in the test. According to the equa-
`tions of Noyes, Whitney, Nernst, and Brunner, the dissolution
`rate depends on a small fluid ‘‘layer,’’ called the hydrodynamic
`boundary layer (dHL), adhering closely to the surface of a solid
`particle that is to be dissolved (solvendum, solute). As can be
`seen from the combined equation, an inverse proportionality
`exists between the dissolution rate and the hydrodynamic
`boundary layer. If the latter is reduced, the dissolution rate
`increases.
`
`Hydrodynamic Basics Relevant to Dissolution
`
`Laminar and Turbulent Flow
`
`Laminar flow is characterized by layers (‘‘lamellae’’) of liquid
`moving at the same speed and in the same direction (Fig. 1).
`Little or no exchange of fluid mass and fluid particles occurs
`across these fluid layers. The closer the layers are to a given
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`130
`
`Diebold
`
`Figure 1 (A) Laminar and (B) turbulent flow: t describes the time
`scale, UA represents the velocity component acting in the direction
`of the flow. Source: From Ref. 10.
`
`surface, the slower they move. In an ideal fluid, the flow
`follows a curved surface smoothly, with the layers central in
`the flow moving fastest and those at the sides slowest. In tur-
`bulent flow, by contrast, the streamlines or flow patterns are
`disorganized and there is an exchange of fluid between these
`areas. Momentum is also exchanged such that slow-moving
`fluid particles speed up and fast-moving fluid particles give
`up their momentum to the slower-moving particles and slow
`down themselves. All, or nearly all, fluid flow displays some
`degree of turbulence. If the fluid velocity exceeds a crucial
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`131
`
`number, flow becomes turbulent rather than laminar since
`the frictional force can no longer compensate for other forces
`acting on the fluid particles. This event depends on the fluid
`viscosity, the fluid velocity, and the geometry of the hydrody-
`namic system and is described by the Reynolds number.
`
`Reynolds Number
`
`The dimensionless Reynolds number (Re) is used to character-
`ize the laminar–turbulent
`transition and is commonly
`described as the ratio of momentum forces to viscous forces
`in a moving fluid. It can be written in the form
`
`Re ¼
`
`r UA L
`Z
`

`
`UA L
`
`n
`
`where n represents the kinematic viscosity of the liquid (r and
`Z are the density and dynamic viscosity, respectively). UA
`describes the flow rate, and L represents a characteristic dis-
`tance or length of the hydrodynamic system, for example, the
`diameter of a tube or pipe. Laminar flow patterns turn into
`turbulent flow if the Reynolds number of a particular hydro-
`dynamic system exceeds a critical Reynolds number (Recrit).
`
`Particle–Liquid Reynolds Numbers
`
`With respect to the hydrodynamics of particles in a stirred
`dissolution medium, the Reynolds numbers determined for
`the bulk flow have to be distinguished from the Reynolds
`numbers characterizing the particle–liquid system. The latter
`hydrodynamic subsystem consists of the dissolving particles
`and the surrounding fluid close to their surfaces. Thus, it is
`the relative velocity of the solid particle surface to the bulk
`flow (the ‘‘slip velocity’’) that counts. However, it is permissi-
`ble to approximate the slip velocity to UA, provided that the
`drug particles are suspended in the moving fluid and the
`density difference between particle and dissolution medium
`is at least 0.3 g/cm3 (8). In this case, L represents a charac-
`teristic length on the (average) particle surface and may arbi-
`trarily be identified with the particle diameter. With respect
`to particle–liquid systems, the laminar–turbulent transition
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`132
`
`Diebold
`
`at the particle surface is decisive. Laminar flow turns turbu-
`lent if Recrit for the flow close to the particle surface is
`exceeded. Thus, Recrit (particle) is not necessarily identical
`with the Reynolds number of the bulk flow—although the lat-
`ter may sometimes serve as a sufficient approximation (9,10).
`
`‘‘Eddies,’’ Dissipation, and Energy Cascade
`
`‘‘Eddies’’ are turbulent instabilities within a flow region
`(Fig. 2). These vortices might already be present in a turbu-
`lent stream or can be generated downstream by an object pre-
`senting an obstacle to the flow. The latter turbulence is
`known as ‘‘Karman vortex streets.’’ Eddies can contribute a
`considerable increase of mass transfer in the dissolution
`process under turbulent conditions and may occur in the GI
`tract as a result of short bursts of intense propagated motor
`activity and flow ‘‘gushes.’’
`The mean velocity of eddies changes at a definitive
`distance called the ‘‘scale of motion’’ (SOM) (11). The bigger
`these eddies are, the longer is the SOM [(9), Sec. 4]. Apart
`from ‘‘large scale eddies,’’ a number of ‘‘small scale eddies’’
`
`Figure 2 ‘‘Eddies’’ (large scale type) downstream of an object
`exposed to flow. Source: Adapted from Ref. 13, Sec. 21.4 (original
`by Grant HL. J Fluid Mech 1958; 4:149).
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`133
`
`exist in turbulent flow. Under turbulent conditions, eddies
`transport the majority of the kinetic energy. Energy fed into
`the turbulence goes primarily into the larger eddies. From
`these, smaller eddies are generated, and then still smaller
`ones. The process continues until the length scale is small
`enough for viscous action to be important and dissipation to
`occur. This sequence is called the energy cascade. At high
`Reynolds numbers the cascade is long; i.e., there is a large dif-
`ference in the eddy sizes at its ends. There is then little direct
`interaction between the large eddies governing the energy
`transfer and the small, dissipating eddies. In such cases,
`the dissipation is determined by the rate of supply of energy
`to the cascade by the large eddies and is independent of the
`dynamics of the small eddies in which the dissipation actually
`occurs. The rate of dissipation is independent of the magni-
`tude of the viscosity. An increase in Reynolds number to a still
`higher value extends the cascade only at the small eddy end.
`Still, smaller eddies must be generated before dissipation can
`occur.
`
`Energy Input e
`
`For closed dissolution systems, it can be hypothesized that the
`hydrodynamics depends on the input of energy in a general
`way. The energy input may be characterized by the power
`input per unit mass of fluid or the turbulent energy dissipa-
`tion rate per unit mass of fluid (e). Considering various paddle
`apparatus, the power input per unit mass of fluid (Fig. 3) can
`be calculated according to Plummer and Wigley [(12), Appen-
`dix B, nomenclature adapted]:
`
`e ¼
`
`p I5 o3
`V
`where e has the dimension length2/time3. o stands for the
`rotations per minute, I is the mean diameter of the paddle
`or impeller, p is a model constant dependent on the hydrody-
`namic flow pattern (laminar or turbulent), and V is the fluid
`volume. As expected, e is influenced mainly by the diameter
`of the impeller and the rotation rate. Based on this equation,
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`134
`
`Diebold
`
`Figure 3 Power input per unit mass of fluid: paddle apparatus,
`900 mL. Calculations shown for extremes of completely laminar
`and completely turbulent hydrodynamic conditions. The actual
`energy input lies in between the two curves, depending on the
`stirring rate. Source: From Ref. 10.
`
`the power input per unit mass of fluid for the compendial pad-
`dle apparatus has been calculated [(10), Chapter 5.6.2]. The
`fluid mass specific energy input rises exponentially with pad-
`dle speed. The exponential form of the observed relationship
`suggests that there is a transition from laminar (p ¼ 0.5) to
`turbulent flow (p ¼ 1.0) within the system, and indicates that
`the energy input to the media and flow pattern in the vessels
`are related.
`The power input per unit mass of fluid is greater for a
`dissolution volume of 500 mL than for 900 mL, at a given stir-
`ring rate. Remarkably, e calculated for laminar conditions
`(p ¼ 0.5) employing 500 mL of dissolution medium (not
`plotted) results in approximately the same hydrodynamic
`effectiveness as when turbulent conditions are assumed
`(p ¼ 1.0) for a dissolution volume of 900 mL (10). This implies
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`135
`
`more effective hydrodynamics for the lower volume. Thus, it
`cannot be assumed that there are no hydrodynamic implica-
`tions when volumes used for a specific dissolution test method
`are changed, but rather, that the change would require
`meticulous validation!
`
`Hydrodynamic Boundary Layer Concept
`
`Concept and Structure of the Boundary Layer
`
`A boundary layer in fluid mechanics is defined as the layer of
`fluid in the immediate vicinity of a limiting surface where the
`layer and its breadth are affected by the viscosity of the fluid.
`The concept of the hydrodynamic boundary layer goes back to
`the work of the German physicist and mathematician Ludwig
`Prandtl (1875–1953) and was first presented at Go¨ttingen and
`Heidelberg in 1904 (Fig. 4). According to the Prandtl concept,
`at high Reynolds numbers, the flow close to the surface of a
`body can be separated into two main regions. Within the bulk
`flow region viscosity is negligible (‘‘frictionless flow’’), whereas
`near the surface a small region exists that is called the
`
`Figure 4 Hydrodynamic boundary layer development on the
`semi-infinite plate of Prandtl. dD ¼ laminar boundary layer,
`dT ¼ turbulent boundary layer, dVS ¼ viscous turbulent sub-layer,
`dDS ¼ diffusive sub-layer (no eddies are present; solute diffusion
`and mass transfer are controlled by molecular diffusion—the thick-
`ness is about 1/10 of dVS), B ¼ point of laminar–turbulent transition.
`Source: From Ref. 10.
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`136
`
`Diebold
`
`hydrodynamic boundary layer. In this region, adherence of
`molecules of the liquid to the surface of the solid body slows
`them down. The hydrodynamic boundary layer is dominated
`by pronounced velocity gradients within the fluid that are
`continuous, and does not, as is sometimes purported, consist
`of a ‘‘stagnant’’ layer. According to Newton’s law of friction,
`pronounced velocity gradients lead to high friction forces near
`the surface of a solid particle. The hydrodynamic boundary
`layer grows further downstream of the surface since more
`and more fluid molecules are slowed down.
`In terms of hydrodynamics, the boundary layer thickness
`is measured from the solid surface (in the direction perpendi-
`cular to a particle’s surface, for instance) to an arbitrarily cho-
`sen point, e.g., where the velocity is 90–99% of the stream
`velocity or the bulk flow (d90 or d99, respectively). Thus, the
`breadth of the boundary layer depends ad definitionem on
`the selection of the reference point and includes the laminar
`boundary layer as well as possibly a portion of a turbulent
`boundary layer.
`
`Laminar and Turbulent Boundary Layer
`
`Apart from the nature of the bulk flow, the hydrodynamic sce-
`nario close to the surfaces of drug particles has to be consid-
`ered. The nature of
`the hydrodynamic boundary layer
`generated at a particle’s surface may be laminar or turbulent
`regardless of the bulk flow characteristics. The turbulent
`boundary layer is considered to be thicker than the laminar
`layer. Nevertheless, mass
`transfer
`rates are usually
`increased with turbulence due to the presence of the ‘‘viscous
`turbulent sub-layer.’’ This is the part of the (total) turbulent
`boundary layer that constitutes the main resistance to the
`overall mass transfer in the case of turbulence. The develop-
`ment of a viscous turbulent sub-layer reduces the overall
`resistance to mass transfer since this viscous sub-layer is
`much narrower than the (total) laminar boundary layer.
`Thus, mass transfer from turbulent boundary layers is
`greater than would be calculated according to the total
`boundary layer thickness.
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`137
`
`Boundary Layer Separation
`
`Both laminar and turbulent boundary layers can separate.
`Laminar layers usually require only a relatively short region
`of adverse pressure gradient to produce separation, whereas
`turbulent layers separate less readily. A few examples of
`turbulent boundary layer separation include golf ball design
`to stabilize trajectory, airfoil design to reduce aerodynamic
`resistance (Fig. 5), and, in nature, in sharkskin to improve
`the shark’s ability to glide. The overall flow pattern, when
`separation occurs, depends greatly on the particular flow.
`The flow upstream of the separation point is fed by recircula-
`tion of some of the separating fluid. Sometimes the effect is
`quite localized, but more often it is not. The consequent
`post-separation pattern is affected by the fact that the sepa-
`rated flow becomes turbulent and so there is a highly fluctu-
`ating recirculation motion over the whole surface of the
`body. With respect to the dissolution of drug particles from
`oral solid formulations, recirculation flow is expected to
`increase mass transfer and can take place even at a low
`Reynolds numbers of Re  10 (13).
`As mentioned, a laminar boundary layer separates a
`greater distance from the surface of a curved body than a
`turbulent one. The laminar boundary layer in the upper
`photograph of Figure 5 is shown separating from the crest
`
`Figure 5 Boundary layer separation: Turbulent vs.
`boundary flow close to an airfoil. Source: From Ref. 89.
`
`laminar
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`138
`
`Diebold
`
`of the convex surface, while the turbulent boundary layer in
`the second photograph remains attached longer, with the
`point of separation occurring further downstream. Turbulent
`layer separation occurs when the Reynolds stresses are much
`larger than the viscous stresses.
`
`Prerequisites for the Hydrodynamic Boundary
`Layer Concept
`
`Originally, the concept of the Prandtl boundary layer was
`developed for hydraulically ‘‘even’’ bodies. It is assumed that
`any characteristic length L on the particle surface is much
`greater than the thickness (dHL) of the boundary layer itself
`(L > dHL). Provided this assumption is fulfilled, the concept
`can be adapted to curved bodies and spheres, including ‘‘real’’
`drug particles. Furthermore, the classical (‘‘macroscopic’’)
`concept of the hydrodynamic boundary layer is valid solely
`for high Reynolds numbers of Re>104 (14,15). This constraint
`was overcome for the ‘‘microscopic’’ hydrodynamics of dissol-
`ving particles by the ‘‘convective diffusion theory’’ (9).
`
`The ‘‘Convective Diffusion Theory’’
`
`The ‘‘convective diffusion theory’’ was developed by V.G.
`Levich to solve specific problems in electrochemistry encoun-
`tered with the rotating disc electrode. Later, he applied the
`classical concept of the boundary layer to a variety of practical
`tasks and challenges, such as particle–liquid hydrodynamics
`and liquid–gas interfacial problems. The conceptual transfer
`of the hydrodynamic boundary layer is applicable to the
`hydrodynamics of dissolving particles if the Peclet number
`(Pe) is greater than unity (Pe > 1) (9). The dimensionless Pec-
`let number describes the relationship between convection and
`diffusion-driven mass transfer:
`
`Pe ¼
`
`UA L
`D
`D represents the diffusion coefficient. For example, low Peclet
`numbers indicate that convection contributes less to the total
`mass transfer and the latter is mainly driven by diffusion. In
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`139
`
`contrast, at high Peclet numbers, mass transfer is dominated
`by convection. The quotient of Pe and Re is called the Prandtl
`number (Pr), or, if we are talking about diffusion processes,
`the Schmidt number (Sc):
`
`¼ Sc
`
`n D
`

`
`Pe
`Re
`
`Pr ¼
`
`The Schmidt number is the ratio of kinematic viscosity to
`molecular diffusivity. Considering liquids in general and
`dissolution media in particular, the values for the kinematic
`viscosity usually exceed those for diffusion coefficients by a
`factor of 103 to 104. Thus, Prandtl or Schmidt numbers of
`about 103 are usually obtained. Subsequently, and in contrast
`to the classical concept of the boundary layer, Re numbers of
`magnitude of about Re  0.01 are sufficient to generate Peclet
`numbers greater than 1 and to justify the hydrodynamic
`boundary layer concept for particle–liquid dissolution systems
`(Re Pr ¼ Pe). It can be shown that [(9), term 10.15, nomen-
`clature adapted]
`
`d  D1=3 n
`
`1=6
`
`ffiffiffiffiffiffiffiL
`UAs
`
`Note that the hydrodynamic boundary layer depends on
`the diffusion coefficient. Introducing the proportionality
`constant K 
`e results in an equation valid for any desired
`hydrodynamic system based on relative fluid motion as pro-
`posed in Ref. 10:
`
`dHL  K 
`e D1=3 n
`
`1=6
`
`ffiffiffiffiffiffiffiL
`UAs
`
`K 
`e consists of a combination of Prandtl’s original proportion-
`ality constant used for the hydrodynamic boundary layer at a
`semi-infinitive plate, Ke, and a constant, K, characterizing a
`particular hydrodynamic system that is under consideration.
`The latter constant has to be determined experimentally.
`
`e ¼ Ke K 
`K 
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`140
`
`Diebold
`
`Among other factors, K is influenced by particle geome-
`try and surface morphology (roughness, edges, corners, and
`defects). For instance, K would equal 1 in the case of a
`smooth semi-infinite plate, and in this case K 
`e is identical
`to Ke. Considering the ‘‘rotating disc system’’ in particular,
`Levich found K to be 0.5. Given that a semi-infinite plate dis-
`solves in a liquid stream and Ke equals 5.2 (which represents
`Prandtl’s proportionality constant in the case of a semi-
` ¼ 2.6), we arrive at the following term
`infinite plate; thus Ke
`for the thickness of Levich’s effective hydrodynamic boundary
`layer (10):
`
`dHL  2:6 D1=3 n
`
`1=6
`
`ffiffiffiffiffiffiffiL
`UAs
`
`The Combination Model
`
`A reciprocal proportionality exists between the square root of
`the characteristic flow rate, UA, and the thickness of the effec-
`tive hydrodynamic boundary layer, dHL. Moreover, dHL
`depends on the diffusion coefficient D, characteristic length
`L, and kinematic viscosity n of the fluid. Based on Levich’s
`convective diffusion theory the ‘‘combination model’’ (‘‘Kombi-
`nations-Modell’’) was derived to describe the dissolution of
`particles and solid formulations exposed to agitated systems
`[(10), Chapter 5.2]. In contrast to the rotating disc method,
`the combination model is intended to serve as an approxima-
`tion describing the dissolution in hydrodynamic systems
`where the solid solvendum is not necessarily fixed but is likely
`to move within the dissolution medium. Introducing the term
`
`dHL  2:6 D1=3 n
`
`1=6
`
`UAs
`ffiffiffiffiffiffiffiL
`
`into the well-known equation adapted from Noyes, Whitney,
`Nernst, and Brunner
`
`dC
`dt
`

`
`A D
`dHL V
`
`ðCs CtÞ
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`141
`
`and employing the proportionality constant k as the apparent
`dissolution rate constant:
`
`k ¼
`
`A D
`dHL V
`results in the combination model according to Diebold (10):
`
` ðCs CtÞ
`
`A V
`
`
`
`A 1=2
`
`L U
`
`dC
`dt
`
`¼ 0:385 D2=3 n
`
`1=6
`
`where Cs represents the saturation solubility of the drug, Ct
`describes the concentration of the dissolved drug in the bulk
`at time t, D stands for the effective diffusion coefficient of
`the dissolved compound, A represents the total surface area
`accessible for dissolution of the drug particles, and V is the
`volume of the dissolution medium employed in the test. Note
`that the apparent dissolution rate constant k is now a
`function of the flow rate that a particle surface ‘‘sees’’ (slip
`velocity) and also a function of L, the characteristic length
`on the particle surface: k(UA; L). The proportionality constant
`k can be determined by appropriately performed dissolution
`experiments or calculated using the following equation:
`
`InðCs C0Þ InðCs CtÞ ¼ k t
`
`where C0 is the initial concentration of the drug at t ¼ 0. Since
`dHL is related to k as demonstrated above, the combination
`model permits calculation of an overall average hydrody-
`namic boundary layer for a given particle size fraction. Thus,
`the proposed relationship provides a tool for a priori predic-
`tion of
`the average hydrodynamic boundary layer of
`non-micronized drugs and hence to roughly forecast (!) disso-
`lution rate in vitro under well-defined circumstances, e.g., for
`the paddle apparatus [(10), Chapter 5.5, pp. 61–62, and
`Chapters 12.3.8 and 13.4.10].
`
`Further Factors Affecting the Hydrodynamic
`Boundary Layer
`
`Apart from the flow rate, of course, properties of the dissolu-
`tion medium as well as the drug compound influence the
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`142
`
`Diebold
`
`effective hydrodynamic boundary layer and hence the intrin-
`sic dissolution rate.
`
`Saturation Solubility (Cs)
`
`Although the saturation solubility (Cs) influences the appar-
`ent dissolution rate constant, it is an intrinsic property of a
`drug compound and can therefore affect the hydrodynamic
`boundary layer indirectly. High aqueous solubility, for exam-
`ple, leads to concentration-driven convection at the surface of
`the drug particles. Thus, forced and natural convection are
`mixed together, and it is challenging to separate/forecast
`their hydrodynamic effects on dissolution rate. In vivo disso-
`lution, however, offers additional problems to the control of
`hydrodynamics. The saturation solubility of a drug in intest-
`inal chyme may vary greatly within the course of dissolution
`in vivo, as has been demonstrated previously (10). The in vivo
`solubility of felodipine in jejunal chyme (37C), for example,
`was determined to be about 10 mg/mL on average (median),
`but varied greatly with time at mid-jejunum, ranging from
`1 to 25 mg/mL or even from 1 to 273 mg/mL, depending on
`the conditions of administration (10). Solubility variations
`within the course of an in vivo dissolution experiment may,
`in such cases, override hydrodynamic effects. Thus, the
`observed time dependency of intestinal drug solubility should
`be taken into account by dissolution models, which otherwise
`may describe dissolution rates in vitro well but fail to do so in
`vivo.
`
`Diffusion Coefficient (D)
`
`The diffusion coefficient is linked to the intrinsic dissolution
`rate constant (ki) as expressed by the term
`
`ki ¼
`
`D
`dHL
`Thus, the thickness of the effective hydrodynamic bound-
`ary layer dHL obviously depends on the diffusion coefficient.
`The diffusion coefficient D further correlates to the diameter
`of the particle or molecule as demonstrated by the relation
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`143
`
`of Stokes and Einstein:
`
`D ¼
`
`kB T
`3 d Z p
`
`where T is the temperature in Kelvin and kB represents the
`Boltzmann constant (1.381  1023 J/K). The term reveals that
`the diffusion coefficient D itself is dependent on the dynamic
`viscosity (Z). In the GI tract, diffusion coefficients of drugs
`may be reduced due to alterations in the fluid viscosity.
`Larhed et al. (16) reported that diffusion coefficients for testos-
`terone were reduced by 58% in porcine intestinal mucus. It has
`also been observed in dissolution experiments that the reduc-
`tion of diffusion coefficients can counteract effects of increased
`drug solubility due to mixed micellar solubilization (17).
`
`Kinematic Viscosity (n)
`
`The viscosity of upper GI fluids can be increased by food
`intake. The extent of this effect depends on the food compo-
`nents and the composition and volume of co-administered
`fluids. Aqueous-soluble fibers such as pectin, guar, and some
`hemicelluloses are able to increase the viscosity of aqueous
`solutions. Increasing the kinematic viscosity of the dissolu-
`tion medium generally leads to a reduction of the effective
`diffusion coefficient and hence results in decreased dissolu-
`tion. For instance, Chang et al. increased the viscosity of their
`dissolution media using guar as the model macromolecule.
`Subsequently, dissolution rates of benzoic acid were reduced
`significantly. However, dissolution rates were not at all
`affected when adjusting the same viscosity using propylene
`glycols (18).
`
`Temperature (T)
`
`The temperature influences the drug’s saturation solubility
`and also affects the kinematic viscosity (density of the liquid!)
`as well as the diffusion coefficient. Therefore, when performing
`dissolution experiments, temperature should be monitored
`carefully or preferably kept constant.
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`144
`
`Diebold
`
`Particle Morphology and Surface Roughness
`
`Faster initial dissolution rates obtained by grinding or milling
`the drug can often be attributed to both an increase in surface
`area and changes in surface morphology that lead to a higher
`surface free energy (19,20). However, an increase in edges,
`corners defects, and irregularities on the surfaces of coarse
`grade drug particles can also influence the effective hydrody-
`namic boundary layer dHL and hence dissolution rate (12,21–
`23). Depending on the surface roughness (R) of the drug par-
`ticle, the liquid stream near the particle surface may be tur-
`bulent even though the bulk flow remains laminar (9,10).
`Irregularities, edges, and defects increase the mass transfer
`in different ways according to the different kinds of hydrody-
`namic boundary layers generated. In the case of a turbulent
`boundary layer, the overall surface roughness is assumed to
`behave in a hydraulically ‘‘indifferent’’ (i.e., does not increase
`mass transfer itself) manner if the protrusions and cavita-
`tions are fully located within the viscous sub-layer (dVS).
`The so-called allowable (¼ indifferent) dimension of such a
`surface roughness (Rzul) can be estimated using an equation
`originally developed for tubes and pipes [(24), Sec. 21 d]:
`
`n U
`
`A
`
`Rzul ¼ 100
`
`For R < Rzul, the surface roughness does not cause per-
`turbations that increase mass transfer.
`In contrast,
`in the case of a laminar hydrodynamic
`boundary layer, the critical dimension of surface roughness
`(Rcrit) can be determined using the following relation:
`
`A
`
`UA L
`
`n ffi
`
`Rcrit ¼ 15
`
`with
`
`ffiffiffiffiffiffiffit=rp
`ffiffiffiffiffiffiffiffiffiffiffiffiffiffin
`r
`ffiffiffit
`rr ¼ 0:332 U2
`
`where t represents the shear stress, r is the fluid density, and
`n stands for the kinematic viscosity. If R > Rcrit, the effective
`
`© 2005 by Taylor & Francis Group, LLC
`
`

`

`Hydrodynamic Considerations
`
`145
`
`hydrodynamic boundary layer close to the particle wall
`becomes turbulent even though the bulk flow still may be
`laminar! In contrast to Rzul, Rcrit depends on the characteris-
`tic length L of the particle surface and is about 10 times
`greater [(24), Sec. 21 d]. In the case of a laminar hydrody-
`namic boundary, Levich (9,25) estimated that Rcrit could be
`exceeded for Re

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