throbber
DRUGS AND THE PHARMACEUTICAL SCIENCES
`
`VOLUME 157
`
`Pharmaceutical
`NH)eal
`
`SHRUTI
`
`IV_AER’s
`
`DiscoveryDevelopment Review Marketing
`Pre-clinical
`_Clinical
`i, vu
`Approval
`
`Continuous Learning — Manufacturing Science
`
`Pre-formulation
`
`Formulation (Clinical)
`Optimization
`
`(Optimization)
`Scale-Up
`
`Manufacturing
`Changes
`
`Quality by Design
`
`&
`Efficacy
`
`Appropriate labeling and risk management
`
`edited by
`Michael Levin
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 1 of 116
`
`
`
`

`

`Pharmaceutical
`Process Scale-Up
`Second Edition
`
`edited by
`Michael Levin
`Metropolitan Computing Corporation
`East Hanover, New Jersey, U.S.A.
`
`New York London
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 2 of 116
`
`

`

`Published in 2006 by
`CRC Press
`Taylor & Francis Group
`6000 Broken Sound Parkway NW, Suite 300
`Boca Raton, FL 33487-2742
`
`© 2006 by Taylor & Francis Group, LLC
`CRC Press is an imprint of Taylor & Francis Group
`
`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: 1-57444-876-5 (Hardcover)
`International Standard Book Number-13: 978-1-57444-876-4 (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
`
`Visit the Taylor & Francis Web site at
`http://www.taylorandfrancis.com
`
`Taylor & Francis Group
`is the Academic Division of T&F Informa plc.
`
`and the CRC Press Web site at
`http://www.crcpress.com
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 3 of 116
`
`

`

`Contents
`
`Introduction . . . .
`Preface . . . . xi
`Contributors . . . . xxi
`
`v
`
`1. Dimensional Analysis and Scale-Up in Theory and Industrial
`Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
`Marko Zlokarnik
`Introduction . . . . 1
`Dimensional Analysis . . . . 2
`Determination of a Pi Set by Matrix
`Calculation . . . . 8
`Fundamentals of the Theory of Models
`and of Scale-Up . . . . 12
`Further Procedures to Establish a Relevance List . . . . 14
`Treatment of Variable Physical Properties by
`Dimensional Analysis . . . . 23
`Pi Set and the Power Characteristics of
`a Stirrer in a Viscoelastic Fluid . . . . 29
`Application of Scale-Up Methods in Pharmaceutical
`Engineering . . . . 31
`Appendix . . . . 52
`References . . . . 53
`
`2. Engineering Approaches for Pharmaceutical Process Scale-Up,
`Validation, Optimization, and Control in the Process and
`Analytical Technology (PAT) Era . . . . . . . . . . . . . . . . . .
`Fernando J. Muzzio
`Introduction and Background . . . . 57
`
`57
`
`xv
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 4 of 116
`
`

`

`xvi
`
`Contents
`
`Model-Based Optimization . . . . 62
`Process Scale-Up . . . . 65
`Process Control
`. . . . 66
`Conclusions . . . . 68
`References . . . . 69
`
`3. A Parenteral Drug Scale-Up . . . . . . . . . . . . . . . . . . . . . .
`Igor Gorsky
`Introduction . . . . 71
`Geometric Similarity . . . . 72
`Dimensionless Numbers Method . . . . 74
`Scale-of-Agitation Approach . . . . 75
`Scale-of-Agitation Approach Example . . . . 78
`Latest Revisions of the Approach . . . . 80
`Scale-of-Agitation Approach for Suspensions . . . . 83
`Heat Transfer Scale-Up Considerations . . . . 85
`Conclusions . . . . 86
`References . . . . 87
`
`. . . . . . . . . . . . . .
`
`4. Non-Parenteral Liquids and Semisolids
`Lawrence H. Block
`Introduction . . . . 89
`Transport Phenomena in Liquids and Semisolids and Their
`Relationship to Unit Operations and Scale-Up . . . . 91
`How to Achieve Scale-Up . . . . 111
`Scale-Up Problems . . . . 123
`Conclusions . . . . 124
`References . . . . 125
`
`5. Scale-Up Considerations for
`. . . . . . . . . . . . . . . . . .
`Biotechnology-Derived Products
`Marco A. Cacciuttolo and Alahari Arunakumari
`Introduction . . . . 129
`Fundamentals: Typical Unit Operations . . . . 134
`Scale-Up of Upstream Operations . . . . 140
`Downstream Operations . . . . 146
`Process Controls . . . . 149
`Scale-Down Models . . . . 150
`Facility Design . . . . 150
`
`71
`
`89
`
`129
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 5 of 116
`
`

`

`Contents
`
`Examples of Process Scale-Up . . . . 151
`Impact of Scale-Up on Process Performance
`and Product Quality . . . . 154
`Summary . . . . 155
`Final Remarks and Technology Outlook . . . . 156
`References . . . . 157
`
`6. Batch Size Increase in Dry Blending and Mixing . . . . . . .
`Albert W. Alexander and Fernando J. Muzzio
`Background . . . . 161
`General Mixing Guidelines . . . . 162
`Scale-Up Approaches . . . . 165
`New Approach to the Scale-Up Problem
`in Tumbling Blenders . . . . 166
`Testing Velocity Scaling Criteria . . . . 173
`The Effects of Powder Cohesion . . . . 175
`Recommendations and Conclusions . . . . 178
`References . . . . 179
`
`7. Powder Handling . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`James K. Prescott
`Introduction . . . . 181
`Review of Typical Powder Transfer Processes . . . . 182
`Concerns with Powder-Blend Handling
`Processes . . . . 182
`Scale Effects . . . . 189
`References . . . . 197
`
`xvii
`
`161
`
`181
`
`199
`
`8. Scale-Up in the Field of Granulation
`and Drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`Hans Leuenberger, Gabriele Betz, and David M. Jones
`Introduction . . . . 199
`Theoretical Considerations . . . . 200
`The Dry-Blending Operation . . . . 201
`Scale-Up and Monitoring of the Wet
`Granulation Process . . . . 202
`Robust Formulations and Dosage Form Design . . . . 214
`A Quasi-Continuous Granulation and Drying Process
`(QCGDP) to Avoid Scale-Up Problems . . . . 214
`IPR2020-00770
`United Therapeutics EX2019
`Page 6 of 116
`
`

`

`xviii
`
`Contents
`
`Scale-Up of the Conventional Fluidized Bed Spray
`Granulation Process . . . . 220
`Summary . . . . 234
`References . . . . 235
`
`237
`
`9. Roller Compaction Scale-Up . . . . . . . . . . . . . . . . . . . . .
`Ronald W. Miller, Abhay Gupta, and
`Kenneth R. Morris
`Prologue . . . . 237
`Scale-Up Background . . . . 238
`Scale-Up Technical Illustrations . . . . 239
`Vacuum Deaeration Equipment Design Evaluation . . . . 241
`Conclusion . . . . 264
`References . . . . 265
`
`10. Batch Size Increase in Fluid-Bed Granulation . . . . . . . . .
`Dilip M. Parikh
`Introduction . . . . 267
`System Description . . . . 273
`Particle Agglomeration and Granule Growth . . . . 284
`Fluid-Bed Drying . . . . 288
`Process and Variables in Granulation . . . . 291
`Process Controls and Automation . . . . 300
`Process Scale-Up . . . . 305
`Case Study . . . . 310
`Material Handling . . . . 311
`Summary . . . . 316
`References . . . . 318
`
`11. Scale-Up of Extrusion and Spheronization . . . . . . . . . . .
`Raman M. Iyer, Harpreet K. Sandhu, Navnit H. Shah,
`Wantanee Phuapradit, and Hashim M. Ahmed
`Introduction . . . . 325
`Extrusion-Spheronization—An Overview . . . . 326
`Extrusion . . . . 328
`Spheronization . . . . 348
`Process Analytical Technologies (PAT) . . . . 361
`Summary . . . . 364
`References . . . . 365
`
`267
`
`325
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 7 of 116
`
`

`

`Contents
`
`xix
`
`371
`12. Scale-Up of the Compaction and Tableting Process . . . . .
`Alan Royce, Colleen Ruegger, Mark Mecadon, Anees Karnachi,
`and Stephen Valazza
`Introduction . . . . 371
`Compaction Physics . . . . 372
`Predictive Studies . . . . 375
`Scale-Up/Validation . . . . 388
`Case Studies . . . . 393
`Process Analytical Technology . . . . 405
`References . . . . 407
`
`13. Practical Considerations in the Scale-Up of Powder-Filled Hard
`409
`Shell Capsule Formulations . . . . . . . . . . . . . . . . . . . . . .
`Larry L. Augsburger
`Introduction . . . . 409
`Types of Filling Machines and Their Formulation
`Requirements . . . . 410
`General Formulation Principles . . . . 418
`Role of Instrumented Filling Machines
`and Simulation . . . . 420
`Scaling-Up Within the Same Design and Operating
`Principle . . . . 421
`Granulations . . . . 429
`References . . . . 430
`
`14. Scale-Up of Film Coating . . . . . . . . . . . . . . . . . . . . . . .
`Stuart C. Porter
`Introduction . . . . 435
`Scaling-Up the Coating Process . . . . 441
`Alternative Considerations to Scaling-Up
`Coating Processes . . . . 479
`Scale-Up of Coating Processes: Overall Summary . . . . 484
`References . . . . 484
`
`435
`
`15. Innovation and Continuous Improvement in Pharmaceutical
`Manufacturing
`. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`Ajaz S. Hussain
`Prologue . . . . 487
`The PAT Team and Manufacturing Science Working Group
`Report: A Summary of Learning, Contributions and Proposed
`IPR2020-00770
`United Therapeutics EX2019
`Page 8 of 116
`
`487
`
`

`

`xx
`
`Contents
`
`Next Steps for Moving Toward the ‘‘Desired State’’ of
`Pharmaceutical Manufacturing in the
`21st Century . . . . 488
`Bibliography and References . . . . 525
`
`Appendix . . . . 529
`Index . . . . 531
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 9 of 116
`
`

`

`4
`
`Non-Parenteral Liquids and Semisolids
`
`Lawrence H. Block
`Duquesne University, Pittsburgh, Pennsylvania, U.S.A.
`
`INTRODUCTION
`
`A manufacturer’s decision to scale-up (or scale-down) a process is ultimately
`rooted in the economics of the production process, i.e., in the cost of material,
`personnel, and equipment associated with the process and its control. While
`process scale-up often reduces the unit cost of production and is therefore
`economically advantageous per se, there are additional economic advantages
`conferred on the manufacturer by scaling-up a process. Thus, process scale-up
`may allow for faster entry of a manufacturer into the marketplace or improved
`product distribution or response to market demands and correspondingly
`greater market-share retention.a Given the potential advantages of process
`scale-up in the pharmaceutical industry, one would expect the scale-up task
`to be the focus of major efforts on the part of pharmaceutical manufacturers.
`However, the paucity of published studies or data on scale-up—particularly
`for non-parenteral liquids and semisolids—suggests otherwise. On the other
`hand, one could argue that the paucity of published studies or data is nothing
`
`a On the other hand, the manufacturer may determine that the advantages of process scale-up
`are compromised by the increased cost of production on a larger scale and/or the potential loss
`of interest or investment income. Griskey (1) addresses the economics of scale-up in some detail
`in his chapter on engineering economics and process design, but his examples are taken from the
`chemical industry. For a more extensive discussion of process economics, see Ref. 2.
`
`89
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 10 of 116
`
`

`

`90
`
`Block
`
`more than a reflection of the need to maintain a competitive advantage
`through secrecy.
`One could also argue that this deficiency in the literature attests to the
`complexity of the unit operations involved in pharmaceutical processing. If
`pharmaceutical technologists view scale-up as little more than a ratio
`problem, whereby
`Scale-up ratio ¼ Large-scale production rate
`Small-scale production rate
`
`ð1Þ
`
`then the successful resolution of a scale-up problem will remain an empirical,
`trial-and-error task rather than a scientific one. In 1998, in a monograph on
`the scale-up of disperse systems, Block (3) noted that due to the complexity
`of the manufacturing process which involves more than one type of unit
`operationb (e.g., mixing, transferring, etc.), process scale-up from the bench
`or pilot plant level to commercial production is not a simple extrapolation:
`
`‘‘The successful linkage of one unit operation to another defines the
`functionality of the overall manufacturing process. Each unit oper-
`ation per se may be scalable, in accordance with a specific ratio, but
`the composite manufacturing process may not be, as the effective
`scale-up ratios may be different from one unit operation to another.
`Unexpected problems in scale-up are often a reflection of the dichot-
`omy between unit operation scale-up and process scale-up. Further-
`more, commercial production introduces problems that are not a
`major issue on a small scale: e.g., storage and materials handling
`may become problematic only when large quantities are involved;
`heat generated in the course of pilot plant or production scale proces-
`sing may overwhelm the system’s capacity for dissipation to an extent
`not anticipated based on prior laboratory-scale experience (3).’’
`
`Furthermore, unit operations may function in a rate-limiting manner
`as the scale of operation increases. When Astarita (4) decried the fact, in
`the mid-1980s, that ‘‘there is no scale-up algorithm which permits us to
`rigorously predict the behavior of a large scale process based upon the beha-
`vior of a small scale process,’’ it was presumably as a consequence of all of
`these problematic aspects of scale-up.
`A clue to the resolution of the scale-up problem for liquids and semi-
`solids resides in the recognition that their processing invariably involves the
`
`b The term unit operations, coined by Arthur D. Little in 1915, is generally used to refer to
`distinct physical changes or unit actions (e.g., pulverizing, mixing, drying, etc.) while unit opera-
`tions involving chemical changes are sometimes referred to as unit processes. The physical
`changes comprising unit operations primarily involve contact, transfer of a physical property,
`and separation between phases or streams.
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 11 of 116
`
`

`

`Non-Parenteral Liquids and Semisolids
`
`91
`
`unit operation of mixing. Closer examination of this core unit operation
`reveals that flow conditions and viscosities during processing can vary by
`several orders of magnitude depending upon the scale of scrutiny employed,
`i.e., whether on a microscopic (e.g., mm to cm) or macroscopic (e.g., cm to m)
`scale. The key to effective processing scale-up is the appreciation and under-
`standing of microscale and macroscale transport phenomena, i.e., diffusion
`and bulk flow, respectively. Transport by diffusion involves the flow of a
`property (e.g., mass, heat, momentum, and electromagnetic energy) from
`a region of high concentration to a region of low concentration as a result
`of the microscopic motion of electrons, atoms, molecules, etc. Bulk flow,
`whether convection or advection, however, involves the flow of a property
`as a result of macroscopic or bulk motion induced artificially (e.g., by mech-
`anical agitation) or naturally (e.g., by density variations) (5).
`
`TRANSPORT PHENOMENA IN LIQUIDS AND SEMISOLIDS AND
`THEIR RELATIONSHIP TO UNIT OPERATIONS AND SCALE-UP
`
`Over the last four decades or so, transport phenomena research has
`benefited from the substantial efforts made to replace empiricism by funda-
`mental knowledge based on computer simulations and theoretical modeling
`of transport phenomena. These efforts were spurred on by the publication in
`1960 by Bird et al. (6) of the first edition of their quintessential monograph
`on the interrelationships among the three fundamental types of transport
`phenomena: mass transport, energy transport, and momentum transport.c
`All transport phenomena follow the same pattern in accordance with the
`generalized diffusion equation (GDE). The unidimensional flux, or overall
`transport rate per unit area in one direction, is expressed as a system prop-
`erty multiplied by a gradient (5)
`
`
`
`

`
`@G
`@t
`
`¼ d
`
`x
`
`@2G
`@x2
`
`
`
`
`
`!
`
`
`
`@G
`@x
`@x
`
`¼ d
`
` 
`
`¼ d
`
`@E
`@x
`
`ð2Þ
`
`where G represents the concentration of a property Q (e.g., mass, heat, elec-
`trical energy, etc.) per unit volume, i.e., G ¼ Q/V, t is time, x is the distance
`measured in the direction of transport, d is the generalized diffusion coeffi-
`cient, and E is the gradient or driving force for transport.
`Mass and heat transfer can be described in terms of their respective
`concentrations Q/V. While the concentration of mass, m, can be specified
`directly, the concentration of heat is given by
`
`c The second edition of Transport Phenomena was published in 2002, 42 years later, an
`indication of the utility of the first edition and its continuing acceptance by the engineering
`discipline.
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 12 of 116
`
`

`

`92
`
`mCpT
`V
`
`¼ rCpT
`
`Block
`
`ð3Þ
`
`where Cp is the specific heat capacity and T is temperature. Thus the speci-
`fication of rCpT in any form of the generalized diffusion equation will result
`in the elimination of rCp, assuming it to be a constant, thereby allowing the
`use of temperature as a measure of heat concentration (5). In an analogous
`manner, momentum transfer can be specified in terms of the concentration
`of momentum u when its substantial derivative is used instead of its partial
`derivative with respect to time
`
`¼ vH2u
`
`Du
`Dt
`
`ð4Þ
`
`where v is the kinematic viscosity. If pressure and gravitational effects
`are introduced, one arrives at the Navier–Stokes relationships that govern
`Newtonian fluid dynamics.
`When the flux of G is evaluated three-dimensionally, it can be repre-
`sented by (5)
`¼ @G
`@t
`
`dz
`@t
`
`ð5Þ
`
`dG
`dt
`
`þ @G
`@x
`
`dx
`@t
`
`þ @G
`@y
`
`dy
`@t
`
`þ @G
`@z
`
`At the simplest level, as Griskey (1) notes, Fick’s law of diffusion for
`mass transfer and Fourier’s law of heat conduction characterize mass and
`heat transfer, respectively, as vectors, i.e., they have magnitude and direc-
`tion in the three coordinates, x, y, and z. Momentum or flow, however, is
`a tensor which is defined by nine components rather than three. Hence,
`its more complex characterization at the simplest level, in accordance with
`Newton’s law, is
`
` 
`
`dvx
`dy
`
`tyx ¼ Z
`where tyx is the shear stress in the x-direction, ðdvx=dyÞ the rate of shear,
`and Z is the coefficient of Newtonian viscosity. The solution of Equation (2),
`the generalized diffusion equation,
`G ¼ f ðt; x; y; zÞ
`
`ð6Þ
`
`ð7Þ
`
`will take the form of a parabolic partial differential equation (5). However,
`the more complex the phenomenon—e.g., with convective transport a part
`of the model—the more difficult it is to achieve an analytic solution to
`the GDE. Numerical solutions, however, where the differential equation is
`transformed to an algebraic one, may be somewhat more readily achieved.
`IPR2020-00770
`United Therapeutics EX2019
`Page 13 of 116
`
`

`

`Non-Parenteral Liquids and Semisolids
`
`93
`
`Transport Phenomena and Their Relationship to Mixing
`as a Unit Operationd
`
`As noted earlier, virtually all liquid and semisolid products involve the unit
`of operation of mixing.e In fact, in many instances, it is the primary unit
`operation. Even its indirect effects, e.g., on heat transfer, may be the basis
`for its inclusion in a process. Yet, mechanistic and quantitative descriptions
`of the mixing process remain incomplete (7–9). Nonetheless, enough funda-
`mental and empirical data are available to allow some reasonable predic-
`tions to be made.
`The diversity of dynamic mixing devices is unsettling: their dynamic,
`or moving, component’s blades may be impellers in the form of propellers,
`turbines, paddles, helical ribbons, Z-blades, or screws. In addition, one can
`vary the number of impellers, the number of blades per impeller, the pitch of
`the impeller blades, and the location of the impeller, and thereby affect
`mixer performance to an appreciable extent. Furthermore, while dispersators
`or rotor/stator configurations may be used rather than impellers to effect mix-
`ing, mixing may also be accomplished by jet-mixing or static-mixing devices.
`The bewildering array of mixing equipment choices alone would appear to
`make the likelihood of effective scale-up an impossibility. However, as diverse
`as mixing equipment may be, evaluations of the rate and extent of mixing and
`of flow regimesf make it possible to find a common basis for comparison.
`In low-viscosity systems, miscible liquid blending is achieved through
`the transport of unmixed material via flow currents (i.e., bulk or convective
`flow) to a mixing zone (i.e., a region of high shear or intensive mixing). In
`other words, mass transport during mixing depends on streamline or laminar
`flow, involving well-defined paths, and turbulent flow, involving innumer-
`able, variously sized, eddies, or swirling motions. Most of the highly turbu-
`lent mixing takes place in the region of the impeller, fluid motion elsewhere
`serving primarily to bring fresh fluid into this region. Thus, the characteriza-
`tion of mixing processes is often based on the flow regimes encountered in
`mixing equipment. Reynolds’ classic research on flow in pipes demonstrated
`
`dReprinted in part, with revisions and updates, by courtesy of Marcel Dekker, Inc., from L. H.
`Block, ‘‘Scale-up of disperse systems: Theoretical and practical aspects,’’ in Pharmaceutical
`Dosage Forms: Disperse Systems (H. A. Lieberman, M. M. Rieger, and G. S. Banker, eds.),
`Vol. 3, 2nd ed., New York: Marcel Dekker, 1998:366–378.
`eMixing, or blending, refers to the random distribution of two or more initially separate phases
`into and through one another, while agitation refers only to the induced motion of a material in
`some sort of container. Agitation does not necessarily result in an intermingling of two or more
`separate components of a system to form a more or less uniform product. Some authors reserve
`the term blending for the intermingling of miscible phases while mixing is employed for materials
`that may or may not be miscible.
`f The term flow regime is used to characterize the hydraulic conditions (i.e., volume, velocity,
`and direction of flow) within a vessel.
`IPR2020-00770
`United Therapeutics EX2019
`Page 14 of 116
`
`

`

`94
`
`Block
`
`that flow changes from laminar to irregular, or turbulent, once a critical
`value of a dimensionless ratio of variables has been exceeded (10,11). This
`ratio, universally referred to as the Reynolds number, NRe, is defined by
`Equations (8a) and (8b),
`
`NRe ¼ Lvr
`Z
`
`NRe ¼ D2Nr
`Z
`
`ð8aÞ
`
`ð8bÞ
`
`where r is the density, v the velocity, L the characteristic length, and Z is the
`Newtonian viscosity; Equation (8b) is referred to as the impeller Reynolds
`number, as D is the impeller diameter and N is the rotational speed of the
`impeller. NRe represents the ratio of the inertia forces to the viscous forces
`in a flow. High values of NRe correspond to flow dominated by motion
`while low values of NRe correspond to flow dominated by viscosity. Thus,
`the transition from laminar to turbulent flow is governed by the density
`and viscosity of the fluid, its average velocity, and the dimensions of the
`region in which flow occurs (e.g., the diameter of the pipe or conduit, the
`diameter of a settling particle, etc.). For a straight circular pipe, laminar
`flow occurs when NRe < 2100; turbulent flow is evident when NRe > 4000.
`For 2100  NRe  4000, flow is in transition from a laminar to a turbulent
`
`regime. Other factors such as surface roughness, shape and cross-sectional
`area of the affected region, etc., have a substantial effect on the critical value
`of NRe. Thus, for particle sedimentation, the critical value of NRe is 1; for
`some mechanical mixing processes, NRe is 10–20 (12). The erratic, relatively
`unpredictable nature of turbulent eddy flow is further influenced, in part, by
`the size distribution of the eddies which are dependent on the size of the
`apparatus and the amount of energy introduced into the system (10). These
`factors are indirectly addressed by NRe. Further insight into the nature of
`NRe can be gained by viewing it as inversely proportional to eddy advection
`time, i.e., the time required for eddies or vortices to form.
`In turbulent flow, eddies move rapidly with an appreciable component
`of their velocity in the direction perpendicular to a reference point, e.g., a sur-
`face past which the fluid is flowing (13). Because of the rapid eddy motion,
`mass transfer in the turbulent region is much more rapid than that resulting
`from molecular diffusion in the laminar region, with the result that the
`concentration gradients existing in the turbulent region will be smaller than
`those in the laminar region (13). Thus, mixing is much more efficient under
`turbulent flow conditions. Nonetheless, the technologist should bear in mind
`potentially compromising aspects of turbulent flow, e.g., increased vortex for-
`mation (14) and a concomitant incorporation of air, increased shear and a
`corresponding shift in the particle size distribution of the disperse phase, etc.
`IPR2020-00770
`United Therapeutics EX2019
`Page 15 of 116
`
`

`

`Non-Parenteral Liquids and Semisolids
`
`95
`
`Although continuous-flow mixing operations are employed to a limited
`extent in the pharmaceutical industry, the processing of liquids and semisolids
`most often involves batch processing in some kind of tank or vessel. Thus, in the
`general treatment of mixing that follows, the focus will be on batch operationsg
`in which mixing is accomplished primarily by the use of dynamic mechanical
`mixers with impellers, although jet mixing (17,18) and static mixing devices
`(19)—long used in the chemical process industries—are being used to an
`increasingly greater extent now in the pharmaceutical and cosmetic industries.
`Mixers share a common functionality with pumps. The power impar-
`ted by the mixer, via the impeller, to the system is akin to a pumping effect
`and is characterized in terms of the shear and flow produced as
`P / QrH
`
`or
`
`H / P
`Qr
`
`ð9Þ
`
`ð10Þ
`
`N
`
`dyne
`
`where P is the power imparted by the impeller, Q the flow rate (or pumping
`capacity) of material through the mixing device, r the density of the mate-
`rial, and H is the velocity head or shear. Thus, for a given P, there is an
`inverse relationship between shear and volume throughput.
`The power input in mechanical agitation is calculated using the power
`number, NP,
`NP ¼ Pgc
`
`
`rN 3D5
`¼ g  cm  sec
`where gc is the force conversion factor gc ¼ kg  m  sec
`2
`2
`, N the
`1), and D is the diameter of the impeller. For a
`impeller rotational speed (sec
`given impeller/mixing tank configuration, one can define a specific relationship
`between the Reynolds number [Eq. (8)]h and the power number [Eq. (10)] in
`which three zones (corresponding to the laminar, transitional, and turbulent
`regimes) are generally discernible. Tatterson (20) notes that for mechanical agi-
`
`tation in laminar flow, most laminar power correlations reduce to NpNRe¼ B,
`where B is a complex function of the geometry of the system,i and that this is
`equivalent to P / Z h N2D3; ‘‘if power correlations do not reduce to this form
`
`g The reader interested in continuous-flow mixing operations is directed to references that deal
`specifically with that aspect of mixing such as the monographs by Oldshue (15) and
`Tatterson (16).
`for mixing is defined in SI-derived units as NRe ¼
`h Here,
`the Reynolds number
`5ND2 r)/Z, where D, impeller diameter, is in mm.; Z is in Pas; N is impeller speed,
`(1.667 10
`in r.p.m.; and r is density.
`i An average value of B is 300, but B can vary between 20 and 4000 (20).
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 16 of 116
`
`

`

`96
`
`Block
`
`
`
` ¼ v/ND;
`Figure 1 Various dimensionless parameters [dimensionless velocity, v
`pumping number, NQ ¼ Q/ND3; power number, NP ¼ Pgc=rN 3D5
`; and dimension-
` ¼ tmN] as a function of the Reynolds number for the analysis of
`
`less mixing time, t
`turbine–agitator systems. Source: Adapted from Ref. 22.
`
`for laminar mixing, then they are wrong and should not be used.’’ Turbulent
`correlations are much simpler: for systems employing baffles,j NP¼ B; this is
`equivalent to P / rN3D5. Based on this function, slight changes in D can
`result in substantial changes in power.
`Impeller size relative to the size of the tank is critical as well. If the ratio of
`impeller diameter D to tank diameter T is too large (D/T is > 0.7), mixing
`efficiency will decrease as the space between the impeller and the tank wall will
`be too small to allow a strong axial flow due to obstruction of the recirculation
`path (21). More intense mixing at this point would require an increase
`in impeller speed, but this may be compromised by limitations imposed by
`impeller blade thickness and angle. If D/T is too small, the impeller will not
`be able to generate an adequate flow rate in the tank.
`Valuable insights into the mixing operation can be gained from a
`consideration of system behavior as a function of the Reynolds number,
`NRe (22). This is shown schematically in Figure 1 in which various dimension-
`less parameters (dimensionless velocity, v/ND; pumping number, Q/ND3;
`power number, NP = Pgc=rN 3 D5
`; and dimensionless mixing time, tmN)
`are represented as a log–log function of NRe. Although density, viscosity,
`mixing vessel diameter, and impeller rotational speed are often viewed by
`
`
`
`
`
`j Baffles are obstructions placed in mixing tanks to redirect flow and minimize vortex formation.
`Standard baffles—comprising rectangular plates spaced uniformly around the inside wall of a
`tank—convert rotational flow into top-to-bottom circulation.
`
`IPR2020-00770
`United Therapeutics EX2019
`Page 17 of 116
`
`

`

`Non-Parenteral Liquids and Semisolids
`
`97
`
`formulators as independent variables, their interdependency, when incorpo-
`rated in the dimensionless Reynolds number, is quite evident. Thus, the sche-
`matic relationships embodied in Figure 1 are not surprising.k
`Mixing time is the time required to produce a mixture of predetermined
`quality; the rate of mixing is the rate at which mixing proceeds towards the
`final state. For a given formulation and equipment configuration, mixing
`time, tm, will depend upon material properties and operation variables. For
`geometrically similar systems, if the geometrical dimensions of the system
`are transformed to ratios, mixing time can be expressed in terms of a dimen-
`sionless number, i.e., the dimensionless mixing time, ym or tmN
`tmN ¼ ym ¼ f NRe; NFrð
`Þ ¼) f NReð

`
`
`ffiffiffiffiffiffi
`pð
`The Froude number, NFr ¼ n=
`Þ, is similar to NRe; it is a measure
`
`ð11Þ
`
`Lg
`of the inertial stress to the gravitational force per unit area acting on a fluid.
`Its inclusion in Equation (11) is justified when density differences are
`encountered; in the absence of substantive differences in density, e.g., for
`emulsions more so than for suspensions, the Froude term can be neglected.
`Dimensionless mixing time is independent of the Reynolds number for both
`laminar and turbulent flow regimes as indicated by the plateaus in Figure 1.
`Nonetheless, as there are conflicting data in the literature regarding the sen-
`sitivity of ym to the rheological properties of the formulation and to equip-
`ment geometry, Equation (11) must be regarded as an oversimplification of
`the mixing operation. Considerable care must be exercised in applying the
`general relationship to specific situations.
`Empirical correlations for turbulent mechanical mixing have been
`reported in terms of the following dimensionless mixing time relationship (24)
`ð12Þ
`
`a
`
` 
`
`T D
`
`ym ¼ tmN ¼ K
`
`where K and a are constants, T is tank diameter, N is impeller rotational speed,
`and D is the impeller diameter. Under laminar flow conditions, Equation (12)
`reduces to
`
`ð13Þ
`ym ¼ H0
`where H0 is referred to as the mixing number or homogenization number. In
`the transitional flow regime,
`H0 ¼ C NReð
`
`Þa
`where C and a are constants, with a varying between 0 and 1.
`
`ð14Þ
`
`k The interrelationships are embodied in variations of the Navier–Stokes equations, which
`describe mass and momentum balances in fluid systems (23).
`IPR2020-00770
`United Therapeutics EX2019
`Page 18 of 116
`
`

`

`98
`
`Block
`
`Flow patterns in agitated vessels may be characterized as radial, axial,
`or tangential relative to the impeller, but are more correctly defined by the
`direction and magnitude of the velocity vectors throughout the system, par-
`ticularly in a transitional flow regime; while the d

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