throbber
CHEMICAL KINETICS
`
`EDITED BY
`
`C.H. BAMFORD
`
`M.A., Ph.D., Sc.D. (Cantab.), F.R.I.C., F.R.S.
`Formerly Campbell-Brown Professor of Industrial Chemistry,
`University of Liverpool
`
`AND
`
`R.G. COMPTON
`
`M.A., D.Phil. (Oxon.)
`University Lecturer in Physical Chemistry
`and Fellow, St. John’s College, Oxford
`
`VOLUME 26
`
`ELECTRODE KINETICS:
`PRINCIPLES AND METHODOLOGY
`
`
`
`ELSEVIER
`AMSTERDAM—OXFORD—NEW YORK—TOKYO
`1986
`
`InnoPharma Exhibit 1104.0001
`
`
`
`

`

`ELSEVIER SCIENCE PUBLISHERS B.V.
`
`Sara Burgerhartstraat 25
`PO. Box 211, 1000 AE Amsterdam, The Netherlands
`
`Distributors for the United States and Canada
`
`ELSEVIER SCIENCE PUBLISHING COMPANY INC.
`52 Vanderbilt Avenue
`
`New York, NY 10017
`
`
`
`ISBN 0-444-41631-5 (Series)
`ISBN 0-444-42550—0 (Vol. 26)
`
`with 120 illustrations and 55 tables
`
`© Elsevier Science Publishers B.V., 1986
`
`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, photo-
`copying, recording or otherwise, Without the prior written permission of the publisher,
`Elsevier Science Publishers B.V./Science & Technology Division, PO. Box 330, 1000 AH
`Amsterdam, The Netherlands.
`
`Special regulations for readers in the USA — This publication has been registered with the
`Copyright Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained
`from the 000 about conditions under which photocopies of parts of this publication may
`be made in the USA. All other copyright questions, including photocopying outside of
`the USA, should be referred to the publishers.
`
`Printed in the Netherlands
`
`InnoPharma Exhibit 1104.0002
`
`

`

`
`This material may be protected by Copyright law (Title 17 U.S. Code)
`
`
`
`Chapter 2
`
`Mass Transport to Electrodes
`
`KEITH B. OLDHAM and CYNTHIA G. ZOSKI
`
`1. Introduction
`
`The investigation of electrode kinetics has one paramount advantage
`over other kinetic studies: the rate of the electron transfer reaction
`
`'
`
`Reactants + n e
`
`Products
`
`(1)
`
`can be measured directly rather than needing to be inferred from concen-
`tration changes. This advantage is a consequence of Faraday’s law, which
`asserts the proportionality of the electron-transfer rate
`
`1'
`nAF
`
`2
`
`)
`
`(
`
`Reaction rate =
`
`to the faradaic current 1' divided by the elctrode area A. In eqn. (2), n
`is the number of electrons and F is Faraday’s constant.
`On the other hand, electrode kinetic studies are at a disadvantage
`compared with investigations of homogeneous kinetics because con-
`centrations are not uniform and surface concentrations can rarely be
`measured directly (optical methods can sometimes provide direct measure-
`ment of the product concentration [1]). This means that the converse
`situation to that in classical homogeneous kinetics exists in electrode
`kinetics: concentration information needs to be inferred from reaction
`rates.
`
`the electrode surface requires a
`
`To calculate concentrations at
`knowledge of
`(a) the stoichiometry of the electrode reaction;
`(b) the bulk concentrations of the species involved;
`(c) the rate of the reaction [or equivalently, because of relationship
`(2), the faradaic current] since the onset of the experiment;
`(d) the laws governing mass transport
`for the particular electrode
`geometry; and
`(e) the prevailing experimental conditions.
`This chapter is concerned with how one uses items (a)-(e) to calculate
`concentrations at the electrode surface. In the electrochemical literature,
`expressions for surface concentrations are seldom regarded as the end
`result of a transport prediction. Instead, one usually assumes that the
`surface concentrations of the species involved in the electrode reaction
`
`References pp. 1 41—1 43
`
`InnoPharma Exhibit 1104.0003
`
`

`

`80
`
`thermodynamic relation (nernstian conditions) or a
`a
`obey either
`particular kinetic expression (volmerian conditions), so that the result of
`the analysis of mass transport can be presented as a relationship between
`the experimentally observable variables: current, cell potential, and time.
`In this chapter, we shall primarily report relationships involving concentra-
`tions, since these are the kinetically significant variables.
`
`1.1 SPECIES INVOLVED IN TRANSPORT
`
`specifically exclude electrode
`shall
`this chapter, we
`Throughout
`reactions that consume or generate insoluble species. Thus,
`the most
`general electrode reaction is
`
`VA A(soln) + VB B(soln). .
`
`. i n e——> VzZ(soln) + VY Y(soln) +. .
`
`.
`
`(3)
`
`. are reactant species
`.
`where the 12s are stiochiometric coefficients, A, B. .
`and Z, Y. .
`. are product species. Usually, all the species are dissolved in
`the electrolyte solution, but an important exception occurs in the
`reduction of certain metal
`ions at a mercury electrode to produce an
`amalgam
`
`M"+(soln) + n e (Hg)———>M(amal)
`
`(4)
`
`Cases in which one of the reactants or products is the material of the
`electrode itself, as in
`
`2 Hg(l) _ 2 e (Hg)—_’ Hg? (301)
`
`or
`
`CuCl§‘(aq) + 2 e (Cu) ——>Cu(s) + 4 Cl‘(aq)
`
`are not excluded.
`
`(5)
`
`(6)
`
`There is, of course, always an ample supply of the electrolytic solvent
`(often water) and of the electrode material (often a metal) at
`the
`interface, but all other species must be transported to and from the
`electrode surface as illustrated in Fig. 1.
`Usually,
`there is no significant impediment to the transport of elec-
`trons,
`through this may not be true of some semiconductor electrodes
`[2—7]. When there is more than one reactant, it is usually possible to
`adjust bulk concentrations (or adopt other experimental strategies such as
`buffering) so that all reactants except one are in such excess that their
`transport poses no difficulty. This is an analog of the “isolation technique”
`familiar to kineticists. The same is true of product species: it is generally
`possible to arrange experimental conditions so that, at most, only one
`product species is subject to a transport restriction.
`Hence, we shall customarily ignore all but one reactant species and all
`but one product species. Moreover, we shall assume that the stoichio-
`metric coefficients of these species are both unity. This is not an essential
`
`InnoPharma Exhibit 1104.0004
`
`

`

`81
`
`919C
`
`thong
`
`electrode
`
`[25
`
`reacton
`
`solution
`
`(amalgam)
`
`Product-.5
`
`T
`91 90 trade
`sur‘Face
`
`Fig. 1. Transport to and from the electrode surface.
`
`assumption, but it does serve to simplify our arguments and covers the
`majority of practical examples. Thus, for a reduction experiment, the
`electrode reaction may be abbreviated to
`
`O(soln) + n e —> R(soln)
`
`(7)
`
`where O (for oxidized species) is the single reactant species we need
`consider and R (for reduced species) is the sole product species under
`consideration. Of course, either 0 or R or both may be ions.
`Sometimes, we shall address an even simpler class of electrode reaction
`in which there is only a single electroactive species of a varible activity.
`The simplest
`instance of this class is the reduction of metal
`ions on
`a cathode composed of that metal, for example
`
`M"+(soln) + n e(M)—>M(s)
`
`(8)
`
`This reaction is the only one treated in Sect. 4 of this chapter.
`It will be our custom to deal with cathodic electrode reactions, i.e.
`with reductions like (7 ) and (8), rather than with the equally important
`oxidation processes. For this reason, cathodic currents will be treated as
`positive*.
`
`1.2 THE ELECTRODE SURFACE
`
`The region extending from the phase boundary out to about 3nm is
`quite unlike the solution beyond. Generalizations valid elsewhere in the
`solution do not necessarily apply here. In this inner zone, the so-called
`double-layer region [9], we may encounter a violation of the electro-
`neutrality condition (see' Sect. 4.1) and large electric fields. Concen-
`trations may be enhanced or depleted compared with the adjacent
`solution.
`
`* This is the usual convention in electroanalytical chemistry, though it is at variance
`with the more logical IUPAC convention [8].
`
`References pp. 141—143
`
`InnoPharma Exhibit 1104.0005
`
`

`

`82
`
`Through phenomena inside the double-layer region may have profound
`effects on the kinetics of electrode reactions [10] , the zone is fortunately
`of little or no consequence in discussions of transport from the solution
`to the electrode or vice versa. To appreciate why this is so, consider
`a typical electrochemical experiment in which species R is being generated
`at an electrode by a current of density 1.0 A m'z. After 1 s of electrolysis,
`the transport zone (the region into which R is being carried) will extend
`into the solution a distance of some 10 ,um, in comparison with which the
`double-layer
`thickness of 3 nm is negligible, Moreover, each electro-
`generated R molecule will
`transit the double-layer region in less than
`200 us.
`in this chapter, concentrations “at the electrode
`When we discuss,
`surface”, we will generally ignore the effect of the double layer. Hence,
`“at the electrode surface” really means “outside the double-layer” or,
`more precisely, “extrapolating to the electrode surface -the concentration
`profile from beyond the double layer”, as illustrated in Fig. 2.
`The narrow double-layer zone is the site of all the chemical and physico-
`chemical processes that attend the electrode reaction. These may include,
`in addition to the electron transfer itself, adsorption and desorption steps,
`as well as chemical transformation between 0 and R (which are the
`species stable in the bulk of the solution) and modified species (with less
`solvation, perhaps, or with different configurations) o and r which are
`adsorbable. Thus, the complete train of events may be
`
`O(soln)
`
`R(soln)
`double-1a er region
`trans-
`transport
`y
`port
`O(soln) fi O(soln) <——> o(ads)X—:r(ads) = r(soln) # R(soln)
`
`or some even more elaborate scheme. Most of these complexities need not
`concern us in this chapter, but should be noted.
`Let [‘0 and PR (with units of mol m'z) denote the amounts of O and
`R (or their modified forms 0 and r) that are adsorbed. The shaded area in
`
`concentration
`
`" surface
`concentration "
`\a
`
`
`> distance
`H————9
`double layer
`
`Fig. 2. Illustration of electrode “surface concentration”.
`
`InnoPharma Exhibit 1104.0006
`
`

`

`83
`
`Fig. 2 illustrates the significance of F. The validity of certain equations
`that we shall develop in Sect. 2.2 requires that F0 and PR remain
`constant. Certain experimental methods, both thermodynamic and
`kinetic*, can be used to investigate the extent of adsorption on electrodes,
`but these are beyond the scope of the present discussion.
`
`1.3 ELECTRODE GEOMETRY
`
`We stipulate the electrode to be smooth (though not necessarily flat)
`and of constant area A. By “smooth” we mean that any undulations in
`the electrode surface should not exceed the thickness of the double layer.
`For an electrode that is less smooth than this, the concept of electrode
`area is somewhat vague and the “effective electrode area” may change
`with time. By prescribing a constant electrode area, we exclude one of
`the most practical electrodes: the dropping mercury electrode treated in
`.Chap. 5.
`Where it is possible to define the coordinate unambiguously, we shall
`use x to denote distance measured normal to the electrode into the trans-
`
`port medium, with x = 0 corresponding to the electrode surface. We
`may distinguish three simple types of electrode geometry: planar, convex,
`and concave, as shown in Fig. 3. Transport to a planar electrode occurs
`along parallel
`lines. For a convex electrode, transport to the electrode
`is convergent, whereas transport away from the electrode is divergent.
`The opposite is
`true for a concave electrode. The case of a convex
`mercury electrode at which reaction (4) occurs is unusual but important;
`there, the transport of both 0 and R occurs convergently.
`is
`The region around the electrode, through which transport occurs,
`filled with solution and should usually be unimpeded by cell walls, other
`electrodes, etc. for a sufficient distance. How far is “sufficient” depends
`upon the mode of transport and on the duration of the experiment. For
`transport by diffusion alone, the requirements are very modest indeed,
`as indicated in Table 1.
`The electrode at which the reaction under study is proceeding is called
`the “working electrode”. It is with this one electrode that we are solely
`concerned. There is, however, an important exception: the experiments
`
`/<:
`
`:C
`/C
`
`(a)
`
`/
`
`/<:
`/V§
`
`/
`
`:<:
`/ fl/
`
`(b)
`
`'
`
`(c)
`
`Fig. 3. Types of electrode geometry. (a) Planar; (b) convex; (c) concave. The arrows
`indicate transport lines to the electrode surface.
`
`* Especially chronocoulometry [11—14].
`
`References pp. 141—143
`
`InnoPharma Exhibit 1104.0007
`
`

`

`84
`
`TABLE 1
`
`Unimpeded distances from the electrode surface
`
`
`
` Duration of experiment/s Unimpeded distance/mm
`
`0.06
`0.1
`0.2
`1.0
`0.6
`10.0
`
`100.0 2.0
`
`discussed in Sect. 4, in which there are two working electrodes. In those
`cases,
`it is important that the two working electrodes be parallel and
`separated by a rather narrow gap.
`
`1.4 FARADAIC AND NON-FARADAIC CURRENTS
`
`Two distinctly different types of current may flow at the electrode
`surface. One kind includes those in which electrons are transferred
`across the interface between the electrode and the solution. This electron
`
`transfer causes a chemical reaction, either oxidation or reduction (Sect.
`1.1)
`to occur. These reactions are governed by Faraday’s law, which
`asserts the proportionality of the electron-transfer rate to the current
`[eqn. (2)] . Accordingly, this current is called a faradaic current.
`A second type of current arises due to the presence of the electro-
`chemical double layer (Sect. 1.2). Additionally, a current may flow due
`to the adsorption or desorption (Sect. 1.2) or species 0 and R as well as
`electroinactive species. In these instances, no chemical reaction occurs
`and consequently electrons are not transferred across the electrode—
`solution interface. However, a current may flow elsewhere and this
`current is called a non-faradaic current.
`Both faradaic and non-faradaic currents may flow when an electrode
`reaction occurs. Thus, the total current which flows is often the sum of
`the faradaic and non-faradaic contributions to the current. Most often,
`it is the faradaic current that is of interest. Many electrochemical tech-
`niques have been developed which minimize or eliminate this non-faradaic
`contribution to the current, but discussion of these is beyond the scope
`of the present chapter.
`In Sect. 1.2, we chose to ignore the effects of the double layer as well
`as complexities due to adsorption and desorption processes. Similarly,
`we will also choose to ignore the presence of non-faradaic currents,
`though in practice they may be important. Hence, throughout this chapter
`only faradaic currents will be considered.
`
`1. 5 KINETIC STRATEGY
`
`Electrochemical studies are performed for many reasons, but in this
`chapter our preoccupation is with experiments carried out for the purpose
`
`InnoPharma Exhibit 1104.0008
`
`

`

`
`surFace
`
`
`
`
`
`F1
`electric
`
`current
`UXQS
`
`
`initial
`conditions
`
`kinetic relations
`
`trons ort relations
`P
`
`85
`
`surFoce
`
`electric
`Potential
`
`concentrations
`O; 0 and R
`
`(boundary
`
`,
`
`gametrlc
`5
`conditions
`
`condition)
`
`Fig. 4. Linkage between kinetics and mass transport. Quantities enclosed by a single
`frame are (generally) functions of time only, while those enclosed by a double frame
`are functions of spatial variables as well as time.
`
`serve, Current “\53231\\$fl9
` predict
`
`
`
`
`current
`
`
`
`
`otential
`
`JM 032
`
`
`
`P
`
`
`
`
`ob
`
`Experiment
`
`initial and
`
`geometric
`conditions
`
`
`interrelationship 0F
`surFace concentrations
`
`
`
`and surFace Fluxes
`surFace Fluxes
`
`
`
`concentration
`
`
`transport
`
`and
`Volmerion
`problem
`
`kinetics
`
`
`
`Flux profiles
`
`
`Fig. 5. Indirect method for determining electrode kinetics. In this method, a particular
`kinetic law has been assumed.
`
`of determining the kinetics of the electrode reaction. At first sight, it is
`not evident how mass transport,
`i.e.
`the motion of the electroactive
`species 0 and R through the solution,
`interacts mathematically with
`the elctrode kinetics, which manifests itself only at the interface between
`the solution 'and- the electrode. The interaction is not direct but occurs
`
`via the surface concentrations (see Sect. 1.2) and surface fluxes (see
`Sect. 2.2), as illustrated in Fig. 4.
`The only variables that are generally accessible in the scheme portrayed
`in Fig. 4 are the current and the potential; in a typical electrochemical
`experiment, one of these variables is imposed on the cell and the other is
`observed. How,
`then, does one learn anything about the kinetics by
`making use of the known laws of transport? There are two strategies for
`doing this.
`The most common strategy is illustrated in Fig. 5. A potential (constant
`or varying) is imposed on the cell and the current—time relationship is
`monitored.
`In the theoretical segment of the study, one assumes a
`
`References pp. 1 41—1 43
`
`InnoPharma Exhibit 1104.0009
`
`

`

`86
`
`
`
`observe
`
`current
`
`impose
`.
`wak%9
`potential
`
`
`Experiment
`
`initial and
`current
`
`
`
`geometric
`conditions
`
`
`
`
`interrelationship 0F
`PFEdIct
` surFace Fluxes
`SUFFOCQ concentrations
`
`
`
`concentration
`
`and
`
`
`Flux proFiles
`
`
`Nernst
`
`
`equation
`
`
`
`transport
`problem
`
`
`
`Fig. 6. Indirect method in which the electrode reaction is assumed to be reversible.
`
`
`
`1 MPOSQ
`
`otentiol
`Experiment
`
`
`
`
`seek kinetic low
`
`
`observe
`
`
`current
`
`semianQgrate
`surface
`concentrations
`
`
`
`Fig. 7. Direct method for determining electrode kinetics. Contrary to the indirect
`method portrayed in Fig. 5, no particular kinetic law has been assumed.
`
`particular kinetic law (usually volmerian kinetics, discussed in Sect. 3.5)
`by means of which an equation linking current, surface concentrations,
`and time is evolved. With the aid of this equation, the transport problem
`can be solved, generating expressions for the concentration and flux
`profiles of O and R. Then, the fluxes of O and R at the electrode surface,
`and hence the current, can be predicted. Finally, comparison of the
`measured current with the predicted current
`is used to verify the
`adequacy of the kinetic assumption and to provide values of the kinetic
`parameters.
`Commonly, the electrode reaction is assumed to be “reversible”, which
`simplifies the mathematics considerably because a direct prediction of
`surface concentrations is possible from the potential alone, as in Fig. 6.
`However, kinetic information is entirely lacking from experiments
`conducted under reversible conditions since the electrode reaction is then
`
`an equilibrium.
`The second strategy which may be used to learn about the kinetics of
`an electrode reaction is illustrated in Fig. 7. As before, a potential
`(constant or varying) is imposed on the cell and a current—time relationship
`is monitored. However, instead of assuming a particular kinetic law, one
`processes the experimental current by semi-integration (see Sects. 5.2 and
`5.4), thus enabling the surface concentrations to be calculated directly.
`Hence, the kinetics can be elucidated by a study that involves only the
`
`InnoPharma Exhibit 1104.0010
`
`

`

`87
`
`electrical variables: potential and current, together with the semi-integral
`of the current. This is the direct method of elucidating electrode kinetics
`and is discussed in Sects. 3.5 and 5.5.
`
`1.6 SYMBOLS AND UNITS
`
`The symbols listed below are used throughout this chapter. In most
`cases, the usage follows the recommendations of the IUPAC Commission
`on Electrochemistry [8] ; however, there are exceptions.
`
`electrode area (m2)
`constant of the j th component (Table 6)
`constant (Table 6)
`capacitance (F)
`capacitance at the input of an operational amplifier (F)
`capacitance of capacitors j and j — 1 (F)
`concentration (mol m'3)
`concentration of electroactive species at the cathode
`surface (mol m'3)
`bulk concentration of the jth ionic species (mol m‘3)
`total initial bulk ionic concentration (mol m"3 )
`bulk concentrations of species 0 and R (mol m"3)
`concentration of species 0 and R at the electrode surface
`(mol m'3)
`'
`concentration at distance x and time t (mol m‘3)
`total
`ionic concentration at distance x and time t
`
`(mol m'3)
`con3centration of species j at distance x and time t(mol
`m" )
`concentration of species 0 and R at distance x and time
`t (mol m‘3)
`Laplace transform of co (x, t) and cR (x, t) (mol m"3 s“)
`steady-state total
`ionic concentration at distance x
`(mol m'3)
`steady-state concentration of species j at distance x
`(mol m'3)
`surface concentrations of species 0 and R at time t (mol
`m'3)
`diffusion coefficient (m2 s' 1)
`diffusion coefficient of species j (m2 s“)
`diffusion coefficient of species 0 and R (m2 s“)
`electrode potential (vs. some reference electrode) (V)
`null or equilibrium potential (V)
`standard potential (V)
`interelectrode potential (V)
`voltage input and output of a circuit (V)
`
`i
`
`cJ-(x, t)
`
`00 (x, t), CK (x, t)
`
`50 (x, s), 51105: s)
`0105)
`
`0,-(x)
`
`080), 0130)
`
`D D
`
`i
`DO ,DR
`
`E E
`
`n
`
`Eo
`
`Ea —Ec
`Ein:Eout
`
`References pp. 141—1 43
`
`InnoPharma Exhibit 1104.0011
`
`

`

`oo 00
`
`electron
`
`Faraday’s constant (96 485 C mol‘l)
`denotes a functional relationship
`constant given by eqn. (161)
`faradaic current (A)
`jth and (i + 1)th faradaic current values (A)
`limiting faradaic current (A)
`faradaic current during the forward and reverse branches
`of a cyclic voltammogram (A)
`exchange current (A)
`transport-free current (Sect. 5.5) (A)
`faradaic current as a function of time (A)
`flux (mol In"2 S”)
`flux of speciesj (mol m‘2 s”)
`flux at distance x at time t (mol In‘2 5—1)
`flux of species j at distance x at time‘ t (mol m‘2 S“)
`flurges of species 0 and R at distance x at time t (mol
`m"
`s‘ )
`steady-state flux of species j (mol In"2 5—1)
`counter or index,j = 1,2,3, .
`. .N
`standard heterogeneous rate constant (m 5‘1)
`heterogeneous rate constants for forward and backward
`reactions (m 5‘1)
`distance between two parallel electrodes (m)
`Laplace transformation symbol
`molecular weight (kg mol‘ 1)
`metal
`metal ion
`
`1/2)
`semi-integral of the faradaic current (A s
`faradaic semi-integral during the forward and reverse
`branches of a cyclic voltammogram (A 51/2)
`total number of ionic species (Sect. 4.0); upper limit
`of counter (Sect. 5.4)
`Avogadro’s constant (6.02205 x 1023 mol'l)
`number of faradays to reduCe one mole of the reducible
`species
`oxidized species
`modified oxidized species at the electrode surface
`arbitrary function in Laplace space [eqn. (142)]
`arbitrary function in Laplace space
`charge (C)
`unit charge (C)
`gas constant (8.31441 J mol'1 K'l)
`reduced species
`resistance (ohm)
`
`InnoPharma Exhibit 1104.0012
`
`‘3'c, + H
`
`
`
`
`
`r-57¢:-H-OnH:'11m.n
`
`l
`
`«a?
`
`1-0
`i*
`
`m)
`
`J J
`
`].
`J(x, t)
`«II-(x, t)
`J0 (x: t)! JR (x9
`
`Jj(x)
`1'
`k0
`
`kf,kb
`
`L Y
`
`

`

`89
`
`input resistance to an operational amplifier (ohm)
`jth and (j —- 1)th resistances in a circuit (ohm)
`radius (m)
`modified reduced species at the electrode surface
`denotes “solid”, as in M(s)
`dummy variable in Laplace space (s’l)
`thermodynamic temperature (K)
`time interval in a cyclic experiment (5)
`time (5)
`lower time limit of semi-integration (s)
`mobility of an ionic species (m2 s'1 V“)
`mobility of a particular ionic species j (m2 5‘1 V'l)
`average velocity (m s”)
`average velocity at time t (m s”)
`local electric field (V m'l)
`distance measured normal to the electrode surface (m)
`charge number
`charge number of a particular ionic species j
`transfer coefficient
`
`surface excess concentration (mol m_2)
`surface excess concentrations of species 0 and R (mol
`m" )
`~
`activity coefficient
`constants given by eqns. (75) and (94)
`time interval (5)
`energy required to carry an ion from distance x = 0 to
`x = x (J)
`viscosity (kg m"1 s")
`overpotential (V)
`equivalent
`ionic conductivity of an ion (m2 ohm“l
`equiv. '1)
`order of a derivative (Table 6)
`electrochemical potential (J mol‘l)
`standard electrochemical potential (J mol'l)
`local ionic strength at distance x and time t (mol m'3)
`local ionic strength at steady state (mol m‘3)
`order of a derivative (Table 6)
`stoichiometric coefficient of species j
`support ratio
`integration variable (5)
`time of sudden change in imposed conditions after an
`experiment commences;
`transition time in a chrono-
`potentiometric experiment (s)
`local potential (V)
`reference potential (V)
`
`a. v
`.5. :U
`
`-.
`
`I
`
`p.-
`
`Suqqmmwwww
`
`GQ! A N V
`
`figiflmax
`
`"1O "1 pa
`
`7
`7, 71
`A
`do —> x)
`
`V33
`
`)1
`fl
`u
`u(x, t)
`M(x)
`V
`V,
`,0
`
`T 7
`
`.
`
`as
`
`References pp. 141—143
`
`InnoPharma Exhibit 1104.0013
`
`

`

`90
`
`¢(x,t)
`¢(x)
`
`local potential at a distance x and time t (V)
`local potential in the steady state (V)
`
`2. Modes of transport
`
`There are three distinct mechanisms by which electroactive solutes
`from the solution may reach the electrode or, conversely, by which
`electrogenerated solutes may be transferred into the solution. These are
`migration, diffusion, and convection.
`Migration is perhaps the easiest to understand. The presence of an
`electric field causes charged particles to move along the field lines; this is
`migration. The force experienced by the particle is proportional to its
`charge and to the electric field (i.e.
`the electrical potential gradient).
`Migration affects ions only; neutral species do not migrate.
`Diffusion does not occur in response to any physical force; it occurs
`because an inhomogeneous solution seeks to maximize its entropy.
`Explained another way, the Brownian motion that all solute particles
`undergo inevitably tends to enrich regions of low local concentration at
`the expense of neighbouring regions of greater concentration. Diffusion
`affects charged and uncharged species equally; both ions and molecules
`diffuse.
`
`the solute particle moves
`In transport by migration or diffusion,
`through a stationary solvent. Convection is a totally different process in
`which the solution as a whole is transported. Solute species reach or leave
`the vicinity of the electrode by being entrained in a moving solution.
`A concise summary of the distinctions between migration, diffusion,
`and convection is that migration occurs in response to a potential gradient,
`diffusion in response to a concentration gradient, and convection in
`response to a pressure gradient.
`
`2.1 CONVECTIVE TRANSPORT
`
`We may distinguish two kinds of convection [15, 16] : forced convection
`and natural convection. Forced convection is the result of some motion
`
`deliberately introduced by the experimenter; natural convection arises
`from changes brought about as a result of the electrolysis itself.
`Stirring the solution,
`rotating the electrode, bubbling gases, and
`pumping solution towards or across the electrode are all methods of
`inducing forced convection. Such methods yield irreproducible results
`unless careful attention is paid to the geometry of the system and to
`ensuring uniformity of the imparted motion. Rotating disks and wall-
`jet devices are among the most reproducible examples of forced con-
`vection . These are the subject of Chap. 5 and will not be further discussed
`here.
`
`InnoPharma Exhibit 1104.0014
`
`

`

`91
`
`The most usual sequence of events leading to natural convection starts
`when electrolysis generates a region close to the electrode in which a
`concentration is significantly enhanced or depleted. Usually, a solution
`that is enriched in a solute has a slightly greater density than the bulk
`solution and therefore, under the influence of gravity, there is a tendency
`for enriched regions to fall. Conversely, a region of depleted concentration
`tends to rise as a result of its slightly diminished density.
`The extent to which a solution is able to respond to these gravitational
`tendencies depends on the geometry of the cell (see Sect. 4). Even under
`the worst conditions, however, the driving force for natural convection
`is small and it takes a considerable amount of time for this small force
`
`to overcome the inertia of the solution mass. Accordingly, natural con-
`vection is unimportant in rapid experiments. In fact, its effect seldom
`appears before about 10 or 20 s, which is long in relation to most electro-
`chemical experiments.
`Henceforth in this chapter, convection will be ignored. In other words,
`we shall treat only experiments in which forced convection is absent and
`in which natural convection is inhibited either by judicious geometric
`design of the cell or by the brevity of the experiment.
`
`2.2 FLUX
`
`The number of moles of a solute being transported across unit area of
`a surface in unit time is known as the flux at that surface; it has units of
`mol m"2 S”. The flux, J, may be equated to the average velocity, v, of
`the individual solute molecules in the direction normal to the surface
`
`multiplied by their concentration, c
`
`J(x, t) = vc(x, t)
`
`(9)
`
`If the surface in question is the electrode and the solute in question is
`the electrogenerated species R,
`then the flux is equal
`to the rate of
`generation of R, i.e. the rate of the electrode reaction
`
`Rate of reaction = JR (0,t)
`
`(10)
`
`Similarly, the flux of the electroactive species 0 at the electrode surface
`is given by
`
`Rate of reaction = — Jo (0, t)
`
`(11)
`
`where the negative sign arises because of the convention that flux is
`treated as positive when it occurs in the direction of increasing x, i.e. away
`from the electrode.
`
`The third component involved in the electrode reaction
`
`O(soln) + n e —> R(soln)
`
`(12)
`
`is electrons. Their flux at the electrode surface must also be proportional
`to the rate of reaction in accordance with Faraday’s law and eqn. (2).
`
`References pp. 141—143
`
`InnoPharma Exhibit 1104.0015
`
`

`

`92
`
`These results together lead to the very important relationship
`
`i(t)
`nAF
`
`JR(0,D = "Jo(0,t) =
`
`(13)
`
`in which the sign of the current i(t) reflects our choice of cathodic current
`as positive.
`Expression (13) implies that, for every n electrons that are Withdrawn
`from the cathode, one molecule (or ion) of R is transported away from
`the electrode. This may be untrue transiently if there is significant
`adsorption of R. More precisely, expression (13) will be invalid whenever
`the amounts of adsorption PR and [‘0 of R and 0, respectively, (see
`Sect. 1.2) are changing significantly with time. Correcting for this effect
`leads to
`
`J0t+d—F—Rt———JOt——£9t—‘
`
`
`i(t)
`nAF
`
`14
`
`the
`it will henceforth be assumed that
`Unless we state otherwise,
`derivatives in expression (14) are negligible, so that expression (13) can
`be used.
`
`2.3 LAWS OF MIGRATION
`
`Consider a region of solution in which the local potential ¢ changes
`along the x axis. Then, — a¢/ax denotes the local electric field X. This
`field acts upon any charge q to produce a force qX acting along the x
`axis. If the charge is that on an ion of charge number z, so that q =
`zF/NA , where NA is Avogadro’s constant, then
`
`zFX
`
`Ah
`
`(15)
`
`_
`Electrostatlc force =
`
`If the ion is in a fluid medium, then the electrostatic force seeking to
`accelerate the ion is opposed by a viscous force trying to slow the ion
`down. Though it
`strictly applies only to spheres of macroscopic
`dimensions, Stoke’s law can provide an approximate expression
`
`Viscous force = 67rnrv(t)
`
`(16)
`
`for the viscous drag, n being the coefficient of viscosity, r the ion’s radius
`and v(t) its velocity an instant t.
`If M/NA is the mass of the ion, then from Newton’s second law of
`motion
`
`M dv
`
`NA dt
`
`= electrostatic force — viscous force =
`
`
`zFX
`
`NA
`
`— 67rnrv(t)
`
`(17)
`
`InnoPharma Exhibit 1104.0016
`
`

`

`
`
`velomty v
`
` limiting velocity
`
` timet
`
` 6—)
`time
`constant
`
`Fig. 8. Velocity versus time curve for an ion in solution.
`
`On integration, one finds
`
`t
`1’0
`
`zFX
`= -——-——-—
`61rNA'qr
`
`1
`
`-—e
`
`[~61rNAnrt
`——-——.——.—
`M
`
`Xp
`
`(18
`
`)
`
`if the ion was initially stationary. This equation corresponds to an ion
`experiencing an initial acceleration of zFX/M but settling down to a
`steady limiting velocity of zFX/(67rNAnr) after times long in comparison
`with the time constant (M/61rNAnr) as shown in Fig. 8.
`For common inorganic ions in aqueous solution, one can take the
`typical values M 2 0.1 kgmol”, 17 1 1.0 x 10'3 kgm"ls"1, z = i (1 or 2),
`r2’0.3 nm and calculate a time constant of the order of 3x 10'14 s.
`
`Evidently, the limiting velocity is acquired virtually instantaneously, so
`that the adjective “limiting” may be dropped and v(t) contracted to v.
`Using the same typical values gives the (velocity/ field) ratio as
`
`(19)
`
`2 F
`= l—'— R 4x10‘3m2V'1s‘l
`67rNA77r
`
`
`
`s X
`
`
`
`This quantity is known as the mobility of the ion and is given the symbol
`u. Some experimental values are given in Table 2 for ions in dilute aqueous
`solution.
`
`Mobility is accorded a positive sign irrespective of the sign of the ion’s
`charge, though of course the direction of motion does depend on the sign,
`cations travelling with the field X (i.e. down the potential gradient a¢/ 8x)
`and anions in the opposite direction. These signs are taken into account
`in the equation
`
`qu _
`121
`
`an EMS
`~ —— ——
`lzl ax
`
`v =
`
`(20)
`
`for the velocity of migration. Util

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