throbber
Laboratory
`Techniques in
`Electroa na lyt ica I
`Chemistry
`
`Second Edition, Revised and Expanded
`
`edited by
`
`Peter T, Kissinger
`Purdue University and
`Bioanalyticai Systems, Inc.
`West Lafayette, Indiana
`
`William R, Heineman
`University of Cincinnati
`Cincinnati, Ohio
`
`Marcel Dekker, Inc.
`
`New York* Basel ¯ Hong Kong
`
`Lupin Ex, 1054 (Page 1 of 19)
`
`

`

`Library of Congress Cataloging-in-Publication Data
`
`Laboratory techniques in electroanalytical chemistry / edited by Peter
`T. Kissinger, William R. Heineman. -- 2nd ed., rev. and expanded.
`p. cm.
`Includes bibliographical references and index.
`ISBN 0-8247-9445-1 (hardcover : alk. paper)
`1. Electrochemical analysis--Laboratory manuals. I. Kissinger,
`Peter T. II. I-Ieineman, William R.
`QD115.L23 1996
`543’.087 de20
`
`95-46373
`CIP
`
`The publisher offers discounts on this book when ordered in bulk quantities.
`For more information, write to Special Sales/Professional Marketing at the
`address below.
`
`This book is printed on acid-free paper.
`
`Copyright © 1996 by MARCEL DEKKER, INC. All Rights Reserved.
`
`Neither this book nor any part may be reproduced or transmitted in any form
`or by any means, electronic or mechanical, including photocopying, micro-
`filming, and recording, or by any information storage and retrieval system,
`without permission in writing from the publisher.
`
`MARCEL DEKKER, INC.
`270 Madison Avenue, .New York, New York 10016
`
`Current printing (last digit)’
`10 9 8 7 6 5 4 3 ’2 1
`
`PRINTED IN THE UNITED STATES OF AMERICA
`
`Lupin Ex. 1054 (Page 2 of 19)
`
`

`

`This materiai may be protected by Copyright faw (Titie 17 U.S. Code)
`
`15
`Solvents and Supporting Electrolytes
`
`Albert J. Fr~ Wesleyan University, Middletown, Connecticut
`
`!.
`
`INTRODUCTION
`
`Almost every electrochemical experiment is carried out in a medium consist-
`ing of a solvent and supporting electrolyte. Many such combinations have been
`used by electrochemical experimenters; as a glance at the Appendix will show,
`selection of a solvent and electrolyte has been the subject of much discussion
`in the literature. Yet the problem, is not intrinsically complicated; a handful of
`solvent-electrolyte combinations would probably suffice for the majority of
`electrochemical applications. For this reason, a few particularly good solvent
`systems (the term "solvent system" is used here to describe the medium con-
`sisting of both solvent and supporting electrolyte) are recommended here, in lieu
`of an extended but uncritical listing of all such systems. The systems to be
`recommended have been chosen from among many that have been reported.
`However, because the novice, will certainly want to develop a feel for what
`makes a "good" solvent system, the first part of the chapter discusses desirable
`criteria, as well as the ways in which electrochemical results may depend on
`the nature of the particular solvent-electrolyte combination employed. Because
`special situations frequently.do require special solutions, Section III describes
`some less common solvents and electrolytes that have been found useful in
`special situations. Finally, the Appendix offers a guide to further readings of a
`more comprehensive nature.
`
`A. Solvent Selection Criteria
`
`There is no universal solvent, and even for a given application one rarely finds
`an ideal system. One must factor some informed guesswork into one’s choice
`of solvent and electrolyte. In order to optimize conditions for an electrode re-
`action, one must consider how its chemical and electrochemical features, for
`
`469
`
`Lupin Ex. 1054 (Page 3 of 19)
`
`

`

`470
`
`Fry
`
`example, the chemical properties of expected intermediates, might be affected
`by the solvent or electrolyte. A solvent is chosen such that its merits outweigh
`its disadvantages in a particular situation. A good solvent system for one type
`of experiment may be wholly unsuitable for other applications. The most im-
`portant properties that the ideal solvent system ought to possess are electrochemi-
`cal inertness (stability over a wide range of potentials), high electrical conduc-
`tivity, good soIvent power, chemical inertness, availability in pure form or ease
`of purification, and low cost, although others must sometimes be considered.
`We now examine each of these in detail.
`
`Electrochemical Inertness
`The solvent system should not undergo any electrochemical reaction over a wide
`range of potentials from very positive (strongly oxidizing) to very negative
`(strongly reducing). The potential at which electrochemical reaction of the sol-
`vent system (either the supporting electrolyte or the solvent itself) commences
`is known variously as the solvent breakdown potential, solvei~t decomposition
`potential, or solvent background limit. Certain systems can be used over a very
`wide potential range; one of the best for this purpose is a solution of tetrabutyl-
`ammonium hexafluorophosphate in acetonitrile, which exhibits anodic and ca-
`thodic breakdown limits of +3.4 and -2.9 V (vs. SCE), respectively [1]. In
`practice, however, one rarely needs both extended positive and negative ranges
`in a single set of experiments; solvents are frequently used that exhibit either a
`very negative or very positive decomposition potential, but not both. Thus, both
`nitrobenzene [2] and methylene chloride [3] have been used for studies of oxi-
`dation processes because they are hard to oxidize, notwithstanding the fact that
`they are quite easily reduced. Conversely, liquid ammonia and methylamine are
`good solvents for electrochemical reductions, although poor for oxidations [4].
`It is reported that potentials as positive as +4.5 V (vs. SCE) can be reached in
`liquid SO2 [5].
`The potential limit is frequently set by the electrochemical behavior of the
`supporting electrolyte, not the solvent. For example, dipolar aprotic solvents
`such as acetonitrile (AN), dimethylformamide (DMF), and dimethyl sulfoxide
`(DMSO) are very difficult to reduce; the reduction of alkali metal ions to the
`corresponding metals (ca. -2 V vs. SCE) is the potential-limiting process in such
`soIvents, and by use of the more difficult-to-reduce tetraalkylammonium salts
`one can reach considerably more negative potentials. The decomposition poten-
`tials of tetraalkylammoniumions at mercury themselves become more negative
`as the alkyl groups become larger (Me4N+: -2.65 V; Bu4N+: -2.88 V) [6]. One
`can frequently extend the accessible cathodic range by using a large tetraalkyl-
`ammonium ion, but double-layer and adsorption effects can negate this gener-
`alization [6,7]. DMF and DMSO are not suitable as solvents for anodic stud-
`ies because they are fairly easily oxidized, but as noted earlier, AN is very
`difficult to oxidize, and in this solvent the decomposition potential of the sup-
`
`Lupin Ex. 1054 (Page 4 of 19)
`
`

`

`Solvents and Supporting Electrolytes
`
`471
`
`¯ porting electrolyte is usually the limiting process. At very positive potentials,
`oxidation of the anion of the supporting electrolyte is often the limiting process.
`Not only should easily oxidized ions such as bromide and iodide not be used
`for anodic studies, but even nitrate and perchlorate salts should be avoided when
`one wishe~ to operate at very positive potentials; at ihese potentials tetrafluoro-
`borate and hexafluoroborate salts are preferred because of their stability to
`oxidizing conditions [1].
`¯ Electrochemical reaction of the solvent or supporting electrolyte need not
`be undesirable. Sometimes one can take advantage of such behavior by using
`the product of such a reaction as a chemical reagent. For example, one can
`generate bromine by anodic oxidation of bromide ion or lithium by cathodic
`reduction of lithium ion. There can be real advantages to electrochemical gen-
`eration of reagents this way instead of using the bulk reagent, for example,
`avoiding the need to handle dangerous substances and being able to add the
`reagent in regulated amounts by careful control of the electrolysis current [8].
`
`Electrical Conductivity
`In order to support passage of an electric current, the solvent system should have
`low electrical resistance. This implies that the solvent should have a moderately
`high dielectric constant (_>10); the prevalence of ion pairing and even multiple
`ionic association in less polar ffolvents results in relatively low ionic mobility
`and conductance. The fact that nonpolar solvents tend not to be very good sol-
`vents for salts in the first place makes it even harder to contemplate using theml
`A number of the most common solvents for organic electrochemistry have sat-
`isfactory dielectric constants (DMF: 36.7; AN: 37.5; DMSO: 46.7), and their
`electrolytic solutions have acceptably high conductances [9-11]. Solvents of
`substantially lower polarity than these, however, have been used successfully
`despite their high electrical resistance (e.g., tetrahydrofuran, THF [12], and
`methylene chloride [3], whose dielectric constants are 7.3 and 8.9, respectively
`[13]). Lower conductivity can be tolerated more readily in low-current appli-
`cations such as voltammetry than in preparative electrolyses. Although volt-
`ammograms can be badly distorted by uncompensated resistance under condi-
`tions of high electrical resistance, the distortion can be largely corrected by using
`high electrolyte concentrations (not always possible in resistive solvents, which
`tend to be poor solvents for salts) and by computational correction of the volt-
`ammogram [11]. A new element has entered the scene in recent years with the
`development of small-diameter (diameter _<10 I.tm) microelectrodes [14]. These
`can be used in highly resistive media [11,15-19], including the absence of
`supporting¯ electrolyte [20], and in fact even in the gas phase [21].
`
`Good Solvent Power.
`An electrochemical solvent must be able to dissolve a wide range of substances
`at acceptable concentrations. In general, this means that electrolytes must be
`
`Lupin Ex. 1054 (Page 5 of 19)
`
`

`

`1
`
`soluble at least to the extent of 0.1 M, while the electroactive substance must
`be soluble to the extent of 0.1-1 mM for application of the various electroana-
`lytical techniques, and higher concentration for preparative electrolyses. Salts
`in which either the cation or anion, and preferably both, are large, such as
`tetraalkylammonium perchlorates, hexafluorophosphates, and tetrafluoroborates,
`generally exhibit the greatest solubility in organic solvents. The latter two are
`becoming more popular than perchlorates because of safety hazards with per-
`chlorates, especially in acidic media. Ethyltributylammonium tetrafluoroborate
`is quite soluble in nonpolar solvents [12], as is tetrabutylammonium hexafluoro-
`phosphate.
`
`Chemical Inertness
`The solvent should not react with the electroactive material nor with interme-
`diates or products of the electrode reaction under investigation. No one solvent
`can reasonably be expected to be unreactive toward all of the many kinds of
`reactive species that can be generated electrochemically. When selecting a sol-
`vent, one does have to make some informed guesses about likely intermediates
`and products, based on chemicaI experience. For example, many organic and
`anodicprocesses generate cationic, that is, electrophilic species [22]; one would
`therefore not use a nucleophilic solvent if one wants such intermediates to be
`long-lived. This is equally true for organic and inorganic species. Acetonitrile
`is frequently used in anodic studies because of its high conductivity and wide
`accessible potential range, but it is moderately nucleophilic; it can react with
`organic intermediates and replace ligands on inorganic intermediates. Attempts
`to characterize anodically gefierated intermediates are therefore frequently car-
`ried out in a less nucleophilic solvent such as methylene chloride or one of the
`"superacid" solvents to be discussed in Section II.A. Conversely, reduction of
`organic substrates frequently forms anionic intermediates, and one must there-
`fore avoid the use of electrophilic solvents, in particular protic substances, to
`avoid acid-base reactions between such intermediates and the solvent. Finally,
`one would like the solvent to be reasonably stable, so that purification, prepa-
`ration, and storage of standard solutions will present no major problems.
`
`Liquid Range
`The solvent should exhibit a convenient liquid range; that is, it should be nei-
`ther too low-boiling nor too high-boiling. The former makes for potential safety
`and fire hazards and.difficulty in storing, solutions, particularly when degassing
`them, and the latter makes purification by distillation more difficult. On the other
`hand, the definition of convenient liquid range depends on the particular experi-
`ment involved, in particular the temperature at which it is to be carried out.
`High-boiling solvents are preferred, for experiments at higher temperatures,
`whereas low-boiling solvents are better for low-temperature electrochemistry
`
`Lupin Ex. 1054 (Page 6 of 19)
`
`

`

`Solvents and Supporting Electrolytes
`
`473
`
`(they retain favorable viscosity properties at low temperatures). Other factors
`to be considered are (a) whether (and how) the solvent must be removed and
`solutes recovered at the end of the experiment, and (b) the particular purifica-
`tion and transfer procedures to be used in the experiment..
`
`Miscellaneous Considerations
`Solvents and electrolytes should also be inexpensive, nontoxic, and nonflam-
`mable. The latter two characteristics are not well satisfied by most organic
`solvents, but with reasonable safety precautions and reasonable ventilation they
`can be used routinely without incident. Another solvent property, viscosity, may
`be of importance on occasion. High viscosities are useful when one wishes to
`extend the time interval over which mass transport occurs purely by diffusion,
`such as for potential-step experiments, but a low-viscosity solvent is preferred
`when efficient mass transport is required, as in preparative electrolyses.
`
`!1. RECOMMENDED SOLVENTS AND ELECTROLYTES
`
`A. Anodic Electrochemistry
`
`Solutions of inorganic salts in water meet many of the preceding desirable sol-
`vent criteria, and have been used for innumerable electrochemical studies. Much
`inorganic electrochemistry has been carried out in water with little difficulty.
`Organic and organometallic compounds either are not very soluble in water or
`are reactive toward it; this has stimulated interest in the use of nonaqueous
`solvents in electrochemistry. Ethanol and methanol are good organic solvents,
`inexpensive, readily available in high purity, dissolve salts readily, and in gen-
`eral have much to commend them for electrochemical purposes. On the other
`hand, alcohols are relatively easily oxidized and are highly nucleophilic. For
`this reason, aprotic organic solvents are more commonly used, and of these, the
`most commonly used are acetonitrile and N,N-dimethylformamide. Acetonitrile
`(AN) is probably the most widely used solvent for organic anodic electrochem-
`istry, and we believe that it is the preferred solvent for most such purposes. It
`is a good solvent for organic compounds and many inorganic and organic salts,
`has a high dielectric constant and a convenient liquid range (mp -45.7°C, bp
`81.6°C), and is transparent in the ultraviolet region. (For purification of AN,
`see Sec. VI.) Recommended electrolytes include tetrabutylammonium hexa-
`fluorophosphate (TBAHFP) and tetrafluoroborate (TBATFB), both of which are
`readily obtained in pure form (See. VI), have high solubilities in AN, and are
`stable to very positive potentials. These salts, particularly the hexafluoro-
`phosphate, are rather insoluble in water and can often be separated from reac-
`tion products by evaporation of the AN followed by extraction with a nonpolar
`solvent, chromatography, or volatilization of the desired product. Sometimes
`
`Lupin Ex. 1054 (Page 7 of 19)
`
`

`

`474
`
`Fry
`
`separation of the electrolyte from an organic product presents difficulties, how-
`ever. An inorganic electrolyte (e.g., NaBF,~, LiC104, or NaC104) may be nec-
`essary in such cases.
`Acetonitrile is rather nueleophilic; this is a disadvantage when the electro-
`chemical reaction produces a reactive electrophilic intermediate, or when the
`substrate is an inorganic or organometallic species containing a replaceable
`ligand. In this case, solutions of TBATFB or TBAHFP in methylene chloride
`have frequently been used as alternatives, although the latter solutions are re-
`ported to decompose slowly [23]. Meyer, however, has recently found that 1,2-
`difluorobenzene is a better solvent than methylene chloride [24]. Other choices,
`although chemically more reactive than the preceding halogenated solvents,
`include trifluoroacetic acid (no electrolyte needed) [25], liquid sulfur dioxide
`[26], an aluminum chloride-alkali halide melt [27], or trifluoromethanesulfonic
`acid [28]. Anodically generated reactive electrophilic inter.mediates are also rela-
`tively long-lived at low temperatures [29] or in very dry AN [30,31].
`
`B. Cathodic Electrochemistry
`
`Water and alcohols are often useful in inorganic electrochemistry. Dimethyl-
`formamide (DMF), acetonitrile, and dimethyl sulfoxide (DMSO) are all excel-
`lent solvents for reductive electrochemistry of organic compounds. TBAHFP and
`TBATFB are good electrolytes for use in these solvents. Despite their rather high
`boiling points (156°C and 189°C, respectively), DMF and DMSO are particu-
`larly good solvents for cathodic electrochemistry. DMF is fairly easily purified
`(See. VI.B) and DMSO is in fact fairly pure as,. received. Both are excellent
`solvents for both organic and inorganic solvents. Water (which is always present
`in polar aprotic solvents) is a poor proton donor in DMSO and DMF and can
`be reduced by appropriate procedures (Sec. VI.A). Probably the most widely
`used organic solvents for cathodic electrochemistry are DMF and acetonitrile.
`DMF has a number of desirable properties, but when wet is prone to slow de-
`composition, liberating dimethylamine. Decomposition is much less of a prob-
`lem in dry DMF, which can be prepared by vacuum distillation from calcium
`hydride, followed by treatment with 3-/~ molecular sieve [32]. Acetonitrile would
`be our primary recommended solvent for cathodic electrochemistry except for
`problems related to its purification. Very pure, dry AN can be obtained by
`proper treatment, but this can be tedious [33]. The efficacy of a number of
`simpler procedures in the literature for. drying AN has been disPUted [34].
`
`I!!. SOME OTHER SOLVENTS
`
`Many solvents and electrolytes have been used in electrochemical applications.
`In the preceding sections we suggested a few solvents for general use. It may
`
`Lupin Ex. 1054 (Page 8 of 19)
`
`

`

`Solvents and Supporting Electr.olytes
`
`475
`
`. happen, however, that none of these is acceptable for a particular application.
`There are many exhaustive discussions of solvent systems available in the lit-
`erature, which we shall not duplicate here (see the Appendix). A few solvents
`deserve specia! mention here, because each has a property likely to be useful
`in certain special situations.
`
`A. Low-Temperature Electrochemistry
`
`Low-temperature applications (Chap. 16) present special problems. With most
`solvents, solvating power decreases (this is a particular problem with support-
`ing electrolytes, which have to be soluble in concentrations of 0.1 M or higher
`for conventional electrochemistry) and viscosity increases as the temperature is
`lowered. Both effects result in higher solution resistance at low temperatures.
`Reilley and Van Duyne [29] and, more recently, Evans [11] and Murray [15]
`and their coworkers have examined a variety of solvent combinations and have
`found several that suffer the least degradation of solvent properties with decreas-
`ing temperature. Reilley and Van Duyne recommended butyronitrile for use with
`cortventional electrodes at temperatures as low as -105°C. The development of
`microelectrodes in recent years has been a boon to those interested in low-tem-
`perature electrochemistry, since these electrodes can function well in highly
`resistive media (Chap. 12). Evans recommended acetonitrile for operations down
`to -40°C and butyronitrile to -80°C for high-speed low-temperature voit-
`ammetry at microelectrodes. Murray found a number of solvent mixtures that
`are suitable for voltammetry at ultra-low temperatures: 1:1 butyronitrile
`(BN):ethyl bromide (with 0.2 M TBAP as electrolyte) is usable to -148°C;
`1:1:1:1 isopentane: methylcyclopentane:BN:ethyl bromide to-158°C; and re-
`markably, 1:1 BN:ethyl chloride, in which voltammetric data could be obtained
`at temperatures as low as -185°C! The use of microelectrodes also permits
`voltammetric measurements in frozen solvent [16] and in solids [17]. Gosser has
`carried out voltammetry in frozen DMSO with conventional electrodes, but has
`suggested that the electrochemistry actually occurs in pools of liquid solvent
`within the frozen medium [18].
`
`B. Solvated Electrons
`
`Solvated electrons are readily prepared in hexamethylphosphoramide (HMPA),
`ammonia, and methylamine [19]. Lithium chloride is the preferred electrolyte
`for this application.
`
`C. Nonpolar Media
`
`Nonpolar media are of interest as electrochemical solvents because they permit
`the investigator to observe the properties of polar intermediates undistorted by
`
`Lupin Ex. 1054 (Page 9 of 19)
`
`

`

`476
`
`Fry
`
`interaction with the solvent medium, but they present major problems because
`of their high electrical resistance. This is compounded by the poor solubility of
`most electrolytes in such solvents. The high electrical resistance causes severe
`signal distortion in low-current experiments (voltammetry) and resistive heat-
`ing (which can be so severe as to cause solvent boiling) in high-current experi-
`ments (preparative-scale electrolyses). As mentioned in the preceding section,
`the use of microelectrodes can largely mitigate the resistance problem in volt-
`ammetry in such media.
`There are hydrocarbon-like media that can permit voltammetric studies. It
`turns out that molten naphthalene and its 1-chloro and 1-methyl derivatives are
`highly electrically conductive and good solvents for organic salts at elevated
`temperatures (1.0 M TBAP solutions can be prepared) [35]. Well-formed ac and
`cyclic voltammograms can be obtained at 150°C. These media were found useful
`for substrates such as phthalocyanines, which are completely insoluble in all
`common solvents near room temperature. A barely explored idea in nonaqueous
`electrochemistry is that of incorporating the functions of solvent and support-
`ing electrolyte into a single substance, that is, a liquid or low-melting organic
`salt. A solution of tetrabutylammonium tetrafluoroborate (TBATFB) in toluene
`separates into two liquid phases; the upper is pure toluene, and the lower is a
`novel "liquid electrolyte" whose composition corresponds to three molecules of
`toluene and one of TBATFB [36]. Fry and Touster examined the properties of
`this substance as a voltammetric solvent and found them to be remarkably similar
`to those of DMF, AN, or DMSO [37]. Tetrahexylammonium benzoate, a liq-
`uid at room temperature, has been found suitable for electrochemical use [38].
`Better candidates might be found among the many low-melting organic salts now
`available [39].
`
`IV. SOLVENT- AND ELECTROLYTE-DEPENDENT PHENOMENA
`
`Electrochemical reactions are unavoidably influenced by the nature of the sol-
`vent and supporting electrolyte. We have already alluded several times to the
`possibility of chemical reaction of electrochemically generated intermediates with
`the solvent. One must always be on guard against this possibility. The nature
`of the problem is exemplified by a study of the electrochemical reduction of
`iodobenzene in DMF containing tetraethylammonium bromide [40]. A variety
`of experiments demonstrated that the h.ighly basic and nucleophilic phenyl an-
`ion is generated in this reaction. Benzene is the major product; presumably it
`is formed by attack by the phenyl anion on traces of moisture in the solvent.
`When the water content in the solvent was lowered, benzene was still found as
`the major product, but now it was found to be formed by proton attack upon
`the cation of the supporting electrolyte (Hoffmann elimination). When the elec-
`trolyte was changed to lithium bromide to prevent Hoffmann elimination, phe-
`
`Lupin Ex. 1054 (Page 10 of 19)
`
`

`

`Solvents nnd Supporting Elect~rol~tes
`
`4 ~7 .
`
`nyl anion proceeded to react with the solvent by a combination of proton ab-
`straction and nucleophilic addition! This is a reminder that carbanions gener-
`ated in dipolar aprotic solvents such as DMF are especially reactive; they can-
`not bind covalently to tetraalkylammonium counte.rions, .and even ion pairing
`to the coUnterion is minimal in such solvents because of their high dielectric con-
`stants. Electrochemically generated phenyl anions might perhaps be longer lived
`in a less reactive solvent (e.g., THF) with a metal salt as electrolyte. The an-
`odic counterpart of this problem has already been mentioned: the fact that an-
`odically generated cations readily attack nucleophilic solvents such as
`acetonitrile, mandating the use of electrophilic solvents when it is desired to
`extend the lifetime of such cations. All of the special solvent systems mentioned
`at the end of Section ii meet this criterion.
`The solvent and electrolyte can affect an electrode process in a number of
`ways other than by reacting chemically with the electrolysis intermediates or
`products. The rate of heterogeneous electron transfer from the electrode to the
`electroactive substance can be affected by. the structure of the electrical double
`layer at the electrode surface, which in turn is dependent on the nature of the
`solvent and supporting electrolyte [41]. This isn’t the only role the electrolyte
`can play when incorporated into the double layer. Acidic cations, such as tri-
`ethylammonium and guanidinium ions, can protonate short-lived intermediates
`generated at the electrode surface before such intermediates can escape into bulk
`solution [42]. Similarly, preferential solvation of small ions by one constituent
`of a mixture can raise the local concentration of this substituent at the electrode
`surface if the ion is in the double layer (anions at potentials positive of the
`potential of zero charge [PZC], cations at potentials negative of PZC). Examples
`of both cathodic [42] and anodic [43] electrochemistry thought to be affected
`this way have been reported. A third type-of surface electrolyte effect has been
`reported, in which optically active products have been obtained from electrode
`reactions in which an optically .active supporting electrolyte has been used [44].
`The degree of asymmetric induction obtained in this fashion has generally been
`small. Reasons for this have been suggested [45].
`The most eornmon electrochemical effects exerted in bulk solution are
`related to association (solvation, ion-pairing, complex formation, etc.) with the
`electroactive substance or electrochemically generated intermediates [4,19]. The
`importance of solvation can be gauged by comparing calculated and measured
`values of the parameter !XEI~2 (defined as the difference, in volts between the
`half-wave potentials of the first and second polarographic Waves) exhibited by
`polycyclic aromatic hydrocarbons (PAH) in dipolar aprotic solvents [46,47]. It
`can be shown that AEIt2 is related to the equilibrium constant for disproportion-
`ation of the aromatic radical anion into neutral species and dianion, that is,
`
`2Ar:~ArH + ArI-I2-
`
`Lupin Ex. 1054 (Page 11 of 19)
`
`

`

`478 F~
`
`The computed value of AE~/2 by a reliable computational method [48] is -5 V
`[49]. The experimental values are generally only about one-tenth of this (0.4-0.6
`V) in DMF, AN, or DMSO, showing that the disproportionation equilibrium
`is significantly perturbed by solvation [46,47,49]; the computed value is essen-
`tially a gas-phase, solvent-free value. The disproportionation equilibrium depends
`on the nature of the supporting electrolyte, demonstrating that ion pairing is
`operative even in solvents as polar as these, although it appears to involve only
`the dianion, not the radical anion. In low-dielectric-constant solvents, where ion
`pairs and even higher aggregates are prevalent, or with electrolytes (such as
`lithium salts) that form strong covalent bonds [50], association between the elec-
`trolyte and charged intermediates plays an even greater role in affecting mea-
`sured potentials. In fact, in the specific case of radical anion disproportionation
`under discussion here, AEIt2 can even become negative; that is, the two one-
`electron waves in polar media coalesce into a single two-electron wave [49,51].
`
`V. EXPERIMENTAL PROCEDURES
`
`A. Obtaining Dry Solvent
`
`Water is the most ubiquitous impurity in organic solvents. Considering how
`sensitive many reactions are to the presence of water, the literature on drying
`solvents is surprisingly full of contradictions and misinformation. Anyone
`doubting this statement is referred to a remarkable series of papers by Burfield
`[32,34,52,53]. The Introduction sections to these papers will provide a healthy
`dose of reality to those who feel they can rely solely on their chemical intuition
`when deciding how to purify their organic solvents. Burfield developed a highly
`accurate method for determining the water content of organic solvents and then
`used it to assess the dryness of a number of common organic solvents, each of
`which was dried by a variety of methods. (The analytical method is based upon
`treatment of the anhydrous solvent with tritiated water, followed by drying and
`analysis of the solvent for residual tritium by scintillation counting [54].) Burfield
`made a number of useful discoveries in the course of this work, including the
`following: (a) a number of common laboratory drying procedures are almost
`totally ineffective with some solvents; (b) a reagent that is good for drying one
`solvent may be rather poor with another; and (c) probably the best general
`method for drying most solvents is by the use of molecular sieves, assuming
`they are used properly, that is, (1) activating 3-,~ molecular sieves for 15 h at
`300°C, (2) allowing them to cool in a P205 dessicator, (3) treating the solvent
`with activated sieve (5-10% w/v) for 6 h to 7 days, and (4) allowing the sol-
`vent to stand over a second batch of fresh molecular sieve. This procedure was
`shown to reduce water levels in the.following solvents to the indicated levels:
`DMF: <1.5 ppm; AN: _<0.6 ppm; DMSO: _<10 ppm. A series of experiments
`
`Lupin Ex. 1054 (Page 12 of 19)
`
`

`

`Solvents and Supporting Elect.rolytes
`
`479
`
`in which the water content of DMF standing over molecular sieve was moni-
`tored by gas chromatography confirmed the fact that several days are necessary
`to achieve the lowest water levels [55]. An important qualification must be added
`at this point, since it was not considered by Burfield: His analysis assumes that
`the only ~olvent impurity is water. This is frequently not the case. DMF con-
`tains readily detectable quantities of dimethylamine, DMSO contains traces of
`dimethyl sulfide, AN can contain acrylonitrile, THF and other ethers contain
`peroxides, etc. Thus it is frequently necessary to subject the solvent to a pre-
`liminary treatment to remove such impurities and lower its water content be-
`fore drying. (Commercial DMF is frequently >0.1 M in water, for example.)
`Peroxides can usuaIly be removed by passing the solvent through a column of
`activated alumina.
`Dipolar aprotic solvents are generally hygroscopic, and contact with the
`atmosphere should be avoided after drying. For critical applications, the solvent
`should be distilled into the electrochemical cell [56] or in a dry box, or passed
`through a column containing the dry agent repeatedly in a sealed system [57].
`Simpler and almost as satisfactory alternatives include drying the solvent over
`a molecular sieve either in a dropping funnel attached to the cell, followed by
`direct addition of the solvent to the cell [55], or in a separate flask, followed
`by solvent transfer using cannula techniques [58]. Supporting electrolytes are
`frequently hygroscopic and should be dried before use; this is readily done in
`an Abderhalden appar

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