throbber
Stereochemistry of
`Organic Compounds
`
`ERNEST L. ELIEL
`Department of Chemistry
`The University of North Carolina at Chapel Hill
`Chapel Hill, North Carolina
`
`SAMUEL H. WILEN
`Department of Chemistry
`The City College of the City University of New York
`New York, New York
`
`With a Chapter on Stereoselective Synthesis by
`
`LEWIS N. MANDER
`Research School of Chemistry
`Australian National University
`Canberra, Australia
`
`A Wiley-lnterscience Publication
`JOHN WILEY & SONS, INC.
`New York
`• Chichester
`• Brisbane
`
`• Toronto
`
`• Singapore
`
`PARAGON - EXHIBIT 2022
`
`

`
`This text is printed on acid-free paper.
`Copyright © 1994 by John Wiley & Sons, Inc. ·
`All rights reserved. Published simultaneously in Canada.
`
`Reproduction or translation of any part of this work beyond
`that permitted by Section 107 or 108 of-the 1976 United
`States Copyright Act without the permission of the copyright
`owner is unlawful. Requests for permission or further
`information should be addressed to the Permissions Department,
`John Wiley & Sons, Inc., 605 Third Avenue, New York, NY
`10158-0012.
`
`Library of Congress Cataloging in Publication Data:
`Eliel, Ernest Ludwig, 1921-
`Stereochemistry of organic compounds I Ernest L. Eliel, Samuel H.
`Wilen, Lewis N. Mander.
`p. em.
`"A Wiley-Interscience publication."
`Includes index.
`ISBN 0-471-01670-5
`1. Stereochemistry. 2. Organic compounds. I. Wilen, Samuel H.
`II. Mander, Lewis N. III. Title.
`QD48l.E52115 1993
`547.1'223--dc20
`
`93-12476
`
`Printed in the United States of America
`
`10 9 8 7 6 5 4 3 2 1
`
`

`
`1071
`
`ascribed to conformational rigidity and to the presence of isotactic helical
`of a preferred helical sense in the copolymer. A similar example of such a
`cooperative phenomenon (with a resulting high optical rotation) involving a

`consisting of mostly achira:l poly(n-hexyl isocyanate) incorporating as
`as 0.12mol% of a nonracemic chiral isocyanate [(S)-(-)-2,2-dimethyl-1,3-
`isocyanate] has been described (Green et al., 1989).
`
`Attachment of achirotopic chromophoric groups such as dehydrophenylalanine and
`azobenzene to polypeptides [e.g., poly(L-glutamic acid)] generates synthetic macro(cid:173)
`molecules whose CD spectra reflect the presence of inherently chiral chtomo(cid:173)
`phores. This spectral feature results because the pendant side chains serve as
`chirality reporters of the secondary structure of the peptides ( Ciardelli and Pieroni,
`1980).
`
`In subsequent studies on dimethyloctene-styrene and similar copolymers, the
`CD could be measured deeper into the UV spectral region. Other CD bands
`typical of isolated benzenoid electronic transitions CB and 1La) signaling the
`chirotopicity of the benzene ring were observed. In contrast to model compounds,
`such as the conformationally restricted 2-phenyl-3,3-dimethylbutane (Salvadori et
`. al., 1972), the copolymer exhibits, in addition, an exciton-like couplet centered at
`about 190 nm. The latter was attributed to a helical conformation in which
`benzene rings in the same chain are sufficiently close to couple. The enantiomer
`. exhibiting negative chirality incorporates a right handed helix ( Ciardelli et al.,
`1972). The CD of such helical copolymers can be calculated by means of a
`classical theory developed by DeVoe (DeVoe, 1969, 1971; see Hug et al., 1974 and
`Section 13-4.a).
`It is only since the 1980s that CD studies have permitted the observation of
`CEs in the vacuum UV region of hydrocarbon polymers devoid of aromatic
`groups. The CD spectra of films of poly-(S)-4-methyl-1-hexene and poly-(R)-3,7-
`dimethyl-1-octene exhibit a CD band at 158 nm that is ascribed to conformations
`containing helical segments having a common helix sense ( Ciardelli and Salva(cid:173)
`dori, 1985).
`
`13-5. APPLICATIONS OF OPTICAL ACTIVITY
`
`a. Polarimetry
`
`The actual measurement of optical activity may be carried out with either manual
`or photoelectric polarimeters. Manual polarimeters have changed relatively little
`, since the first instruments were developed some 140 years ago (Lowry, 1964, p.
`180). Photoelectric polarimeters, the type nowadays commonly found in research
`laboratories, have greatly reduced the tedium formerly associated with the
`measurement of optical rotation with manual instruments. Moreover, photoelec(cid:173)
`tric polarimeters are much more accurate and sensitive, permitting the rapid and
`meaningful recording of quite small absolute rotation values a to about ±0.002°
`and, consequently, the use of smaller samples. Polarimeters fitted with microcells
`may even serve advantageously as detectors in HPLC resolutions (Mannschreck,
`
`

`
`1072
`
`Chiroptical Propertie{· .... ·
`
`·.
`Eigelsperger, and Stuhler, 1982; Mannschreck, 1992; Pirkle, Salvadori, et al.;.
`1988; Lloyd and Goodall, 1989). A laser-based polarimetric HPLC detector h~s · ·
`been shown to be sensitive to as little as 12 ng of sample (Yeung et al., 1980). For· .
`the advantages of the use of CD detectors in HPLC, see Salvadori, Bertucci, and
`Rosini (1991); Mannschreck (1992).
`The laser polarimetric detector has been adapted to HPLC analysis not only
`of optically active samples but also, in a different way, as a universal detector for
`achiral, that is, optically inactive, substances. In this technique, termed "indirect .
`polarimetry," the mobile phase is optically active, containing, for example.
`(-)-2-methyl-1-butanol or ( + )-limonene, and the detector output due to the
`optically active solvent is zeroed. Under these conditions, any optically inactive ·
`fraction passing through the detector cell is sensed since the concentration of the'
`optically active solvent is thereby reduced. The response of the detector is
`universal, like that of a refractive index detector, but is more sensitive than the. · · · ·
`latter (Bobbitt and Yeung, 1984, 1985; Yeung, 1989). The simultaneous measure- .
`ment of absorbance and optical rotation during the liquid chromatographic . .
`resolution of chiral substances on enantioselective stationary phases make pbs: . ·
`sible the determination of the enantiomer composition in spite of extensive peak-- ~­
`overlap (Mannschreck et al., 1980; Mannschreck, Eigelsperger, and Stuhler,· ·. ·
`1982; compare Drake, Gould, and Mason, 1980).
`·:"·
`For a brief discussion of polarimetry and its instrumentation, see the review·.:· ·
`by Lyle and Lyle (1983); for a more extensive treatment, see Heller and Curme':·
`(1972).
`.
`The measurement of optical activity has traditionally been the method -of ; ·
`choice to establish the nonracemic character of a sample of a chiral compound·.·' ..
`and, when quantitatively expressed as a ratio [a] I [a] max, of its enantiomeric
`composition (optical purity). In contemporary practice, chiroptical measurements
`have to a large extent been replaced by NMR and by chromatographic analyses :.:
`for the purpose of determining enantiomeric compositions (cf. Chapter ,6).'~:
`Nevertheless, the use of [a] for this and other purposes continues. The reasons. ·
`are that the measurement is easy to carry out and one may wish to compar~.
`experimental values of [a] with those in the literature. While substantial coliec-: .
`tions of optical rotation data exist, for example those in various handbooks and< .
`chemical supplier catalogs, it should not be assumed that values of [a] provide·d'·:
`are those of enantiomerically pure compounds. A consistent set of spe<;tfi~­
`rotation data for amino acids including temperature coefficients has been com~.
`piled by ltoh (1974).
`.
`..
`Optical activity has been used (a) to determine if a given unknown substap:ce. ~-.
`is chiral or achiral; (b) to ascertain the enantiomeric composition of _chi~al.
`samples, either qualitatively or quantitatively; (c) to study equilibria; the m.uta~:
`rotation or change in rotation of equilibrating stereoisomers as a function ofJime0 . _·.
`is one such phenomenon (Eliel, 1962; for a recent example, see Arjona;: :
`Perez-Ossorio, et al., 1984); and (d) to study reaction mechanisms. Qt~~l': · ·
`chiroptical techniques, namely, ORD and CD, have increasingly replaced pol:ari-: ..
`metry in these applications, especially in the past 20 years. For review( of
`applications of polarimetry, see Lowry (1964), Eliel (1962), Legrand and Roumep. -·
`(1977), and Purdie and Swallows (1989).
`· · ··
`Polarimetric methods remain useful for quality control in pharmacology
`
`

`
`Applications of Optical Activity
`
`1073
`
`food-related industries (Lowman, 1979; Chafetz, 1991); there are also numerous
`applications in forensic, clinical, pharmaceutical, and agricultural chemistry (Pur(cid:173)
`die and Swallows, 1989). The percentage of sucrose in commercial samples is still
`being determined by polarimetry (saccharimetry); in ·the trade this is called
`"direct polarization". The cost of raw sugar is based on the results of the
`polarimetric analysis; if the analyte solution is dark, the raw sugar is first clarified
`· by precipitation of the dark side products with basic lead acetate (Cohen, 1988).
`An example of application d (above) is the methanolysis of the tosylate of
`(R)-( + )-C6H 5CH2SCH2CH(CH3)CH20H that leads to a partially racemized
`methyl ether. The intervention of a cyclic (symmetrical, and hence achiral)
`intermediate, via neighboring group participation, was inferred (Eliel and Knox,
`1985).
`The magnitude of rotation a, in degrees, fundamentally depends on the
`number of molecules of the sample being traversed by the linearly polarized light
`as well as on their nature, hence optical activity is not a colligative property.
`Values of a are affected by many variables, among which are wavelength, solvent,
`concentration, temperature, and presence of soluble impurities. It must also be
`mentioned that large molecules, such as proteins, may spontaneously orient
`themselves in solution, and consequently no longer be isotropic. The measure(cid:173)
`ment of the rotatory power of such substances may then be complicated by the
`occurrence of linear dichroism (see Heller and Curme, 1972, p. 67).
`As already pointed out in Section 1-3 (q.v.), rotation magnitudes are usually
`normalized to a quantity called the specific rotation [a] that was introduced by
`Biotin 1835 (Biot, 1835; cf. Lowry, 1964, p. 22), Eq. 13.28,
`
`[a]= a/tp = altc
`(13.28)
`where e is the length of the cell in decimeters, p (for undiluted liquids) is the
`density in grams per milliliter (g mL - 1
`) and c is the concentration also in grams
`per milliliter. The units of [a] are 10- 1 deg cm2 g- 1 (see also Eq. 1.1 and Section
`1-3).
`Comparison of specific rotations of homologues, and of organic compounds
`generally, is more significant if a modified Biot equation is used in which the
`quantity called the molar rotation [<I>] depends on the number of moles of
`substance traversed by the linearly polarized light, Eq. 13.29.
`
`[<I>]= [a]M/100
`
`(13.29)
`
`where M is the molecular weight. The units of [<I>] are 10 deg cm2 mol- 1 (see also
`Eq. 1.2) (IUPAC, 1986).
`The cumulative effect of the above-mentioned variables on [a] or [<I>] is
`potentially very large. A practical consequence is that precise reproduction of
`published rotation values, from laboratory to laboratory, or even from day to
`day in the same laboratory, is difficult to achieve (Lyle and Lyle, 1983).
`This sensitivity to numerous variables (Schurig, 1985) and the absence of
`major tabulations of critically evaluated absolute rotation data is responsible
`for the decreasing reliance on optical activity as a measure of enantiomeric
`composition .
`
`.
`·~·.
`
`. ·
`. .
`
`

`
`1074
`
`Chiroptical Properties·
`
`Much preliminary work may be necessary to increase the low accuracy of routine
`optical rotation measurements. For an outstanding example of a study of some of
`the variables affecting the rotation of a-methylbenzylamine, mandelic acid, and
`ephedrine, see the dissertation by Zingg (1981).
`
`.
`
`The sign of rotation is often the only experimental criterion for the specifica- ·
`tion of configuration. It is important to stress how frequently and how easily this.
`property may change for a given substance, for example, (R)-2-hydroxy-1,1'~
`binaphthyl has [a]~0 +4.77 [c = 0.86, tetrahydrofuran (THF)], +13.0 (c = 1.12,
`.
`w
`THF), and [a] 0 -5.2 (c = 1.03, CH30H) (Kabuto et al., 1983). Even tartaric ·
`acid, one of our configurational standards, does not exhibit an invariant sense of .
`rotation: [<I>] 578 -12.9 (24°C), -0.9 (57°C), and +10.8 (94°C) (all c = 10; ··
`dioxane); +21.3 (24.7°C, H20), +6.6 (24°C, EtOH), +0.3 [25.3°C, N,N-

`dimethylformamide (DMF)], -12.9 (24°C, dioxane), and -14 (25.2°C, Et20) (ll,ll
`c = 10) [both sets of data measured on (R,R)-tartaric acid] (Hargreaves and
`Richardson, 1957).
`When the sign of rotation shows a strong solvent, concentration, wavelength·,. · · ·
`or temperature dependence, the association of such sign with a given configura~ ..
`tional descriptor is arbitrary. In the case of 4-amino-1-( diethylamino )pentane·
`(Fig. 13.70), 37 [a] is less than 0 in ethanol (365-589 nm) but greater than 0 when
`measured in the absence of solvent (neat). The configuration of the sample
`derived from L-glutamic acid was referred to as (R)-(-) rather than (R)-( +}. · ·
`because of the much larger magnitude of the rotation of the neat sample over that.
`in ethanol (Craig et al., 1988). This, and the other examples cited, serve to
`emphasize the crucial importance of specifying and recording the precise ex-·
`perimental conditions of measurerp.ent of optical rotations and of chiroptical. ·.
`properties in general. In particular, confusion can arise when the sign of rotation
`is related to a given configuration and the solvent is not specified.
`Occasionally, specific rotations of samples are very small. When that situation · ··
`arises, especially in resolutions or in stereoselective syntheses in which strongly
`rotating reagents are used, exceptional care must be taken to insure that th~. ·.
`rotation of the product is not spurious. A small amount of impurity having a large'· ·
`[a] may overwhelm (or at least seriously falsify) the rotation of a sample having·.a · ·
`small [a] (for a problem case, see Baldwin et al., 1969; see also Goldberg et at.;
`1971). Achiral contaminants, particularly solvents, will also affect the opti¢ai ··
`activity of a sample (Lyle and Lyle, 1983; Schurig, 1985).
`. ..
`Traces of achiral compounds, including solvent residues, normally would be.
`expected to reduce [a] (by dilution of the sample) and hence to artificially lower . ·.
`the optical activity of nonracemic samples (but not the enantiomeric purity of
`chiral solute). However, the converse may also be observed, for example, the.:
`optical activity of 1-phenylethanol at 589 nm is increased when acetophenone, ·a
`possible contaminant, is present in the sample (Yamaguchi and Mosher, 197~)~.
`The enhanced optical activity of the alcohol comes about because

`.
`properties are induced in the achiral ketone by the alcohol; in the example,
`induction is superimposed on and swamps the typical and opposite dilution e:ffe~t ·
`(cf. Section 13-4.e).
`The case of low rotation warrants further comment. There are two sltlJati.O!lS
`in which no optical rotation is observed with enantiomerically enriched ., .. ., .. ..,~,,~, ,,
`
`

`
`· ... Applications of Optical Activity
`
`1075
`
`(a) the experimental device used (by implication this includes the eye) is of
`insufficient sensitivity, and (b) the specific conditions of measurement are such
`·.·that a is, in fact, accidentally equal to zero.
`In the first situation, the measurement threshold is such that there is no clear
`signal (rotation) distinguishable from iJ1.Strumental noise. The condition is one of
`· operational null. Progressive dilution of a solution of an optically active com(cid:173)
`. pound eventually leads to a sample that is no longer palpably optically active
`when the operational null threshold is crossed. Such a sample no longer reveals its
`enantiomeric excess; the sample is said to be cryptochiral (Mislow and Bickart,
`1976/1977).
`
`The term cryptochiral is not to be confused with the analogous expression
`"stereochemically cryptic" that refers to a stereoselective chemical reaction whose
`stereochemical outcome is hidden (Hanson and Rose, 1975; cf. also Section 8-5).
`
`Notable examples of enantiomerically enriched compounds that are crypto(cid:173)
`chiral as a consequence of inherently low optical rotation magnitude are shown in
`Figure 13. 71. The cryptochirality condition may conceivably be lifted by measur(cid:173)
`ing a different chiroptical property, for example, vibrational circular dichroism
`(VCD) (Section 13-6).
`The second type of cryptochirality arises when the measurement of rotation
`accidentally takes place in the vicinity of a change in sign (see below and above).
`For an example involving a change in concentration of dimethyl a-methylsucci(cid:173)
`nate, CH30 2CCH(CH3)CH2C0 2CH3 , see Berner and Leonardsen (1939). At a
`certain concentration, the measured rotation is necessarily zero (crossover point)
`and the sample is then accidentally cryptochiral. Note that a distinction between
`stochastic achirality (cf. Section 13-2.a) and cryptochirality cannot be made unless
`
`38
`
`CH3 H
`I
`I
`H3c-c-c-oH
`I
`I
`CH3 D
`
`40
`
`C2Ha
`
`I
`n-CeH13-C-n-C4He
`I
`
`n-C3H7
`
`41
`
`Sanderson and
`Mosher (1966)
`
`Wynberg, et al.
`(1965)
`
`Figure 13.71. Compounds illustrating cryptochirality.
`
`

`
`1076
`
`Chiroptical Properties
`
`the former be lifted by a change in measuring device or the latter by a change in
`conditions, the latter being easier to achieve. Other examples of the second type
`of cryptochirality have been given above and in Section 13-4.
`The dependence of optical rotation on the wavelength of the light, ORD, has
`been discussed in Sections 13-2.b and 13-4.
`
`Effect of Temperature
`
`The effect of temperature on chiroptical properties may be ascribed to the
`following phenomena (Legrand and Rougier, 1977): (a) changes in density of the
`solute and/or the solvent that alter the number of molecules being observed; (b)
`changes in the population of vibrational and rotational energy levels of the chiral
`solute; (c) displacement of solute-solvent equilibria; (d) displacement of con(cid:173)
`formational equilibria; and (e) aggregation and microcrystallization of the chiral
`solute (cf. enantiomer discrimination, Section 6-2).
`In general, [a] changes 1-2% per degree Celsius, but larger changes (up to
`10% per degree Celsius) are not unknown, for example, [aJn of aspartic acid,
`H02CCH(NH2)CH2C02H, in water (c=0.5%) is 4.4 at 20°C, 0 at 75°C, and
`-1.86 at 90°C. The change in sign at 75°C (temperature of cryptochirality, see
`above) is noteworthy (Greenstein and Winitz, 1961, p. 78).
`An early example of an increase in specific rotation with increasing tempera(cid:173)
`ture that was ascribed to a shift in a conformational equilibrium is that of
`2-butanol (Horsman and Emeis, 1965). Other examples are discussed in Section
`13-4.e).
`Strong dependence of the optical rotation on temperature may be found even
`among hydrocarbons, for example, 3-phenyl-1-butene whose neat rotation,
`[a]~ -5.91, for the enantiomerically pureR enantiomer increases linearly 0.18°/
`oc from 16 to 29°C. Here, it is likely that the temperature exerts a strong
`conformational bias (Cross and Kellogg, 1987).
`
`Effect of Solvent
`
`The "nonspecific" influence of solvent on the specific rotation may be corrected
`by calculation of a quantity called the specific rotivity fi' that includes the
`refractive index of the solvent ns (see Heller and Curme, 1972):
`
`fi' = [3a]l(n; + 2)
`
`(13.30)
`
`Several examples of dramatic changes in the angle of rotation as a function of
`solvent have been given above. Many instances of changes in sign of [a] have also
`been recorded for amphoteric substances, such as the amino acids, as the pH is
`changed (Greenstein and Winitz, 1961, p. 1727). An exceptional example of the
`effect of solvent on [a] is given in Figure 13. 72. Given examples such as these, it
`is disconcerting how frequently the mention of the solvent is omitted from
`experimental descriptions of the optical rotation.
`Care in choosing the solvent to be used in the measurement of [a] is
`necessary in view of the several specific types of interaction that are possible
`between solute and solvent. In general, one recognizes the intervention of .·
`hydrogen bonds when oxygen-containing solutes, such as carboxylic acids, aide-
`
`

`
`1077
`
`Forma midi
`
`~
`lNJ
`bH3
`
`Ethylene bromide
`
`-6
`-8-
`
`-10
`-12
`-14
`
`-16
`
`-18
`- 20100
`
`eo
`
`60
`
`40
`p
`Figure 13.72. Specific rotation of nicotine in various solvents ( p =concentration of solute in grams
`· · per 100 grams of solution). At p = 100, the "bulk" rotation [a] 100 should be a constant, as observed,
`· · and at p = 0, [a] should tend to the intrinsic rotation {a} (p. 1079). [Adapted with permission from
`• Winther, C. (1907), Z. Phys. Chern., 60, 621.]
`
`and ketones, and alcohols, are dissolved in hydroxylic solvents; in some
`'cases reactions, such as hemiacetal formation, may occur. In addition, dipole(cid:173)
`. dipole interactions and changes in conformer populations are important sources of
`·.solvent-induced variations in rotation magnitude (Lyle and Lyle, 1983).
`The effect of intermolecular solute association of polar solutes on [a] in
`· nonpolar solvents has already been.,pointed out (see above). Solute-solute
`' association effects may be leveled out or suppressed in polar solvents by competi~
`·. · tion with (concentration-independent) solute-solvent association. Polar solvents,
`, such as ethanol, may break up solute-solute association leading to a smaller
`concentration dependence of [a], as is found with nicotine (Fig. 13.72). Such
`· , findings illustrate the desirability of using methanol or ethanol as a solvent in
`polarimetry.
`In some instances, hydrogen bonding is known to be responsible for changes
`•. in [a] with concentration and/or solvent. Compounds 42 and 43 (Fig. 13.73)
`. exhibit a remarkable solvent dependence of the sign of [a] 0 for the RRISS (syn)
`, 42 diastereomer that is not found in the case of the RS I SR (anti) diastereomer 43 .
`. The sign of [a]~ of (4R,5R)-42 is(+) in methanol and(-) in chloroform. This
`
`

`
`1078
`
`Chiroptical Properties·
`
`OR
`HsC&~~1oH21
`
`OCH2C&Hs
`
`Rx;
`
`OH
`
`43 (R= H) S,R shown
`
`44
`
`42 (R = H) R,R shown
`42a (R=Ac)
`
`Figure 13.73. Structures 42-44.
`
`difference has been ascribed to a conformational change: the predominant
`methanol-solvated (OH/OCH2C6H 5 ) anti conformer gives way to a (OH/
`OCH2C6H 5 ) gauche intramolecularly hydrogen-bonded conformation in chloro~
`form. Such sign reversal is not seen in the benzyl ether-acetate derivative 42a, in
`the corresponding diol (the latter appears to prefer the gauche conformation ·
`regardless of solvent) or in the diol acetonide. Reversal of the sign of [a] 0 would.
`seem to be precluded in the predominant zigzag (all-anti) conformation of the
`molecular skeleton. A similar sign reversal was observed in a series of 2-alkoXy.
`alcohols 44 (Fig. 13.73) presumably for the reasons advanced above. The free dibl.
`(S)-1,2-dodecanediol exhibits [a]~ -10.1 (EtOH) but +0.9 (CHC13 ). This
`suggests that here, too, the intramolecularly hydrogen-bonded conformer prevails c
`in CHC13 (Ko and Eliel, 1986).
`Another example of optical rotation sign reversal is found with compound 45
`(Fig. 13.74): [a] 0 +14.0 ± 0.6 (CHC13 ) and -2.7 ± 0.6 (CH30H) (Suga et al.;
`1985). The molar rotation [<I>] 0 was found to be independent of concentration .irt
`12 solvents of varying polarities. The principal factor responsible for the s,ign·
`reversal was intramolecular hydrogen bonding between the carbonyl and hydroxyl· :
`groups in nonpolar solvents (confirmed by IR and 13C NMR measurements)and ·.
`its absence in the presence of strongly solvating media (alcohols, CH3CN,··'or
`acetone) as a result of competing intermolecular hydrogen bonding between·t~e ..
`solvent and the solute (Suga et al., 1985). A sign reversal of aCE was also noted·
`in the ORD of 46 (Fig. 13.74) in CHC13 and CH30H but this was not reflected in' .
`the sign of [a] 0 . The latter fact points up once again the desirability of carrying···
`out studies of chiroptical properties over a range of wavelengths and preferably.·
`into regions that reveal the responsible CEs.
`,.,. : ·. ·
`A particularly clear-cut example of a conformational equilibrium that ·is.·
`responsible for changes in chiroptical properties over the range of 210-350nm.is
`shown by ketone ( + )-47 (Fig. 13.75). As the solvent is changed from cyclohexane.· ,.
`to acetonitrile or methanol, the effect of increasing solvent polarity on. the. • ·
`dipole-dipole repulsion (as well as solvation effects) between the adjac~nt .
`permanent dipoles (C=O and C-Br) causes a conformational change: the broniine , ·
`changes from axial to equatorial and significant changes in the CD ·.
`· ·
`
`45
`Figure 13.74. Structures 45-46.
`
`46
`
`

`
`Applications of Optical Activity
`
`1079
`
`s=b-··
`
`H
`
`0
`
`47
`48
`Figure 13.75. Structures 47-48.
`
`(Kuriyama et al., 1967). Striking changes in [a] 0 of propylene oxide, including
`sign reversal, are observed as the solvent is changed from benzene to water. In
`this instance, we are cautioned against ascribing the effect of solvent directly to
`conformational changes (Kumata et al., 1970).
`
`Effect of Concentration
`
`Equation 13.28 suggests that the specific rotation should be independent of
`concentration. It is not hard to find evidence that this constancy holds only over
`very narrow concentration ranges and, in some solvents, not at all (the example
`of nicotine is found in Fig. 13.72; other examples may be found in the book by
`Lowry, 1964, Chapter VII).
`As early as 1838, Biot suggested that the specific rotation followed a linear
`relationship, such as that of Eq. 13.31, where a and b are constants
`
`[a]= a+ be
`
`(13.31)
`
`and cis the concentration (Lowry, 1964, p. 90). Constant a has been equated with
`a new quantity called the "intrinsic rotation" {a}, a true constant corresponding
`to the specific rotation in a given solvent at infinite dilution; [a Jc__,0 = {a} (Heller
`and Curme, 1972, p. 163). Obviously, {a} can only be calculated since ex(cid:173)
`perimentally, as the concentration is reduced the rotation must vanish.
`The intrinsic rotation is the specific rotation for a system free of solute-solute
`interactions. However, solute-solvent interactions are maximized in {a}, which
`can differ greatly from solvent to solvent, for example, for nicotine (Fig. 13.72).
`Since, obviously a = oo at 0% solute, the values at very low concentrations must
`be extrapolated. Conversely, as the concentration of solute increases, solute(cid:173)
`solute interactions become dominant and the effect of the solvent eventually
`vanishes: [a Jc__, 100 = [a lneat tends to a constant value that is identical for all
`solvents.
`rotation of (S)-2-phenylpropanal,
`the specific
`A
`recent report on
`CH3CH(C6H 5)CH=O, makes it clear that even at relatively low concentrations in
`benzene (c = 1-4), changes of the order of 1-2% are found in [a] 0 as the
`concentration is doubled (see Table 13.3; Consiglio et al.; 1983). The accurate
`determination of optical purities is thus seen to be dependent on the careful
`measurement of rotations as well as on comparison of the resulting [a] values
`with reference [a] values measured in the same solvent, at the same temperature,
`and at the same concentration (Consiglio et al., 1983). The reader is also
`
`

`
`1080
`
`Chiroptical Properties
`
`Influence of Dilution on [a]~ for {S)-2-Phenylpropanal•.b
`TABLE 13.3.
`Concentration (g/100 mL- 1
`
`)"
`
`[a]~t
`
`[a]~
`166.6
`161.8
`Neat
`182.2
`177.9
`46.43
`195.4
`190.5
`18.57
`201.9
`196.6
`9.29
`207.9
`202.7
`7.43
`205.8
`211.3
`3.72
`214.7
`209.1
`1.49
`• Reprinted with permission from Consiglio, G., Pino, P., Flowers, L. 1., and
`Pittman, C. U., Jr. (1983), J. Chern. Soc. Chern. Cornrnun., 612. Copyright©
`Royal Society of Chemistry, Science Park, Milton Road, Cambridge CB4 4WF,
`UK.
`b Optical purity 68%.
`c Benzene solution.
`
`reminded that [a ]0
`reflects, but is not necessarily linearly related to, the . .
`enantiomeric composition (Horeau effect: Horeau and Guette, 1974; see Section
`· ..
`6-S.c).
`Some very subtle effects are manifested by the rotatory power. For example·:.: ·
`[a ] 436 of menthol and of menthol-0-d diverge as the solute concentration i~ .
`increased 100-fold (in cyclohexane). This differential concentration effect reveals
`that intermolecular hydrogen bonding is subject to a thermodynamic isotope :
`'
`effect (Kolbe and Kolbe, 1982).
`In the absence of experimental information, calculations, or obvious structui:~ ,
`al features, the detailed aspects of chiroptical data including sign reversal of CBs .
`cannot be explained. A detailed study of the chiroptical properties of alkyL:·
`substituted succinic anhydrides 48 (R' = H, R =alkyl; Fig. 13.75) revealed a great'..
`sensitivity of the CD (CEs at ca. 220 and 240 nm) to solvent polarity; to'
`temperature, and to the size of the alkyl substituent. The results could generally··
`be accommodated by a sector rule (for anhydrides with local C2v symmetry of th.e- • ·
`chromophore, that is, a planar anhydride group) (Gross, Snatzke, and Wessling;:· .•
`1979), but the absence of information about conformational equilibria' and . ·
`solvation effects made it impossible to offer explanations for the detailed features:
`of the CD (Sjoberg and Obenius, 1982).
`··.r
`
`b. Empirical Rules and Correlations. Calculation of Optical Rotation
`
`Ever since simple curiosity about optical activity gave way to its ap]pU<;attoR; ,
`efforts have been made to calculate the magnitude and sign of the optical ro!a#qn.
`in relation to structure and configuration.
`.
`One of the very oldest correlations between structure and rotatory pow~r
`that of Walden who observed that the molar rotations of diastereomeric salt~ ·
`dilute solution are additive properties of the constituent ions (Walden, . ·
`Jacques et al., 1981, p. 317). Arithmetic manipulation of the molar rotattcms.o!'.:.
`diastereomeric salts, such as those obtained in a resolution, may thus permit
`to estimate the enantiomeric purity achieved during a resolution mediated,
`these salts.
`
`

`
`The molar rotations [<1>] 0 of two of the nonenantiomeric diastereomeric salts of
`a-methylbenzylamine mandelate are -182.3 and +169.7 (in H 20). From these
`data, the molar rotations of the constituent ions are calculated to be [<1>] 0 = ~
`([-182.3] + [+169.7]) = -6.3, that is, ±6.3 and [<1>] 0 = ~ ([-182.3]- [+169.7]) =
`-176, that is, ±176. The latter value is assigned to the mandelate ion [literature
`values of [<1>] 0 for sodium and potassium mandelates are 182 and 178, respectively
`(both in H 20; Ross, et al., 1937)]. Combination of this value with that for
`ephedrinium hydrochloride, [<1>] 0 69.8 (H 20) (Overby and Ingersoll, 1960) yields
`the molar rotations of the two ephedrine mandelate diastereomers, [<1>] 0 = 176 +
`69.8 =246 and 176-69.8 = 106. The experimental values of these rotations ob(cid:173)
`tained by Zingg, Arnett, et al. (1988) are ±250 and ±107, respectively (both in
`H 20).
`
`Analogously, molar rotations of inclusion compounds would be expected to
`additive properties of the host and guest molar rotations. In both cases,
`Ji,U~~·····vity of rotations would not necessarily obtain when strong intermolecular
`takes place.
`In the case of covalent compounds, early correlation attempts also made use
`of the concept that the rotations of compounds containing several chiral centers
`be calculated by adding rotation contributions from each of these centers.
`concept, incorporated in van't Hoff's empirical "Principle of Optical Super(cid:173)
`!i?PIOSJLtlcm." that individual chiral centers in a chiral compound make independent
`... vu•u .. •uu•vu" to the molar rotation (van't Hoff, 1893; Kuhn, 1933, pp. 394, 423) is
`successfully being applied in very limited contexts.
`The relative configurations of the diastereomeric (R)-0-methylmandelate
`esters of 49 were assigned by application of the van't Hoff principle. Contribu(cid:173)
`. tions to the specific rotations from the octalin portions of the ester molecules were
`· estimated from the rotations of ( + )-dihydromevinolin 50 and lactone 51 (Fig.
`13.77) to be approximately 100 (148.6- 48.8) while that of the (R)-mandelate
`was independently known to be strong and positive [(S)-(-)-methyl
`uuc:.uu•v•ate] has [a]~ -124 (Barth, Mosher, Djerassi, et al., 1970).
`, the configurational assignments were ±75 f

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