throbber
3986
`
`J. Am. Chem. Soc. 1992, 114, 3986-3988
`empty p orbitals with the occupied s orbitals, i.e., hybridization.
`A ligand field of Dy symmetry is shown to induce strong bonding
`between two Be atoms which are weakly bound in the Be2 mol-
`ecule.
`
`orbitals. Upon bonding to the two CO molecules, the Be 2s orbital
`mixes with its low-lying 2p orbitals to form sp2 hybrid orbitals,
`two of which are used to form bonds with the carbon lone pair
`orbital of the CO, while the third is occupied by one of the un-
`paired electrons. The second unpaired electron occupies a b,
`orbital perpendicular to the plane of the molecule, which is a Be
`2p type orbital that is involved in a ir back-bonding interaction
`with the CO ir* orbital.
`The 3B) state of Be(CO)2 is analogous to the lowest 3B, state
`of methylene on the basis of the nature and type of the highest
`occupied orbitals. Formation of Be2(CO)4 from two Be(CO)2
`molecules in their 3B, states results in a Be-Be double bond. An
`analysis of the two highest occupied orbitals of (CO)2BeBe(CO)2
`shows significant bonding between the two Be atoms. The highest
`occupied (b3u) orbital is bonding between the Be atoms, with
`each Be also involved in back-bonding with the t* orbitals of
`the CO. The second highest occupied orbital (ag) is a a type
`bonding orbital. The triplet state arising from the one-electron
`excitation to the * orbital of (CO)2BeBe(CO)2 is 16.2 kcal below
`the singlet state at the HF level of theory. The inclusion of electron
`the order of the states with the 'A,
`correlation effects reverses
`below the triplet state by 10.6 kcal at the MP2 level of theory.
`The binding energy of (CO)2BeBe(CO)2 is computed with
`respect to dissociation to two Be('S) and four CO molecules as
`well as to two ^ Be(CO)2 fragments. The computed binding
`energies are given in Table I. Even at the HF level, the interaction
`energy between the two Be(CO)2 fragments in (CO)2BeBe(CO)2
`is 15.5 kcal. Note that the Be2 molecule shows no binding at this
`level of theory.11
`*1The complete fragmentation of (CO)2BeBe-
`(CO)2 into two Be atoms and four CO molecules requires 36.0
`kcal at the HF level of theory. At the MP4(SDQ) level, the
`strength of the Be-Be bond in this molecule is 50.0 kcal, and the
`complete fragmentation requires 68.3 kcal. The contributions of
`electron correlation to the binding energy of both Be(CO)2 and
`(CO)2BeBe(CO)2 are large, indicating the need to reoptimize the
`geometry of these molecules to include electron correlation ef-
`fects.10 This trend is consistent with other studies of CO binding
`to metal atoms.12
`At the HF level of theory the BeBe distance in (CO)2BeBe-
`(CO)2 is only 1.938 Á, which is considerably shorter than the
`experimental bond distance of 2.45 Á in Be2,13 the shortest Be-Be
`distance of 2.226 Á in Be metal,14 and the Be-Be bond length
`of 2.124 Á in HBeBeH at the same level of theory.15
`The
`computed harmonic vibrational frequency for the Be-Be stretch
`for this molecule of 942 cm"1 is more than 4 times larger than
`the experimental value of the vibrational frequency of Be2 at 223
`cm"1 and is also larger than the Be-Be stretching frequency of
`645 cm"1 in HBeBeH at the same level of theory. The barrier
`for rotation about the Be-Be bond, computed at the HF geometry
`of the ground state by the GVB method16 correlating only the pair
`of electrons in the bond, is 6.70 kcal. The computed properties
`of the Be-Be bond, bond distance, bond energy, vibrational
`stretching frequency, and barrier to rotation, all lend support to
`the idea of a double bond between the Be atoms in this molecule.
`The nature of the bonding and the stability of (CO)2BeBe(CO)2
`demonstrate that it is indeed possible to form strong double bonds
`between Be atoms. The bonding in this molecule where two
`formally s2 closed subshells interact arises from the mixing of the
`
`(10) A detailed account of the electron correlation effect on the geometries
`of the singlet and triplet states, binding energy, and the singlet-triplet energy
`separation will be published separately.
`(11) Harrison, R. J.; Handy, N. C. Chem. Phys. Lett. 1986,123, 321 and
`references therein.
`(12) (a) Balaji, V.; Sunil, K. K.; Jordan, K. D. Chem. Phys. Lett. 1987,
`(b) Blomberg, M. R. A.; Brandemark, U. B.; Johansson, J.; Si-
`136, 309.
`eghbahn, P. E. M.; Wennerberg, J. J. Chem. Phys. 1988, 88, 2344.
`(c)
`Blomberg, M. R. A.; Brandemark, U. B.; Siegbahn, P. E. M.; Wennerberg,
`J.; Bauschlicher, C. W. J. Am. Chem. Soc. 1988, 110, 6650.
`(13) Bondybey, V. E. Chem. Phys. Lett. 1984, 109, 436.
`(14) Donahue, J. Structure of Elements; John Wiley: New York, 1974.
`(15) Sana, M.; Leroy, G. Theor. Chem. Acta 1990, 77, 383.
`Ill
`In Modern Theoretical
`(16) Bobrowitz, F. W.; Goddard, W. A.,
`Chemistry: Methods of Electronic Structure Theory; Scheafer, H. F., Ill,
`Ed.; Plenum: New York, 1977; Vol. 3.
`
`It is a pleasure to acknowledge Profs. K. D.
`Acknowledgment.
`Jordan, L. C. Allen, and J. M. Schulman and Dr. V. Balaji for
`many insightful discussions and Ms. D. M. Corsi for critical
`comments on the manuscript.
`
`Proton Affinities of the 20 Common -Amino Acids*
`Greg S. Gorman, J. Paul Speir, Cheryl A. Turner, and
`I. Jonathan Amster*
`Department of Chemistry, University of Georgia
`Athens, Georgia 30602
`Received December 27, 1991
`We report the first measurement of the gas-phase basicities
`(-AG of protonation) of all of the 20 common -amino acids, from
`which we can derive their proton affinities (- /7 of protonation).
`The basicities are determined by observing the occurrence
`or
`of reaction 1, and its reverse reaction, in which
`nonoccurrence
`A represents an amino acid, B a reference base, and AH"1" and
`BH+ their respective protonated forms. The observation of proton
`A + BH+ — AH"1" + B
`(1)
`transfer between the protonated molecule of a base and an amino
`acid implies a negative free energy for reaction 1. The reverse
`reaction is observed if the free energy of reaction 1 is positive.
`The -amino acids are low-volatility compounds and thus are
`difficult subjects for the type of gas-phase equilibrium experiments
`generally utilized to measure gas-phase basicity or proton affinity.1
`To date, values of the proton affinity of only 6 of the 20 common
` -amino acids have been published.2·3 Meot-Ner et al. used
`high-pressure mass spectrometry to measure the gas-phase basicity
`of six amino acids (glycine, alanine, valine, leucine, phenylalanine,
`and proline), and they also used variable-temperature studies to
`directly measure the enthalpy of the proton-transfer reaction and
`thus the proton affinities of three of the amino acids.2 Locke and
`Mclver measured the gas-phase basicities of glycine and alanine
`with ion cyclotron resonance spectrometry.3 In these studies, for
`the cases where enthalpies were not measured directly, proton
`affinity values were obtained from the gas-phase basicity data by
`estimating the entropy of the reaction, which was found to be a
`small or negligible correction.
`In this study, we use Fourier transform ion cyclotron resonance
`spectrometry to observe the reaction of laser-desorbed, neutral
`amino acid molecules with a series of protonated reference bases,
`as in reaction 1. With this technique, called laser desorption/
`chemical ionization (LD/CI), low-volatility or nonvolatile com-
`pounds can be utilized as the neutral partner in studies of ion-
`molecule reactions.4 The reverse of reaction 1 is examined by
`forming a protonated molecule of an amino acid by matrix-assisted
`laser desorption of the amino acid,5 ****followed by reaction with a
`• Author to whom correspondence should be addressed.
`* Dedicated to Professor Fred W. McLafferty on the occasion of his re-
`tirement.
`(1) The experimental methods for making equilibrium gas-phase basicity
`measurements have been reviewed by Aue, D. H.; Bowers, . T. In Gas Phase
`Ion Chemistry: Bowers, . T., Ed.; Academic Press: New York, 1979; Vol.
`2, Chapter 9.
`(2) Meot-Ner, M.; Hunter, E. P.; Field, F. H. J. Am. Chem. Soc. 1979,
`101, 686-689.
`(3) Locke, M. J.; Mclver, R. T., Jr. J. Am. Chem. Soc. 1983, 105,
`4226-4232.
`(4) Amster, I. J.; Land, D. P.; Hemminger, J. C; Mclver, R. T., Jr. Anal.
`Chem. 1989,61, 184-186. Speir, J. P.; Gorman, G. S.; Cornett, D. S.; Amster,
`I. J. Anal. Chem. 1991, 63, 65-69. Speir, J. P.; Gorman, G. S.; Amster, I.
`J. In Laser Ablation, Mechanisms and Applications: Miller, J. C., Haglund,
`R. F., Eds.; Springer-Verlag: Berlin, 1991; pp 174-179.
`(5) Karas, M.; Hillenkamp, F. Anal. Chem. 1988, 60, 2299-2301. Beavis,
`R. C.; Chait, B. T. Rapid Commun. Mass Spectrom. 1989, 3, 432-435.
`Hillenkamp, F.; Karas, M.; Beavis, R. C; Chait, B. T. Anal. Chem. 1991, 63,
`1193A-1202A.
`
`0002-7863/92/1514-3986$03.00/0
`
`© 1992 American Chemical Society
`
`See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
`
`Downloaded via UNIV OF CALIFORNIA LOS ANGELES on July 11, 2018 at 20:36:33 (UTC).
`
`
`
`-3986-
`
`
`
`
`MAIA Exhibit 1026
`MAIA V. BRACCO
`IPR PETITION
`
`

`

`Communications to the Editor
`
`J. Am. Chem. Soc., Vol. 114, No. 10, 1992
`
`3987
`
`Table I. Gas-Phase Basicities (GB) and Proton Affinities (PA) of
`the 20 a-Amino Acids (italics) and Reference Bases
`GB (350 K)“·6
`reference base
`amino acids
`200.7 ± 1.5
`acetophenone
`202.4 ± 3.2
`204.1 ± 1.5
`205.4 ± 2.8
`206.6 ± 1.5
`209.2 ± 1.5
`210.3 ± 2.6
`211.3 ± 1.5
`212.3 ± 2.5
`213.2 ± 1.5
`214.8 ± 3.1
`216.3 ± 1.5
`217.4 ± 2.6
`218.4 ± 1.5
`219.5 ± 2.6
`
`pyrrole
`
`2-fluoropyridine
`methylamine
`
`3-fluoropyridine
`
`ethylamine
`
`isopropylamine
`
`pyridine
`
`PA“C
`210.0 ± 1.5
`211.7 ± 3.2
`213.4 ± 1.5
`214.6 ± 2.7
`215.7 ± 1.5
`219.6 ± 1.5
`220.2 ± 2.1
`220.7 ± 1.5
`222.1 ± 2.9
`223.4 ± 1.5
`225.0 ± 3.1
`226.5 ± 1.5
`227.2 ± 2.2
`227.8 ± 1.5
`228.5 ± 2.2
`
`Gly
`
`Cys
`
`Ser, Asp
`Ala, Val
`Leu, lie
`
`Phe, Tyr, Asn
`
`Thr, Met, Gin,
`His
`
`Pro, Trp
`
`Glu
`
`Lys
`
`terf-butylamine
`
`trimethylamine
`
`diethylamine''
`
`di-n-propylamine1'
`triethylamine*'
`tri-n-butylamine‘,
`
`229.2 ± 1.5
`220.5 ± 1.5
`231.0 ± 3.3
`221.9 ± 2.9
`232.7 ± 1.5
`223.3 ± 1.5
`240.6 ± 1.9
`223.7 ± 1.9
`233.5 ± 1.5
`224.1 ± 1.5
`242.6 ± 3.4
`226.0 ± 3.4
`237.3 ± 1.5
`227.9 ± 1.5
`239.6 ± 1.5
`230.2 ± 1.5
`233.8 ± 1.5
`243.2
`>243.2
`>233.8
`Arg
`“All values are in kcal/mol. 6Relative GB values of the reference
`bases are from ref 7 and are adjusted for 350 K. The absolute GB
`values are based on a value of 198.2 kcal/mol for NH3 at 350 K. CPA
`values of the reference bases are from ref 7.
`* Relative GB and PA
`values are from ref 1 and are referenced to the values of GB and PA of
`trimethylamine from ref 7.
`
`neutral volatile base.6 Use of a series of reference bases with
`small differences in gas-phase basicity allows an amino acid to
`be bracketed within a narrow range. Most of the reference bases
`used in this study were chosen from a recently published ladder
`of proton affinities and gas-phase basicities.7 The study that
`produced the revised ladder of basicities eliminated systematic
`in the upper end of the proton affinity scale previously in
`errors
`use8 and yielded measurements in agreement with proton affinity
`values obtained from absolute determinations at various positions
`in the ladder.
`The gas-phase basicities and proton affinities measured for the
`20 -amino acids are listed in Table I, along with the literature
`values for the reference bases used in this study. For most of the
`amino acids, the entropy of reaction 1 can be calculated from
`changes in the rotational symmetry numbers ( ) of the reactants
`and products according to eq 2.9 The amino acids and reference
`AS,0i = R In [ +/ + ]
`(2)
`bases used in this study have low symmetry, yielding rotational
`entropy values of approximately ±2 cal/mol-K. At the temper-
`ature of the system in this study, 350 K,10 *the contribution of
`entropy to the free energy of reaction 1 is ±0.7 kcal/mol. The
`uncertainty in the assignment of the gas-phase basicities of the
`amino acids, determined by the spacings between the gas-phase
`
`(6) For formation of protonated molecules by matrix-assisted laser de-
`sorption, amino acids were mixed 1:1000 with sinapinic acid and desorbed with
`the 248-nm line of an excimer laser. The desorbed ions were thermalized by
`20-100 collisions with argon, admitted through a pulsed valve, prior to their
`reaction with the reference bases.
`(7) Meot-Ner, M.; Sieck, W. J. Am. Chem. Soc. 1991, 113, 4448-4460.
`(8) Lias, S. G.; Liebman, J. F.; Levin, R. D. J. Phys. Chem. Ref. Data
`1984, 13, 695-808.
`(9) Kebarle, P. Anna. Rev. Phys. Chem. 1977, 28, 445-476.
`(10) The cell temperature was determined by measuring the reaction rate
`of Ar+ with Na, suggested as a “thermometer” reaction by Bartmess, J. E. In
`Structure / Reactivity and Thermochemistry of Ions·, Ausloos, P., Lias, S. G.,
`Eds.; Reidel: Dordrecht, The Netherlands, 1987; pp 367-371. The kinetic
`energy dependence of this reaction has been reported by Dotan, I.; Lindinger,
`W. J. Chem. Phys. 1982, 76, 4972-4977.
`
`Table II. A Comparison of the Proton Affinity Values of Six
`a-Amino Acids with Published Values"
`
`Locke and
`Mclver**
`210.3
`213.1
`
`Meot-Nerc
`current study6
`204.0-207.4
`208.2
`glycine
`alanine
`211.5-214.0
`212.2
`211.5-214.0
`valine
`213.9
`leucine
`214.0-216.5
`214.5
`216.5-218.1
`215.1
`phenylalanine
`218.8-222.1
`218.4
`proline
`“All values are in kcal/mol. 6Values from this study have been ad-
`justed to match the proton affinity ladder from ref 9. CPA values are
`from ref 2 and are based on the proton affinity ladder of ref 9.
`8 PA
`values are from ref 3 and are adjusted to match the proton affinity
`ladder from ref 9.
`
`basicities of the reference bases, is ±2.5 kcal/mol on average. The
`contribution of entropy, caused by changes in rotational symmetry,
`to the free energy of reaction 1
`is much smaller than the un-
`certainty in the assignment of the free energy and can be neglected,
`so that AH = AG for reaction 1. This assumption has been verified
`in a previous study of amino acid proton affinities.2 A notable
`exception to this simplification occurs when an intramolecular
`hydrogen-bonded cyclization occurs, as has been observed for
` , -diamines.11 Lysine (2-carboxy-1,5-pentanediamine) fulfills
`the geometric requirement for such a cyclization. The formation
`of the hydrogen bond in cyclized diamines causes a substantial
`increase in their gas-phase basicities and proton affinities relative
`to monoamines of similar polarizability. The difference in gas-
`phase basicity between lysine and the structurally similar mo-
`noamine, leucine (2-carboxyisopentylamine), 11.2 ± 3.2 kcal/mol,
`is approximately the same as that between 1,5-pentanediamine
`and n-pentylamine, 10.7 ± 1.5 kcal/mol,1 suggesting the intra-
`molecular cyclization of protonated lysine. The entropy of cy-
`clization for lysine should be approximately equal
`to that of
`-20.5 cal/mol-K).1 The enthalpy
`1,5-pentanediamine (AScyc =
`of protonation of lysine via reaction 1 is equal to the free energy
`-7.2 kcal/mol at 350 K. The proton affinity value
`plus TAS =
`of lysine in Table I has been adjusted to account for cyclization
`of the protonated molecule.12
`The difference in gas-phase basicities between glutamic acid
`and aspartic acid (
`= 13 kcal/mol) is much greater than for
`the corresponding amides, glutamine and asparagine (
`= 2
`kcal/mol). The structures of these amino acids differ by one
`methylene group in the side chain, and so only a small difference
`in gas-phase basicity is expected, as is observed for the amides.
`The anomalously high gas-phase basicity of glutamic acid suggests
`cyclization of its protonated form. The proton affinity assignment
`has been adjusted to account for the entropy of cyclization ( 56 0
`-21.5 cal/mol-K, as for 4-hydroxybutanamine,1 TAScyc = -7.5
`Interestingly, the gas-phase basicity of glutamine does
`kcal/mol).
`not suggest cyclization of that amino acid.
`Arginine is the most basic of the amino acids, presumably
`because of the strongly basic guanido group present in its side
`It is more basic than any of the reference bases that are
`chain.
`used in this study. As a result, we can only set a lower limit on
`the gas-phase basicity of this amino acid, 234 kcal/mol, and on
`the proton affinity, 243 kcal/mol. Although arginine has the
`
`=
`
`(11) Aue, D. H.; Webb, J. M.; Bowers, . T. J. Am. Chem. Soc. 1973,
`95, 2699-2701. Yamdagni, R.; Kebarle, P. J. Am. Chem. Soc. 1973, 95,
`3504-3510. Meot-Ner, M.; Hamlet, P.; Hunter, E. P.; Field, F. H. J. Am.
`Chem. Soc. 1980, 102, 6395-6399.
`(12) The adjustment of the proton affinity values is valid only if endo-
`thermic but exoergic reactions are rapid compared to the ion-molecule col-
`lision rate, as has been observed in high-pressure mass spectrometry mea-
`surements; see ref 11, Meot-Ner et al. There is controversy as to whether
`endothermic processes can be driven by entropy in low-pressure experiments,
`as has been discussed by Henchman, M. In Structure j Reactivity and Ther-
`mochemistry of Ions· Ausloos, P„ Lias, S. G., Eds.; Reidel: Dordrecht, The
`Netherlands, 1987; pp 381-399. Although ions are detected under low-
`pressure, collisionless conditions (1CT9 Torr) in the experiments reported here,
`the ion-molecule reactions occur under multiple collision conditions, for which
`the reaction entropy must be considered. For example, protonated amino acid
`molecules undergo 20-100 collisions with a reference base.
`
`
`
`-3987-
`
`
`
`
`

`

`kH
`
`3988
`
`J. Am. Chem. Soc. 1992, 114, 3988-3989
`appropriate geometry to form an intramolecular hydrogen bond,
`there is insufficient data on the gas-phase basicities of guanidine
`compounds to determine whether cyclization occurs for this amino
`acid on the basis of the measurements reported here.
`Table II compares prior measurements of the proton affinities
`of six amino acids with the measurements reported here. To
`facilitate the comparison, the proton affinity values in the table
`are referenced to the proton affinity ladder used in the earliest
`study.2 We find excellent agreement with Meot-Ner’s mea-
`surements of the proton affinities of alanine, valine, leucine, and
`proline and close agreement for glycine and phenylalanine.2
`Likewise, we find excellent agreement with the proton affinity
`value of alanine reported by Locke and Mclver.3 Our value for
`the proton affinity of glycine is lower than, but similar to, that
`measured by Locke and Mclver. The comparisons show no
`systematic differences between ours and prior measurements. Our
`relative ordering of the proton affinities of the amino acids agrees
`with that proposed by Bojensen with a few notable exceptions.13
`Bojensen places histidine as the second most basic amino acid,
`while we find it sixth on our
`In Bojensen’s study, glutamine
`list.
`is more basic than glutamic acid, opposite to our
`finding. We
`find alanine, methionine, and threonine to be more basic than does
`Bojensen. While the exact reason for the differences between the
`two measurements is not known, it should be emphasized that the
`measurements reported here are based on well-established ion-
`molecule techniques and theory. Bojensen relates ion abundances
`in a metastable ion mass spectrum to thermodynamic properties
`of the product ions, a technique proposed by Cooks and co-workers
`for measuring relative proton affinities of closely related, mono-
`functional compounds.14 The discrepancies between our mea-
`surements and those of Bojensen may be a limitation of Cook’s
`method, which has never been tested with structurally diverse,
`polyfunctional molecules such as the amino acids.
`Acknowledgment. This work was supported by the National
`Science Foundation (CHE-9024922). Acknowledgement is made
`to the donors of the Petroleum Research Fund, administered by
`the American Chemical Society, for partial support of this work.
`
`~
`
`Table I. Rate Constants0 for the Reaction of Muonium Atoms with
`Aromatic N-Heterocyclic Solutes in Water at ~295 K. Comparison
`with Published Data on ‘H4
`WC
`solute
`*m/*h
`fcH/N
`33 ± 3
`benzene
`5.5
`(1.5)'
`9
`3.7
`58 ± 4
`pyridine
`7.4
`7.84
`7.8
`12
`50 ± 3
`2.7'
`12.5
`pyridazine
`1.4
`18.5
`37 ± 2
`0.46
`40.2
`pyrimidine
`0.927
`9
`77 ± 5
`3.0*
`25.7
`pyrazine
`19
`1.5
`"All k values are in units of 10s dm3 mol'1 s'1; kM/C and fcH/N are
`See footnotes c to g for pH dependence.
`described in the text.
`' Calculated as kH per C atom. 4The 7.8 refers to nat-
`4 Reference 8.
`ural pH. At pH ~1, however, fcH =
`1.7 X 108 M'1 s'1 for pyridine
`where >99% of the solutes were protonated; so protonation decreases
`kH by ~4.5-fold. This contrasts with Mu, where kM increased from 58
`to 74 (X108) on changing from pH ~7 to pH 1.2, for pyridine. Thus
`kM/kH increases from 7.4 at pH ~7 to 44 at pH ~1.
`'kH measured
`1, where ~99% of the solutes have one of their N atoms
`at pH ~
`protonated; so the number of unprotonated N atoms is half that rele-
`fkH measured at pH ~1, where ~65% of the
`to the fcM data.
`vant
`solutes have one of their N atoms protonated.
`* kH measured at pH
`1, where ~30% of the solutes have one of their N atoms protonated.
`study it shows a different type of reaction.
`Recently we found that Mu produces the free radical which
`arises from addition into the aromatic ring of pyrazine at its C
`atoms.4 Now we report the rate constants measured for that
`reaction and Mu’s reaction with other N-heterocyclic compounds.
`Table I gives our observed rate constants (kM) for reaction of Mu
`with benzene, pyridine, and the 1,2-, 1,3-, and 1,4-diazines. These
`data were obtained by measuring, the chemical decay rate of
`muonium using the muon-spin-rotation technique5 on millimolar
`aqueous solutions of these compounds at natural pH. They are
`compared in Table I with the published data for reaction of ‘H
`atoms with the same solutes under similar conditions. Since Mu
`adds to a ring C,4 whereas
`attaches at the more electronegative
`N atoms,6 these data have been normalized in columns 3 and 5
`of Table I by dividing fcM by the number of C sites on the solutes
`and kH by the number of N sites.
`Mu is seen to react 3.7 times faster than with benzene. This
`is not inconsistent with Mu’s 3-fold-higher mean thermal velocity
`stemming from its one-ninth atomic mass:
`a kinetic isotope effect
`expected for diffusion-limited reactions.3 The overall rate constants
`ratio, kM/kH, then increases to 7.4 due to the presence of one ring
`N, and up to 40 with two N atoms. This ratio rises partly by ku
`increasing and partly by kH decreasing.
`The enhancement of kM by ring nitrogens implies that Mu is
`“nucleophilic”, because N draws electron density from the C atoms
`where Mu reacts. Preliminary results even suggest that there is
`a direct proportionality between -log (fcM) and the Hückel mo-
`lecular orbital localization energy on each C.7 By contrast, the
`presence of the ring N of pyridine seems to switch H’s reaction
`site from C to N (as seen through ESR data6). This results in
`a minor overall decrease in kH (but an increase per atom where
`reaction takes place). The presence of a second N atom then
`reduces the rate further. Also,
`has already been shown to
`display a negative Hammett p parameter,1·2 and the differences
`between the diazines have been interpreted as consistent with this
`electrophilic character.1
`information to be gleaned from the pH
`There is additional
`dependence of these rate constants arising from protonation of
`the solute. For pyridine, which is the only N-heterocycle for which
`data are available, kM/kH jumps from 7.4 at pH ~7 to 44 at pH
`1. Almost all of this jump (see footnotes to Table I) emerges
`from a decrease in kH. This consequently corroborates the view
`that reaction of H, but not Mu, occurs on the N atom.
`
`(13) Bojensen, G. J. Am. Chem. Soc. 1987, 109, 5557-5558.
`(14) McLuckey, S. A.; Cameron, D.; Cooks, R. G. J. Am. Chem. Soc.
`1981, 103, 1313-1317.
`
`Different Reaction Paths Taken by Hydrogen Isotopes
`Zhennan Wu, John M. Stadlbauer,7 8and David C. Walker*
`Chemistry Department and TRIUMF
`University of British Columbia
`Vancouver, V6T 1Y6 Canada
`Received September 6, 1991
`Kinetic isotope effects usually reveal differences in degree but
`not in kind. However, in studying the chemistry of muonium
`(hydrogen’s light isotope with a positive muon as nucleus) we find
`it to behave as a nucleophile in its reactions with aromatic N-
`heterocyclic solutes in water. This contrasts with the electrophilic
`character of .1,2 Muonium (Mu) has one-ninth the atomic mass
`of but virtually the same reduced mass, ionization energy, size,
`etc. as all hydrogen isotopes. Hitherto Mu has shown kinetic and
`spectroscopic isotopic effects which are understandable in terms
`of its different atomic mass from that of ;3 but in the present
`
`’ Permanent address: Chemistry Department, Hood College, Frederick,
`MD 21701.
`(1) (a) Neta, P.; Schuler, R. H. J Am. Chem. Soc. 1972, 94, 1056.
`(b)
`Neta, P. Chem. Rev. 1972, 72, 533.
`(2) Pryor, W. A.; Lin, T. H.; Stanley, J. P.; Henderson, R. W. J. Am.
`Chem. Soc. 1973, 95, 6993.
`(3) (a) Walker, D. C. Muon and Muonium Chemistry·, Cambridge
`(b) Walker, D. C. Can. J. Chem. 1990,
`University Press: Cambridge, 1983.
`86, 1719.
`
`~
`
`(4) Wu, Z.; Barnabas, . V.; Stadlbauer, J. M.; Venkateswaran, K.;
`Porter, G. B.; Walker, D. C. J. Am. Chem. Soc. 1991, 113, 9096.
`(5) Cox, S. F. J. J. Phys. C 1987, 20, 3187.
`(6) Barton, B. L.; Fraenkel, G. F. J. Chem. Phys. 1964, 41, 1455.
`(7) Z. Wu et al., preliminary unpublished data.
`(8) Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Natl.
`Stand. Ref. Data Ser. (U.S., Natl. Bur. Stand.) 1988, NSRDS-NBS.
`
`0002-7863/92/1514-3988S03.00/0
`
`© 1992 American Chemical Society
`
`
`
`-3988-
`
`
`
`
`

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