You are on page 1of 3

Notes 433

2 shows the spectrum, which is attributed to the Experimental Section


CDS radical on the basis of the line intensities and Preparation of S,S-Dideuterio-^,4-pentanedione, 3,3-
splitting constant. Comparison of Figures 1 and 2 Dideuterio-2,4-pentanedione was prepared by stirring
shows that CD3 formation by solvent cleavage cannot
·

2,4-pentanedione (Eastman Organic Chemicals) with


account for any significant portion of the central line excess deuterium oxide (99.7 mol % D20), extracting
of Figure 1. with methylene chloride, and finally fractionally dis-
4. Summary and Discussion. The esr spectrum of tilling the deuterated product. This exchange process
protonated methyl radical in a glass of TMPD and was repeated three times.
perdeuterated 3MP at 77°K is shown to arise from the Anal. Caled for CsH«D202: C, 58.80; total hydro-
ß scission of TMPD, with a biphotonic mechanism in- gen, 9.87. Found: C, 58.60, 58.64; total hydrogen
dicated. It is of interest to note that the CH3 signal ap-
·

assuming H:D =
3:1, 9.75, 10.1. Analysis was by
pears only after prolonged irradiation, but the trapped Micro-Tech Laboratories, Inc., Skokie, III.
electron (seen at a much lower microwave power) The purity of the product was confirmed by nmr,
appears after only a few minutes. Since ß scission does uv, ir,3 and mass spectra.9 Substitution of deuterium
occur in this perdeuterated glass, this pathway should at the 3 position of 2,4-pentanedione was ~93% as
also be viable in the protonated systems. Whether or shown by peak height measurements of the nmr signals
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

not it is the dominant mechanism for producing the from the methylene protons of 2,4-pentanedione and
methyl radical fragment in the protonated systems re- 3,3-dideuterio-2,4-pentanedione, under similar instru-
mains to be established. A similar study of perdeu- ment settings. The hydroxyl group and the vinyl
Downloaded via UNIV DE VALLADOLID on May 2, 2019 at 09:07:21 (UTC).

terated TMPD in protonated 3-MP could greatly carbon have the same per cent of protium within the
clarify this question. experimental error of nmr measurements.
No attempt was made to increase the methylene
Acknowledgments. The authors wish to acknowledge
deuterium content to the theoretical limit since approxi-
the assistance of Richard Kornfield in obtaining the
mass spectrum of the perdeuterated 3-MP. mately 2% deuterium substitution into the terminal
methyl groups was observed by nmr after the last ex-
change reported above.
Spectra. The keto-enol equilibrium constants were
determined by electronically integrating, with a Varían
Keto-Enol Equilibria in 2,4-Pentanedione and Associates, Inc., A-60 spectrometer, the nmr signals of
the terminal methyl groups of the two tautomers.
3,3-Dideuterio-2,4-pentanedione
Integrations were optimized according to the guidelines
of Jungnickel and Forbes.10 Uv data were obtained
by David W. Thompson and A. L. Allred* with a Cary Model 14 spectrophotometer.
Department of Chemistry and Materials Research Center,
Northwestern University, Evanston, Illinois 80301 Results
(Received April 8, 1989) The equilibrium and thermodynamic constants
Publication costs borne completely by The Journal of Physical Chemistry derived from nmr data fot 2,4-pentanedione and 3,3-
dideuterio-2,4-pentanedione are summarized in
Table I.
The goal of this research was to investigate hydrogen The ultraviolet spectrum of 2,4-pentanedione exhibits
isotope effects on intramolecular hydrogen bonds. The a maximum at 271 µ due to the enol tautomer. Sub-
per cent of each enol tautomer, apparently containing stitution of deuterium into the 3 position does not alter
intramolecular hydrogen bonds, in D20 solutions of 3- noticeably either the position or the shape of the enol
deuterio-3-methyl-2,4-pentanedione1 ·2 and 1-deuterio-
2-acetylcyclohexanone8 is less than the per cent enol F. C. Nachod, Z. Phys. Chem., Abt. A, 182, 193 (1938).
(1)
tautomer present in the corresponding completely pro-
(2) F. A. Long and D. Watson, J. Chem. Soc., 2019 (1958).
tonated compound in H20. Values of acid ionization
(3) T. Riley and F. A. Long, J. Amer. Chem. Soc., 84, 522 (1962).
constants show that there is less intramolecular hydro- (4) G. Dahlgren, Jr., and F. A. Long, ibid., 82, 1303 (1960).
gen bonding for the deuterio maléate ion in D20 than for (5) C. J. Creswell and A. L. Allred, ibid., 84, 3966 (1962), and
the hydrogen maléate ion in Hs0.4 Interpretation of references therein.
the above results is difficult owing to the possibility of (6) (a) G. R. Plourde, Dies. Abstr., 22, 1400 (1961); (b) R. West,
personal communication.
different solvent effects from H20 and D20. Deute-
(7) D. W. Thompson, Ph.D. Thesis, Northwestern University,
rium, in several systems,6-7 forms stronger intermolecular 1968.
hydrogen bonds than protium. In this research, keto- (8) H. Ogoshi and K. Nakamoto, J. Chem. Phys., 45, 3113 (1966).
enol equilibria were measured by nmr for pure /3-dike- (9) “Catalog of Mass Spectral Data,” American Petroleum Insti-
tute, Project No. 44, Carnegie Institute of Technology, Pittsburgh,
tones and 0.5 mole fraction solutions in cyclohexane and Pa.
by uv spectroscopy for dilute solutions in cyclohexane. (10) J. L. Jungnickel and J. W. Forbes, Anal. Chem., 35, 938 (1963).

The Journal of Physical Chemistry, Vol. 75, No. 3, 1971


434 Notes

Table I: Equilibrium and Thermodynamic Constants for Enolization of


2,4-Pentanedione and 3,3-Dideuterio-2,4-pentanedione from Nmr Measurements

AS,
AH, cal mol”1
% keto K T, °c koal mol-1 deg”1

O 0

ch8cch2cch8 21.8 ± 0.3“ 3.59 37.3 -2.4 ± 0.2s -5.2 ± 0.6s


(pure) 13.6 ± 0.5 6.35 2.5
10.5 ±0.3 8.52 -19.0
O 0

ch8ccd>cch, 19.1 ±0.4 4.24 37.3 -0.10 ±0.04 2.6 ± 0.2


(pure) 18.5 ± 0.3 4.41 2.5
18.6 ±0.3 4.38 -19.0
O 0
II II

ch8cch2cch8 10.4 ± 0.2 8.62 37.3 -1.8 ±0.1 -1.4 ± 0.2


. 5 mole fraction in CeHu) 7.4 ± 0.3 12.51 0.0
5.8 ± ~1 16.2 -18.5
O O
II II
1! ¡1

CHsCCDeCCH, 11.0 ±0.2 8.09 37.3 0.09 ± 0.12 4.5 ± 0.4


. 5 mole fraction in CeHu) 10.7 ± 0.3 8.34 0.0
11.4 ±0.3 7.77 -18.5
1

Standard deviation. 6
Probable error.

absorption band. With the assumption that the molar ated enol tautomer being less ordered as a consequence
absorptivity for 2,4-pentanedione does not change upon of less intramolecular hydrogen bonding. Both inter-
substitution of deuterium into the 3 position, uv data at molecular hydrogen bonding and the existence of some
25° for dilute (co. 10~6 M) cyclohexane solutions of 2,4- nonhydrogen-bonded enol tautomer would make AS
pentanedione and 3,3-dideuterio-2,4-pentanedione un- more positive. Free hydroxyl groups in the gas phase
ambiguously show that Ku is significantly larger than and in carbon tetrachloride solutions have been ob-
KD. served by ir.11 Evidence for molecular association
comes from both nmr spectra and cryoscopie measure-
Discussion
ments. Rogers and Burdett16 concluded from the
The magnitudes, Table I, of AHenotation for both pure chemical shifts of 2,4-pentanedione in various concen-
2,4-pentanedione and a 50:50 2,4-pentanedione-cyclo- trations of cyclohexane and from temperature studies
hexane solution are much more negative than those of that intermolecular association is important. Our
the deuterated analog. As reported earlier,11 the in- chemical shift measurements confirm the work of Rogers
tensity of the 1712 cm-1 peak, due to carbonyl stretch- and Burdett and furthermore show sigmoid variations
ing, in the infrared spectrum of 2,4-pentanedione in at concentrations below 0.2 mole fractions, thus prob-
cyclohexane is considerably smaller than for the deu- ably explaining why others12·13 reported no chemical
terium analog, Also, uv studies of dilute solutions, shifts for this system. Freezing point depressions of
co. 10-8 M, of 2,4-pentanedione in cyclohexane at
2,4-pentanedione-cyclohexane solutions are consistent
25° show that the enol content is significantly larger with intermolecular association. For example, the
than in a comparable solution of 3,3-dideuterio-2,4-
apparent molecular weight of 2,4-pentanedione is 105
pentanedione. If stabilization of the enol tautomer in 0.10 m solution and 113 in 0.18 m solution, whereas
by intramolecular hydrogen bonding is an important the true molecular weight is 100.
factor in determining the overall enthalpy of enolization, The large differences in enol-keto ratios and thermo-
protium apparently forms a stronger hydrogen bond dynamic constants for 2,4-pentanedione and 3,3-dideu-
in this system than does deuterium.
terio-2,4-pentanedione are difficult to explain. Isotope
For the normal protium-containing compound, both
pure and in cyclohexane solutions, the AH and AS val- (11) S. Bratoz, D. Hadzi, and G. Rossmy, Trans. Faraday Soc., 52,
ues in Table I are comparable to other recently reported 464 (1956).
values.12-14 (12) L. W. Reeves, Can. J. Chem., 35, 1351 (1957).

The entropy of enolization is much more positive for (13) G. Allen and R. A. Dwek, J. Chem. Soc. B, 161 (1966).
(14) J. Kuo, Dies. Abstr. B, 27, 3064 (1967).
3,3-dideuterio-2,4-pentanedione than for 2,4-pentane- (15) . T. Rogers and J. L. Burdett, Can. J. Chem., 43, 1516
dione. This difference can be attributed to the deuter- (1965).

The Journal of Physical Chemistry, Vol. 76, No. 8, 1971


Notes 435

effects of the keto form are doubtless small and mainly Research Projects Agency of the Department of De-
confined by zero point energy differences of vibrations fense, through the Northwestern University Materials
involving the methylene group. The maximum dif- Research Center.
ference in bond dissociation energies for [Z?0(O-H) —

Eo(C-H)] -

[Eo(O-D) So(C-D)]16 is too small to


~

account for the differences in enthalpy values in Table


I. Conceivably, enolization of 3,3-dideuterio-2,4-pen- Comparative Ligand Field Studies of
tanedione is less favored as a consequence of the mean Manganese(II) Spectra
O-D bond length being less than that of OH, and there-
fore the hydrogen bond, -D-O, being longer and
by A. Mehra
weaker. Normal bonds, for example -O-D, involving
deuterium always are shorter than analogous protium College of Environmental Sciences, University of Wisconsin—Green
Bay, Green Bay, Wisconsin 54305 (Received May 18, 1970)
bonds due to anharmonic vibrations.17 Williams re-
Publication costs assisted by the University of Wisconsin—Green Bay
ported18 the following bond distances and angles from
an investigation by X-ray crystallography

2.468 A A comparative study of Mn2+ spectra in different


ligand coordinations is undertaken in this work. The
large number of bands in the spectrum of Mn2+ allows a
free variation of various parameters and provides a
good check on ligand field calculations. Spectra in
hydrated, fluoride, chloride, and bromide complexes
are investigated and regularities have been observed.
The energy levels from d6 configuration of Mn2+ and
the general features of observed spectra have been
reviewed earlier.1-3 The observed band positions in
several cases are given in Tables I and II. Several
schemes for fitting the band energies have been sug-
gested.4-8 One of the best is that of Heidt, Koster,
and Johnson4 using free variation of the three crystal
If this planar, chelated enol structure had involved field parameters B, C, and Dq. However, it is found
average17 bond lengths (rc_c = 1.537 A, rc=o = 1.335, that these three parameters are not enough to give
rC-o = 1.426, rC-o = 1.215) and bond angles of good agreement with all the band energies. A similar
either 120° or 105°, the 0-0 separation would be 2.50 case is encountered for the free ion where the use of
Á and the hydrogen atom would be >0.30 Á from the only two electrostatic parameters B and C does not give
0-0 line. The actual O-H distance above is ~0.22 Á good agreement for all the terms.9-11 Consequently,
greater than the average O-H distance17 in nonhydrogen several correction terms for fitting the free ion term
bonded compounds. Thus, formation of the enol ring values have been suggested. Among these, Trees’
apparently is accompanied by bond strains, and the correction term is found to be most important.10·11
small difference in O-H and O-D bond lengths may be Following its success for the free ion, we have included
unusually important. this term in our ligand field calculations. The signifi-
The data presented here indicate that acidity con- cance of this correction11 and the matrix elements in
stant data4 and keto-enol equilibrium data1-3 for
/3-diketones in light and heavy water are indeed best (1) N. 8. Hush and R. J. H. Hobbs, Progr. Inorg. Chem., 10, 259
explained in terms of protium forming a stronger intra- (1968).
molecular hydrogen bond than deuterium in these sys- (2) D. S. McClure, Solid State Phys., 9, 399 (1969).
tems and that solvent differences between light and (3) C. . Ballhausen, "Introduction to Ligand Field Theory,”
McGraw-Hill, New York, N. Y., 1962.
heavy water are not dominant. (4) L. J. Heidt, G. F. Koster, and A. M. Johnson, J. Amer. Chem.
Soc., 80, 6471 (1958).
Acknowledgments. D. W. T. expresses appreciation
(5) L. E. Orgel, J. Chem. Phys., 23, 1004 (1955).
for a predoctoral fellowship from the Division of General
(6) J. W. Stout, ibid., 31, 709 (1959).
Medical Sciences, United States Public Health Service. (7) W. Low and G. Rosengarten, J. Mol. Spectrosc., 12, 319 (1964).
This research was supported in part by the Advanced (8) A. Mehra and P. Venkateswarlu, J. Chem. Phys., 45, 3381
(1966).
(16) K. B. Wiberg, Chem. Rev., 55, 713 (1955). (9) J. C. Slater, “Quantum Theory of Atomic Structure," McGraw-
(17) L. E. Sutton, “Tables of Interatomic Distances and Configura- Hill, New York, N. Y., 1960.
tion in Molecules and Ions” (Special Publication No. 18), The (10) R. E. Trees, Phys. Rev., 83, 756 (1951); 84, 1089 (1951); and
Chemical Society, London, 1965. 85, 382 (1952).
(18) D. E. Williams, Acta CrystaUogr., 21, 340 (1966). (11) G. Racah, ibid., 85, 381 (1952).

The Journal of Physical Chemistry, Vol. 76, No. S, 1971

You might also like