The stereochemistry of protonation. XI - Journal of the American

Feltenberger , Arthur Nagel , Wendy Simon , Melani Suarez-Contreras , Norma J. Tom and Brian T. O'Neill ... Howard E. Zimmerman and Alexey Ignatch...
0 downloads 0 Views 796KB Size
6091 Table 111. Pmr Chemical Shifts for D-Biotin. and Its Triprotonated Form Solvent CP cs CaC7 CYC1OC11 FSOaH-SbFrS02 CFaCOOH

4.0-4.5



3.10

3.47

Cl 2

CO*H,+

OH

NH

12.28 12.54

9.64

7.04 7.21

5.60

1.7-2.6

3.29

4.81

1.73

2.62

a Chemical shifts in FS03H-SbF5-S02solution are given in parts per million from external TMS. internal TMS as standard.

!I

4

H H H

+

e

...

Shifts in CFCOOH are referred to

ureido ring. Both of these m o d e s of a t t a c k would be less favorable, on steric grounds, t h a n in D-biotin, a n d i n d e e d L-biotin shows no physiological activity. 28

OH

HN&+

...

PH

Experimental Section

: (CH~)&H~COOH~+ 9-11 12

XVIII Relatively strong interactions between the sulfur atom in D-biotin and t h e p r o t e i n to which it is a t t a c h e d h a v e been ~ u g g e s t e d . ~The ~ , ~ fact ~ that protonation of the sulfur occurs trans to t h e valeric acid side c h a i n suggests that b i n d i n g of t h e sulfur with t h e p r o t e i n should o c c u r from the s a m e d i r e ~ t i o n . ? ~A t t a c k from this side of the t h i o l a n e ring is particularly favorable in D-biotin in which the valeric a c i d side c h a i n is cis to t h e ring j u n c t i o n . In L-biotin, any interaction with the sulfur w o u l d h a v e to occur with a p p r o a c h from either the same side as the side c h a i n or the s a m e side as the (23) A . C. Mildvan and M. C. Scrutton, Biochemisfry, 6, 2978 (1967). (24) The steric requirements of the proton in such cases is a question that has not been completely resolved. In the protonation of ketonesP5 and cyclic ethers26 protonation from the least hindered side is always found, and our results accord with the view that in a conformational rivalry between the proton and nonbonding electrons, the final equilibrium position is primarily determined by the interaction properties of the ~roton.27 (23) G. A . Olah, M. Calin, and D. H. O’Brien, J . Amer. Chem. SOC., 89, 3586 (1967). (26) G. A. Olah, P. J. Szilagyi, and J. Sommer, to be published. (27) Footnote 17 in ref 17a.

1,l-Dimethylurea was prepared from nitrourea and dimethylamine, Methylisourea p-toluenesulfonate was prepared from urea and methyl p-toluenesulfonate. Other urea and guanidine bases used were commercially available and, where necessary, were recrystallized before use. Solutions of the bases in fluorosulfuric acid-antimony pentafluoride-sulfur dioxide were prepared as described earlier. Proton spectra were obtained using a Varian Associates Model HA-100 spectrometer with a capillary tube of TMS as the internal lock signal. Chemical shifts, referred to this standard, were obtained using a frequency counter, Temperature measurements were made by means of a thermometer constructed to fit into a 5-mm nmr tube and are considered accurate to within + 1’. 16N spectra were obtained by the indor method using the apparatus described previ~usly.~The reference used for determination of 16N shifts was a saturated aqueous solution of 16NH4NOafor which v E / v x was found to be 9.8686442. Rotational rates were computed at coalescence using the relationship

k = 117

r3’z

= K -

where Av is the chemical shift separation in hertz in the absence o exchange.

Acknowledgment. This work was possible through a grant of the N a t i o n a l Institutes of Health. Hoffman L a R o c h e , Inc., is t h a n k e d for a gift of D-biotin. (28) P . Gyorgy, Vifamins, 1, 527 (1954).

The Stereochemistry of Protonation. XI Howard E. Zimmerman and Patrick S. Mariano

Contribution f r o m the Chemistry Department of the University of Wisconsin, Madison, Wisconsin 53706. Received March 15, 1968 Abstract: The stereochemistry of ketonization of exocyclic cyclohexane enols and tautomerism of aci-nitro compounds was reinvestigated. The enol of l-acetyl-4-phenylcyclohexanewas found to afford predominantly the less stable cis isomer on ketonization. Similarly, l-aci-nitro-4-phenylcyclohexanetautomerized preferentially to cis-l-nitro-4-phenylcyclohexane when large proton donors were used. With smaller donors, the trans isomer was preferred. It was concluded that substituents a t carbon 2 are not necessary for kinetic control leading t o the less stable stereoisomer. The relationship of ground-state conformations to the transition states utilized is presently discussed. It is concluded, in keeping with our earlier postulate concerning kinetic control of ketonization, that approach of the proton donor in the least hindered way is usually the most important factor.

o m e y e a r s ago, w h e n t h e difference between kinetic and t h e r m o d y n a m i c control of reactions was just b e c o m i n g clear, we n o t e d t h a t many organic reactions

S

involved unstable enol intermediates and that stereo-

chemical preferences observed in these reactions are c o n t r o l l e d b y t h e kinetics of ketonization. O u r hypothesis, which seems to b e generally accepted, w a s that in many cases t h e less stable stereoisomeric product Zimmerman, Mariano

/ Ketonizing

Exocyclic Enols

6092

results from a kinetically controlled ketonization in which proton attack occurred from the less hindered direction. The cases initially discussed' and subsequently investigated2 included exocyclic six- and five-ring enols, endocyclic five-ring enols, and less common examples. Only a few of the molecules considered were six-ring enols containing a bulky substituent at C-2. Recently, Malhotra and suggested that our views on ketonization of 2-substituted exocyclic six-ring enols must be abandoned. The purpose of this report is to clarify the picture, put it in the proper perspective, and provide some recent results. The enols 1 of 1-benzoyl-2-phenylcyclohexane and the mi-nitro tautomer 2 of 1-nitro-2-phenylcyclohexane were postulated by Malhotra and Johnson to exist in 2-phenyl axial conformations l a and 2a due to strain.3b The predominance of cis product1,2e was rationalized with our hypothesis of less hindered attack, in this case axial, of the proton donor on conformations l a and 2a4 (see Chart I). Thus the Malhotra-Johnson Chart I. cis-1-Benzoyl-2-phenylcyclohexane and cis-1-Nitro-2-phenylcyclohexane Formation

le

2e'

A

,o-

H la

2a

b~

mechanism consists of application of our basic theory of less hindered approach to a different ground-state conformation. More important, however, are the transition states in proceeding from enol or mi-nitro tautomer to product. Provided that the activation energy for the ketonization reaction is large compared with the barrier to conformational inversion, the proportion of conformations reacting is not necessarily related to the population of the ground-state conformations; instead, the stereochemistry utilized depends on the relative transition-state energies.8 In the transition stateg (note Chart 11) (1) H. E. Zimmerman J . Org. Chem. 20, 549 (1955). (2) (a) H. E. Zimmerman, J . Am. Chem. Soc., 78, 1168 (1956); (b) H. E. Zimmerman and H. J. Giallombardo, ibid., 78, 6259 (1956); (c) H. E. Zimmerman, ibid., 79, 6554 (1957); (d) H. E . Zimmerman and T. W. Cutshall, ibid., 80, 2803 (1958); (e) H. E. Zimmerman and T. A. Nevins, ibid., 79,6559 (1957); (f) H. E. Zimmerman and T. W. Cutshall, ibid., 81, 4305 (1959); (9) H. E. Zimmerman and A . Mais, ibid., 81, 3644 (1959); (h) H. E. Zimmerman and W. H. Chang, ibid., 81, 3634 (1959). (3) (a) S. I