13
Indian Journal of Chemistry Vol. 42B, December 2003, pp. 3089-3101 Theoretical perspectives on the Cope rearrangement in bullvalene systems Pebam Munindro Singh & R H Duncan Lyngdoh* Department of Chemistry, North-Eastern Hill University, Shillong 79 3 022, In dia Received 6 August 2002; accepted (revised) 4 March 2003 Using the se mi -empiri ca l AM I SCF-MO method, transition states have been located For Cope rearrangements in vari- ous bullvalene systems, including degenerate unsubstituted and difluorosubstituted cases as well as non-dege nerate mon- ofluoro substituted cases . Transition states have been located for non-degenera te cases using the routine SADDLE keyword of the MOPAC package. By assuming transition state symme try abo ut the reac ti on coo rdinate for degenerate cases, compu- ta ti onal time is substantia ll y saved with no loss of essen ti al accuracy. All th ese tran si ti on states in corporate a boat-shaped cyc li c six-membered moiety. Ac ti va ti on energy barriers range between 24.0 and 32.5 kca llmol, rather hi gh co mpared w ith experimental values re poned , and thus systemati ca ll y co rrec ted. Th e non-degcneratc cases have transition states whose posi- tions along the reac ti on coordinate (as gauged by appropriate geo metry markers) agree well with their s li ghtly "late" charac- ter predicted from reac ti on endoth er micity by the Hammond postulate. Th e eFfects of fluorine substitution on ac ti va ti on en- ergy and trans iti on state geo metry as predicted by th ese ca lc ul a ti ons accord we ll with chemical intuition based on the induc- tive effect. The appli ca tion of a systema ti c correction to the activation barriers leads to more reasonable values. For the non- degenerate monofluorosubstituted cases, these co rrected activation barriers are used to construct an energy profile diagram to represent a ll the interco nversions, which leads us to inf er a minimal energy requirement of 19 .22 kcallmol for co mplete scrambling of th e ca rbons during Cope rea rrangement of th ese systems. The Cope rearrange ment l.2, long conside red as a typi- cal pericyclic reaction, is believed to be a one-step concerted process occun·ing via a cyclic transition state and abiding by the conservation of orb it al sy m- metrl. The thermal prerequisite and stereochemistry of this [3,3] sigmatropic rearrangement have been satisfactorily explained by frontier orbital analysis and the correlation diagram method 4 . 5 . This theoretical study examines the Cope rearra ngement in some bull- valene systems, including degener ate cases with reac- tant and product id entica l. Cope rearrangements in 1,5-hexadienes are r3 ,3 ] sigmatrop ic carbon -ca rbon bond migrations over a diene framework, subjected to intense experimental studl· 14 . Originally described as a "no mechanism" reaction 1 5, the Cope rearrangement was then labelled as a pericyclic concerted reaction via a cyclic aro- matic transition state. Later se mi-empirical theoretical work raised th e possibility of it being an asynchro- nous process involving biradicaloid intermediates l6 . More recently, accurate theoretical calculations l7 in- dicate a concerted process as a more correct descrip- ti on. Probably the actual situation depend s upon the particular system involved, including substituent and conformation al effects, as well as the theoretical method used. The degenerate Cope rean·a ngement is a special case where the reactant and product are identical, e.g. for unsubstituted or symmetrically substituted acyclic 1,5-h exa diene systems, besides complex systems like se mibullvalene and bullvalene involving va le nce tautomerism. Bullvalene, first synthesized by Schr6der l8 , involve s a Cope rea rr angeme nt (Figure 1) attracting much attention due to the phenomenon of comp l ete scrambling of its 10 carbon atoms by multi- ple Cope rearrangements 19 which distribute the pos i- tion of the cyclopropane ring among a ll th e carbons. At room temperature four 13C NMR signals are ob- tained as expected, but at about 100 D C they coalesce into only one signaI 2o . 22 , signifying indisting ui shabil- ity of all 10 carbo ns when the rapid rate of intercon- version between the 10 !/3 di fferent tautomers be- comes comparable to the NMR time scale . Th e mole- cule adopts a "fluxional" structure, giving only one sharp singlet in its proton NMR spectr um as well 23 . M h . I k 24 .39 . I . uc expenmenta wor , mall1 y uSin g spec- troscop/ O.22.30.39, has followed its discovery. Th eo - retical stud/ 9A I regarding the bullvalene Cope rear- rangement has been relatively less intense. Apart from the interesting feature of "fluxionality", we also note that, unlike the acyclic 1 ,5-hexadiene ca ses, bullva- lene possesses considerably less conformation al flexi- bility, thus restricting the scope for different reaction

Theoretical perspectives on the Cope rearrangement ...nopr.niscair.res.in/bitstream/123456789/21769/1/IJCB 42B(12) 3089 … · Indian Journal of Chemistry Vol. 42B, December 2003,

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Indian Journal of Chemistry Vol. 42B, December 2003, pp. 3089-3101

Theoretical perspectives on the Cope rearrangement in bullvalene systems

Pebam Munindro Singh & R H Duncan Lyngdoh*

Department of Chemistry, North-Eastern Hill University , Shillong 793 022 , India

Received 6 August 2002; accepted (revised) 4 March 2003

Using the semi-empirica l AM I SCF-MO method, transiti on states have been located For Cope rearrangements in var i­ous bullvalene sys tems, including degenerate unsubstituted and difluorosubstituted cases as well as non-degenerate mon­ofl uoro substituted cases. Transition states have been located for non-degenerate cases using the routine SADDLE keyword of the MOPAC package. By assuming trans ition state symmetry about the reac tion coordinate for degenerate cases, compu­tati onal time is substantiall y saved with no loss of essenti al accuracy. All these transi tion states incorporate a boat-shaped cyclic six-membered moiety . Acti vation energy barriers range between 24.0 and 32.5 kcallmol , rather hi gh compared with ex perimental values reponed, and thus sys te mati call y corrected. The non-degcneratc cases have transition states whose posi­tions along the reacti on coordinate (as gauged by appropri ate geometry markers) agree well with their s lightly " late" charac­ter predicted from reaction endothermici ty by the Hammond postulate. The eFfects of fluorine substitution on ac ti vati on en­ergy and transition state geometry as predicted by these calculations accord well with che mical intuition based on the induc­tive effect. The appli cation of a systematic correction to the activation barriers leads to more reasonable values. For the non­degenerate monofluorosubstituted cases, these corrected activation barriers are used to construct an energy profile di agram to represent all the interconversions, which leads us to infer a minimal energy requirement of 19.22 kcallmol for complete scrambling of the carbons during Cope rearrangement of these systems.

The Cope rearrangement l.2, long considered as a typi­cal pericyclic reaction , is believed to be a one-step concerted process occun·ing via a cyclic transition state and abiding by the conservation of orbital sy m­metrl. The thermal prerequi site and stereochemistry of thi s [3,3] sigmatropic rearrangement have been satisfactorily explained by frontier orbital analysis and the correlation diagram method4

.5. This theoretical

study examines the Cope rearrangement in some bull­valene systems, including degenerate cases with reac­tant and product identical.

Cope rearrangements in 1,5-hexadienes are r3 ,3 ] sigmatropic carbon-carbon bond migrations over a diene framework, subjected to intense experimental studl·14. Originally described as a "no mechanism" reaction 15, the Cope rearrangement was then labelled as a pericyclic concerted reaction via a cyclic aro­matic transition state. Later semi-empirical theoretical work raised the poss ibility of it being an asynchro­nous process involving biradicaloid intermediates l 6

.

More recently , accurate theoretical calculations l7 in­dicate a concerted process as a more correct descrip­tion. Probably the actual situation depends upon the particular system in volved, including substituent and conformational effects, as well as the theoretical method used.

The degenerate Cope rean·angement is a special case where the reactant and product are identi cal, e.g. for unsubstituted or sy mmetrically substituted acyc lic 1,5-hexadiene systems, besides complex systems like semibullvalene and bullvalene involving va lence tautomerism. Bullvalene, first synthesized by Schr6der l8

, involves a Cope rearrangement (Figure 1) attracting much attention due to the phenomenon of complete scrambling of its 10 carbon atoms by multi­ple Cope rearrangements 19 which distribute the posi­tion of the cyclopropane ring among all the carbons. At room temperature four 13C NMR signals are ob­

tained as expected, but at about 100DC they coalesce into only one signaI2o.22, signifying indi stingui shabil­ity of all 10 carbons when the rapid rate of intercon­version between the 10 !/3 di fferent tautomers be­comes comparable to the NMR time scale. The mole­cule adopts a " fluxional " structure, giving only one sharp si ng let in its proton NMR spectrum as well 23 .

M h . I k 24.39 . I . uc expenmenta wor , mall1 y uSing spec-troscop/O.22.30.39, has followed its discovery . Theo-

retical stud/9A I regarding the bullvalene Cope rear­rangement has been relatively less intense. Apart from the interesting feature of "fluxionality", we also note that, unlike the acyclic 1 ,5-hexadiene cases, bullva­lene possesses considerably less conformational flexi­bility, thus restricting the scope for different reaction

3090 INDIAN J. CHEM ., SEC B, DECEMBER 2003

Figure 1 - Degenerate Cope rearrangement of bullvalene showing the reactant (an ' product) and transit ion state .

~ )( ~

/ A B

t F t ----

F

~ x

c D

Figure 2 - Positional isomers of monoOuorobullvalene show ing the direct inte rconversion. Cross marks shows that there is no direct interconversion between the isomers.

pathways during Cope rearrangement. Study of the acyclic] ,5-hexadiene cases42

.43 has identified a num­ber of differing reaction paths (synchronous or asyn­chronous) and transition states (chair, boat, skew, bi­radicaloid) depending upon the initial structure and conformations assumed for the reactant and product geometries, besides the level of theoretical methodol­ogy employed.

Unsubstituted bullvalene itself presents a degenerate case (Figure 1). Non-degenerate cases

include the effects of monosubstitution so that any Cope rearrangement in a monosubstituted bull valene will be non-degenerate. Monofluoro substituted bull va lenes exist as four positional isomers A, B, C and D (Figure 2), and we focu s here on the possible inter-conversions via Cope rearrangements between these four isomers. Six unique inter-conversions may be envisioned, viz., A~B, A~C, A~D, B~C,

B~D and C~D. Of these, only the A~D, C~B and C~A cases can proceed Via single Cope

SINGH et al.: COPE REARRANG EMENT IN BULLVALENE SYSTEMS 309 1

F F

A D

F F

F

c B

F F

c A

Figure 3 - Non-degenerate Cope rearrangement between the positional isomers.

rearrangements (Figure 3), the others occurring only through multiple Cope rearrangements. The net result is that, at a particular temperature like the unsubsti ­tuted case, all fo ur isomers should inter-convert at a rate comparable with the NMR time scale so that complete scrambling of all the carbons occurs. Thus, no spectroscopic differentiation between the four isomers is evident, a single C i3

, HI or FI9 peak being observed20.23.44.

Our study of degenerate Cope rearrangements in­cludes difluorosubstituted bullvalenes, which exist as twelve positional isomers (six corresponding to three antipodal pairs) hav ing 27 unique isomerisat ion proc­esses44

• We choose only cases where di substituti on occurs such that the rearrangement can proceed in degenerate fashion. This leaves only five difluoro­substituted bullvalenes which undergo degenerate Cope rearrangements, labell ed as E, F, G, H and I in

Figure 4 which also portrays the transition state and product for each case.

We seek to compare Cope rearrangement feasib il ­ity for these different bullvalene systems through study of the reaction pathways and transition states, laying focus on transition state stability and geometry . The transition states geometries for the non­degenerate cases (i nvolving monosubstituted f1uoro­bullvalenes) are also gauged to see if they are "early", "mjdway" or " late" along the reaction coordinate. This is related to the endothermicity or exothermicity of the net reaction in each case as per the Hammond postulate45

. Finally, we attempt to correct the AM I values for the activation energies, and use these cor­rected va lues to plot an energy level profile diagram for cases of multiple Cope rearrangements in mon­ofluorosubstituted bullvalenes.

3092

F F

F

F

INDIAN J. CHEM .• SEC B. DECEMBER 2003

F

I , 1',

I

F

F F

F

F

F

F

E

F F

F

G

F

H

""'-F

F J

Figure 4 - Degenerate Cope rearrangement of difluorosubstituted bull valcne showing the Reactant (and product) and transition state.

SINGH el al.: COPE REARRANGEMENT IN BULLVALENE SYSTEMS 3093

Theoretical methodology Both the AM 146 and PM347 semi-empirical SCF­

MO methods of the MOPAC 6.0 package48 were ini­tiaIly used to calculate energy-minimjsed geometries and wave-functions for the reactants, products and transition states in the unsubstituted bullvalene case. Since the PM3 heats of formation for all species were found invariably higher than the AMI values, we opted for the AM 1 method to carry out all further cal­culations. AIl equilibrium geometries were optimised by the Davidon-Fletcher-Powell algorithm49

. We used the PRECISE keyword of the MOPAC package to sharply reduce the threshold for gradient norm during energy minimisation, resulting in improved accuracy and reproducibility.

Location of transition states Different methods were applied to locate transition

states for degenerate and non-degenerate cases. Non­degenerate cases were located by using the standard SADDLE keyword of the MOPAC package, which interpolates between reactant and product geometries through a reverse search strategy, resulting in the transition state. Degenerate cases were treated in two ways as described below.

We firstly assumed symmetry about the reaction coordinate for the transition state in any degenerate Cope rearrangement. We had earlier applied this con­cept of symmetry to rapidly locate transition states in some degenerate cases of the 1,5-sigmatropic hydro­gen shift50 in 1,3-pentadienes. Following this concept, a putative transition state with sy mmetry different from reactant or product may be located by adopting a trial geometry of the appropriate symmetry, which is then simply energy-minimised subject to this symme­try constraint. This procedure results in a transition state if the starting point for optimisation occurs close enough to the transition state itself, or if that is the only possible option on the energy hypersurface. For transition states having the same symmetry as the re­actant, we ensured that the starting geometry for en­ergy-minimisation was closer to the transition state than the reactant. This was performed by simply adopting the energy-minimised transition state for the unsubstituted bullvalene case as a basis to build up starting geometries for the degenerate substituted cases. The advantage of this symmetry constraint method lies in substantial reducti on of computational effort expended in evaluating integrals, gradients and second derivatives .

The other method to locate transition states for de­generate cases was to simply apply the SADDLE keyword described earlier. This procedure was com­putationally more rigorous than the symmetry­constrained search described above, but is more gen­eral in its application, besides being less ambiguous.

Whatever be the method for locating transition states, confirmation of a given stationary point as a transition state was performed by applying the FORCE keyword, which set-up and diagonali sed the Hessian (second derivative) matrix . The emergence of only one negative eigenvalue confirmed the given structure as a transition state5

'.

Results and Discussion We present our results for both degenerate and

non-degenerate cases of Cope rearrangements in bull­valene systems as given below:

Degenerate Cope rearrangement in unsubstituted bullvalene

Table I presents AM 1 and PM3 values of the heats of formation flHJR) and some geometrical parameters for unsubstituted bullvalene, the PM3 energy being higher than the AMI value by 1.44 kcallmol. Only four different carbon-carbon bond types exi st in bull­valene, represented as a, b, c and d (Figure Sa). All other bonds are equivalent to one of these due to the C3v symmetry of bullvalene. To these four C-C bonds, we note a fifth carbon-carbon distance e (Figure Sa) which forms a new type a C-C bond after the rear­rangement (Figure 1). We also note the di tance f between the two vertices, besides the angles f}J and f}2

of Figure Sa, which interchange during the rear­rangement.

The AM 1 values for lengths of bond types a, b, c and d are respectively 1.5185, 1.4598, 1.3386 and 1.4953 A (PM3 values respectively 1.5160, 1.4685, 1.3354 and 1.4976 A). These values are compared in Table I with those obtained from higher level theo­retical calculations24 and with the results of experi­ment using X-ray and neutron diffraction methods52

,

with which there is substantial qualitative agreement. A normal C-C si ngle bond is indicated for bond type a within the cyclopropane moiety , despite the strain effect, and a usual C=C double bond for bond type c, relatively strain-free. Lengths for the single-bond types band d are shorter than the single bond type a

since the termini of bond types band d are connected to doubly-bonded Sp2 carbons. The carbon-carbon distance e (AM 1 value 2.4580 A) indicates ab ence of

3094 INDIAN J. CHEM., SEC B, DECEMBER 2003

(a) (b)

Figure 5 - Geometry markers for (a) reactant , (b) transition state and (c) carbon numbering.

Table I - AM I and PM3 heats of formation and geometrical parameters of un substituted bullvalene* (reactant) along with results of higher level theory and ex periment#

Method l1.Hj (R) a b c d e f 8, flz

AMI 75.40 1.5185 1.4598 1.3386 1.4953 2.4580 3.0340 60.00 122.04

PM3 76.84 1.5160 1.4685 1.3354 1.4976 2.4630 3.0330 60.00 122.22

HF/6-3 1 +G** 1.5178 1.5192 1.3255 1.4831

HF/6-311++G- 1.5366 1.5244 1.3369 1.4845

MP2/6-31 +G** 1.5326 1.5102 1.3529 1.4677

X-ray diffr. 1.5352 1.5157 1.3450 1.4727

Neutron diffr 1.5330 1.5160 1.3420 1.4730

*Heats of formation in kcallmol; bond lengths in angstrom; angles in degrees #For theoretical results, see ref 24 ; for experimental values, see ref 52.

Table II-AM I and PM3 data* for degenerate Cope rearrangement in unsubstituted bullva­lene and geometrical parameters for the transition stat.!

Method l1.HITS) Vi Ell g II f 8

AMI (Sad) 104.46 -0.2260 29.06 2.1908 1.3800 1.4882 2.9900 94.76

AMI (Sym) 104.45 -0.2218 29 .05 2. 1906 1.3804 1.4875 2.9900 94.75

PM3 (Sad) 108.08 - 1.1 431 31.24 2.1027 1.3823 1.4899 2.9930 89.76

PM3 (Sym) 108.08 - 1.11 32 31.24 2.1025 1.3820 1.4899 2.9920 89.76

*Heats of formation and activation energy in kcallmol; Hessian eigenva lue in 105dynes/cm; bond lengths in angstrom; angle in degrees

covalent bonding, while the distance f between the vertices has an AM 1 value of 3.0340 A. The cyclo­propane angle (J, is necessarily exactly 60° by symme­try , while the angle (J2 to which it converts after the rearrangement has an AM 1 value of 122.04°.

Table II gives AM 1 and PM3 values for the heat of formation !1Hr(TS) of the transition state, the acti­vation energy Ea, and the single negative Hessian ei­genvalue Vi, where the SADDLE and symmetry­constrained methods are labelled "Sad" and "Sym"

respectively (as for all tables in this paper). These two methods both yield practically the same values for heat of formation, activation energy and single nega­tive eigenvalue (for both AM I and PM3 Hamilto­nians). The symmetry-constrained approach is much less compute-intensive than the SADDLE strategy , using less than 10% of the computational time, clearly pointing to the effectiveness of the symmetry­constrained method. The values for the activation en­ergy (29.05 kcallmol for AM 1 and 31.24 kcallmol for

SINGH et al.: COPE REARRANGEMENT IN BULLY ALENE SYSTEMS 3095

PM3) may be compared with the experimental value53

of 12.8 kcallmol, which indicates that the semi­empirical AMI and PM3 Hamiltonians yield values appreciably larger in comparison. A systematic error of about 15-16 kcallmol had been suggested by De­war and Jie40 in calculating AM 1 activation barriers for these kinds of tricyclic systems. Since absolute values are not what are sought here, but rather trends among a series, this semi-empirical methodology may yet yield meaningful results.

Table II further presents some PM3 and AM I determinants of transition state geometry (Figure 5b) in the degenerate Cope rearrangement of unsubstituted bullvalene, where both the SADDLE and the symmetry-constrained approaches converge to the same results. The transition state symmetry (C2v)

differs from that of the reactant (C3v), but the labelling system of Figure 5b also applies to more general cases (see later). Important characteristics of transition state geometry are the bond lengths and bond angles involved in the opening of the cyclopropane ring and shift of the 1t-electron system during the reaction. Apart from one ethylenic double bond and the C-C single bonds not directly involved in the rearrangement, there remain three types of carbon-carbon bonds in the transItIon state, represented by the bond types g, hand i (Figure 5b). The bond type g in the transition state (equivalent to g' opposite) corresponds to the bond type a in the reactant (Figure Sa), while the bond type h, here corresponds to the bond type b in the reactant. Due to symmetry, the bond h here is equivalent to the adjacent bond h' (Figure 5b) which corresponds to the bond type c in the reactant (Figure Sa). The bond type i in the transition state corresponds to the bond type d in the reactant. The distance f between the vertices is again taken note of. The angles 8 and e' noted here in the transition state are the same on both sides (Figure 5b) for degenerate cases, corresponding to the angles 8, and 82 of the reactant (Figure Sa). We quote values of geometry markers or heats of formation here as derived from the symmetry­constrained methodology rather than from the standard SADDLE keyword whenever differences arise (all very small in any case).

The AMI value of 2.1906 A (PM3 value 2.1025 A) for the bond length g in the transition state is much larger than the AMI value of 1.5185 A (PM3 value 1.5160 A) for this bond (type a) in the cyclopropane ring of the reactant, indicating cleavage of this bond

during the reaction. The AMI bond length value for the bond types hand h' is 1.3804 A (PM3 value 1.3820 A), intermediate between the single bond length for bond type b (AM 1 value 1.4598 A) and the double bond length for bond type c (AMI value 1.3386 A), indicating that single-double bond inter­conversion occurs symmetrically here in the transition state. The bond length i (AMI value 1.4875 A; PM3 value 1.4899 A) in the transition state differs little from that of the corresponding bond d in the reactant (AMI value 1.4953 A). This is also noted for the di s­tance f between the two vertices, having an AM 1 value of 2.9900 A (PM3 value 2.9920 A).

The bond angles angles 8 and e' on both sides of the transition state have an AMI value of 94.76°(PM3 value 89.76°). The increase from the 8, value of 60° in the reactant is due to lengthening of the cyclopropane C-C bond a of the reactant as it cleaves during the reaction. The decrease from the AMI 82 value of 122.04° in the reactant is due to shortening of the non­bonded carbon-carbon distance e in the reactant as a new bond forms.

We see that the bullvalene Cope rearrangement centers around a non-planar cyclic six-membered ring moiety which is boat-shaped, the two allylic moieties connected by two bonds of the bond type g which un­dergo breaking and forming (of the bond types a and e in the reactant). It also resembles an open book of C2v symmetry with a dihedral angle of about 130.210 between the two halves. It thus differs from the chair­shaped or skew-shaped transition states reported for the Cope rearrangement in acyclic 1,5-hexadiene sys­tems, which can incorporate a boat-shaped species as we1l54

. The conformational flexibility of the 1,5-hexadiene system allows for a variety of shapes for the transition state, each depending upon the con­former chosen for reactant and product. The bull va­lene system is conformationally less flexible , with only this boat-shaped transition state envisioned so far for the Cope rearrangement here.

Non-degenerate Cope rearrangments in monosubsti­tuted jluorobullvalenes

From the six inter-conversions among the four monosubstituted fluor bullvalene isomers (Figure 2), only the A~D, the C~B and the C~A cases pro­ceed through single Cope rearrangements (Figure 3) . The first two of these involve symmetrical transition states (Cs symmetry) which are calculated by both SADDLE and symmetry-constrained strategies. The other three inter-conversions A~B, B~D and C~D

3096 INDIAN 1. CHEM., SEC B, DECEMBER 2003

occur only through multiple Cope rearrangements (see later).

Table III gives the AM I values for the heats of formation I'1Hj(R) along with the symmetry point group of the four positional isomers A, B, C and D (Figure 2). For the purpose of this paper, bracketed point groups in the tables, e.g. [e2v], indicate arrival at this symmetry after geometry optimjsation even with­out imposing any symmetry constraints. Point groups without brackets indicate that symmetry constraints were imposed. The energies I'1Hi of each isomer rela­tive to the most stable isomer are also given in Table III. The lowest energy isomer is C, and the stability order is C > B > A > D. Fluori ne substitution in the cyclopropane ring is well in accord with those already predicted55

, to have destabilising effect. Fluorine sub­stitution on any s/ carbon terminus of a C-C double bond tends to stabilise the system while substitution far away from the cyclopropane ring destabilises the system.

Table IV presents the AM I heats of formation I'1HiCTS) of the transition states for the non-degenerate Cope rearrangements involving these monosubstituted fluorobullvalenes, along with their symmetry point groups and the reaction enthalpy I'1Hr for the forward reaction. Values of activation energies for both for­ward and reverse reactions are presented as Ea(F) and Ea(R) respectively. Table IV also furnishes the Hes­sian negative eigenvalue Vi confirming the transition state. The computational strategies are labelled SAD­DLE (Sad) or symmetry-constrained (Sym).

Table III-AM I data* on heats of formation of four positional isomers for monosubstituted f1uorobullvalene

Isomers A 8 C D

6Hr(R) 33.57 29.98 29.25 35.21

Point group Cs C, C, C.,

6H; 4.32 0.73 0.0 5.96

*Enthalpies in kcallmol

The symmetry-constrained approach applied to the A---+D and the C---+B conversions results in substantial saving of computational time with no loss of essential accuracy for values of the various parameters li sted in Table IV. The heats of formation of the transition states range from 57.56 kcallmol to 64.77 kcal/mol , the most stable one being that for the C---+A conver­sion, and the least stable being that for the A---+D con­version, reflecting n the differing effects of position of the fluorine substituent in the transition state.

Calculated AM I va lues for the activation eneigy E" range from 23.99 to 31.20 kcal/mol, comparable to the AM I value for the unsubstitu ted ca e (29.05 kcal/mol) , but certainly on the large s ide when com­pared with experim ntal expectation. The forward and reverse reaction feasibilities as given by the activation

barriers EaCF) and Ea(R) predict that the CHA inter­conversion would be kinetically most facile, and the AHD one least facile, regardless of reaction direc­tion. Note that the order of absolute magnitude for the negative Hessian eigenvalues runs parallel to that for the activation energies, which we have noted in our other theoretical work56

.

The positive values of react ion enthalpy I'1Hr pre­dict endothermicity for the three conversions A---+D, C---+B and C---+A CI'1 Hr = 1.64, 0.73 and 4.32 kcallmol respectively). This endothermicity depends , of course, on the direction taken for the reaction, where our choice is as above. The reaction enthalpies give the ordering for degree of endothermicity as : C---+A > A---+D> C---+B. Since the reaction enthalpies for these reactions are small compared with the activation en­ergies, the Hammond postulate predict the transition states to be quite different from ei ther reactant or product, but located approximately midway along the reaction pathway and yet a little closer to the product, viz. having some small degree of " late" character. We confirm these inferences by examining aspects of transition state geometry relating to bond breaking and making, besides angle inter-conversions occur-

Table IV-AM I data* on inter-conversion rcactio ns (non-degcnerate Co pc rcarrangements) for mono-substitutcd f1uorobullvalencs

Reaction 6HfTS) 6Hr £ ,,( F) £,,( R) II; Pt Gp (TS)

A--+D (Sad) 64.77 1.64 31.20 29.56 -0.2656 [Cs]

A--+D (Sym) 64.77 31.20 29.56 -0.2651 C.,

C--+8 (Sad) 58.43 0.73 29.18 28.45 -0.1955 [C,l

C-+8 (Sym) 58.43 29.18 28.45 -0. 1939 I c.,j C--+A (Sad) 57.56 4.32 28.31 23.99 -0.0880 C,

*Heat of formation and act ivati on energies in kcallmol ; 11,' in 105 dynes/cm

SINGH el al.: COPE REARRANGEMENT IN BULLVALEN E SYSTEMS 3097

ring during the reaction, which we define by refen'ing to the general case of transition state geometry given in Figure Sb. These aspects include lengths of the bonds g and g', lengths of the bonds hand h', as well

the values of the angles Band 8', as given for the gen­eral case of Figure Sb. Each of these pairs of geome­try markers have exactly equal values for the "mid­way" transition state in the degenerate Cope rear­rangement for unsubstituted bullvalene. They are not equal for non-degenerate cases, though, and the ratios

gig', hlh' and BI8' may give some idea of the position of the transition state along the reaction pathway. Late transition states are expected to have values of gig' and BI8' larger than unity and values of hlh' smaller than unity, and conversely for early transition states. For the C-tA case, due to the transition state asym­

metry , two sets of hand h' bonds exist with two val­

ues for the hlh' ratio.

Table V gives AMI values of these geometry markers along with the gig', h/h' and BI8' ratios for the three processes. The values of these ratios all closely approach unity, indicating the approximately "midway" character of all these transition states. The

precise values of the gig' and· hlh' ratios indicate that these transition states all possess some degree of

"late" character. The values of the gig' ratio for the processes A-tD, C-tB and C-tA (1.00320, 1.00502 and 1.00270 respectively) are all slightly larger than unity and indicate that the transition states are "late"

to a small extent. The hlh' ratios are all slightly smaller than unity , being 0.99877, 0.99812 and 0.98029 (0.99573) for the processes A-tD, C-tB and C-tA respectively, likewise inferring the slightly " late" character of all these transition states. This in­ference, however, is not drawn uniformly from the values of the BI(f ratio, which are larger than unity for the processes C-tB and C-tA as expected, but smaller than unity for A-tD. Apparently, the bond length ratios are more effective in corroborating the

Hammond postulate since the bonds involved are di­rectly present in the rearranging cyclic moiety of the

transition states, while the angles Band 8' are not. In most aspects, transition state geometries for

these 3 non-degenerate Cope rearrangements closely resemble that for the degenerate unsubstituted case, invoking the same boat-shaped type of moiety at the center. This may be seen by comparing the values of the g, g' , h, h', Band 8' geometry markers for these non-degenerate transition states with the correspond­ing values of the g, hand B markers for the unsubsti­tuted case (Table II).

The effect of fluorine substitution upon transiti on state stability and activation energy Ea depends upon the site of substitution in the transition state, and may be gauged by comparison with the unsubstituted case having an Ea value of 29.05 kcallmol. Figure Sc pre­sents a numbering system for identifying substitution sites in the transition state, given as C l

, C2, C3and C.J.

Substitution at the C l position is on a double bond not involved in the realTanging moiety of the transition state, and has only a slight destabilising effect. Fluo­rine substitution at the C2 position (the A-tD case) increases En to 31.20 kcallmol (Table IV). Thi s may be explained by invoking the -1 inductive effect of the fluorine at a cyclopropane ring which has been pre­dicted to have a destabilising effect on the ri ng55 ow­ing to a increased 2s character in the orbital s present in the carbon atom as it bonds. Fluorine substitution at the C' position (the C-tA) case lowers Ea to 28.31 kcal/mol. This stabilising effect is due to the lone­pairs on fluorine being in conjugation with the cyclo­propane ring bond being formed thereby ass isting its formation during the pericyclic process.

Degenerate Cope rearrangements in dijluorosubsti­tuted bullvalenes

Table VI presents AM I values for t..HI R) and t..HjTS) , the heat of formation of the reactants and transition states, E", the activation energy, along with

Table V- AM I dala* for geometri cal paramelers o f lransition stat.es in non-degenerate Cope rea rrangemenls in vol ving monofluorosubstiwted bullval ene systems

Reaction g , II h' f) rJ' gig' hilI' ()IrJ' g

A--+ D (Sad) 2 . 1950 2.1870 1.3784 1.3807 93.62 94.57 1.00366 0.99870 0 .98995 A--+ D (Sy m) 2.1940 2 .1 870 1.3785 1.3802 93.62 94.59 1.00320 0.99877 0.98975

C --+ B (Sad) 2.2030 2 .1920 1.3788 1.38 15 95.47 95 .07 1.00502 0 .99804 1.0042 1 C --+ B (Sym) 2.2030 2 .1920 1.3788 1.38 15 95.46 95.08 1.00502 0 .998 12 1.00400 C --+ A (Sad) 2 .2260 2 .2200 1.3724 1.4000 96.76 95 .90 1.00270 0 .98029 1.0091 8

1.3768 1.3827 0.99573

*Bond lengths in angstro m; angles in degrees

3098 INDIAN 1. CHEM., SEC B, DECEMB ER 2003

Table VI-AM 1 data* on reactants and transition states for degenerate Cope rearrangements involving difluorosubst ituted bullvalene systems

Reaction PtGp PtGp I'1.H/ R) I'1.HPS) Ea II; T (R) (TS)

E-.E (Sad) [Csl rC2,,] -6. 10 25.63 31.73 -0.3079 2895

E-.E (Sym) [Cs ] Cb · 25.63 31.73 -0.3082 525

F-.F (Sad) rC .• ] [C2v] - 14.25 14.25 28.50 -0.1683. 2848

F-.F (Sy m) [Cs ] C2" 14.25 28.50 -0.1 69 1 9 10

G-.G (Sad) rC,] ICb ,] -14.99 17.55 32.54 -0.3806 2780

G-.G (Sym) [Cs] C21. 17.55 32.54 -0.3793 805

H-.H (Sad) C, rC.1 - 11.88 11.31 23. 19 -0.0297 2805

I -. I (Sad) C, rC2] - 12.24 10.85 23.09 -0.0300 2685

*Enthalpies in kcallmol: v;in 10 5 dynes/cm; T in seconds'

Table VII-AM I data* on geometrical parameters for transi tion states of degenerate Cope rearrangements invol ving difluorosubstituted bullvalene systems

Reaction g h ()

E-.E (Sad) 2. 1900 1.3782 1.5044 93.42

E-.E (Sym) 2. 1900 1.3783 1.5043 93.43

F-.F (Sad) 2.2050 1.3798 1.4853 95.77

F-.F (Sym) 2.2040 1.3797 1.4853 95.79

G-.G (Sad) 2.1 880 1.395 1 1.4854 94.76

G-.G (Sy m) 2.1 880 1.395 1 1.4853 94.76

H-.H (Sad) 2.2463 1.3927 1.5070 97.67

1.3787 1.484 1

1 -. 1 (Sad) 2.2580 1.3940 1.5052 96.96

1.373 1 1.4858

*Bond lengths in angstrom and angles in degrees

the Hessian negative eigenvalue and computational time T for the 5 degenerate Cope rearrangements in­volving the fi ve difluorosubstituted bullvalenes E, F , G , H and I portrayed in Figure 4, listing also the symmetry point groups for the reactants and transition states. Heats of formation for the reactants range be­tween -14.99 and -6.10 kcallmole, g iving the stability order G > F > I > H > E for the 5 isomers. The Cope rearrangements involving the systems E, F and G have transition states with C2v sy mmetry , so we used the symmetry-constrained method to locate them be­sides the routine SADDLE keyword. Trans ition states fo r the H and I cases have C and C2 symmetry re­spectively, and may also be calculated using the sy mmetry-contrained method.

Both SADDLE and symmetry-constrained methods lead to essentially the same results regarding geometry and energy of the transition states of C2v

symmetry (for systems E , F and G ), demonstrating immense reduction of computational time when the latter method is employed. The transi tion states have heats of formation flHjTS) ranging from 10.85 to 25.63 kcallmole, g iving the stability order I> H > F > G > E. The calculated AMI values for the acti vation energies range from 23 .09 to 32.54 kca llmol, yielding the order of magnitude G > E > F > H > I, implying the reverse order for kinetic facility of the reaction. Here again, the AM 1 values for the activation enthalpies are apprec iably larger than what would be expected from experiment or higher level theoretical calculations.

Table VII presents geometrical parameters pertaining to the transition states for degenerate Cope rearrangements in these 5 difluorosubstituted bul­lvalenes. These geometry markers correspond directly to those of the general case portrayed in Figure 5b,

SINGH el al.: COPE REARRANGEMENT IN BULLVALENE SYSTEMS 3099

Table VIII-Corrected ac tivation enthalpy values* for Cope rearrangements in bullvalene systems

System ullsubstd A-)O C-)B

Ea (AM I) 29.05 31.20 29.18

Ea (corrected) 12.75 14.90 12.88

* All enthalpies in kcallmo1

the characteristic geometry changes being represented by the bond types g and h along with the bond angle B. The values of these markers indicate that the transition states for these 5 degenerate Cope rearrangements again closely resemble that of the unsubstituted bullvalene case (Table II) , giving the same boat shape. Note that for the H and I cases, the lowering of symmetry in the transition states calls for two sets of h values, where the first referring to the h bond towards the viewer (Figure 4) and the second referring to that directly behind it.

The chief difference between these degenerate re­actions and the unsubstituted case is the symmetrical difluorosubstitution. The effect of fluorine substitu­tion upon transition state geometry is gauged from the lengths of the bonds directly attached to a fluorine atom. Fluorine substitution at either terminus of an h type bond has the effect of lengthening it. The fluo­rine-bonded It type bonds in the G~G, H~H and I~I cases show this increase, with lengths from 1.3927 to 1.3951 A. Other It type bonds are shorter, ranging between 1.3731 and 1.3797 A. The - / induc­tive effect of fluorine diminishes the h bond electron density , reducing its partial double bond character further. A similar bond lengthening effect occurs for the i type bonds, which are fluorine substituted in the E ---+E case, as well as on one side each for the H~H and 1---+1 cases.

The effect of fluorine substitution upon the activa­tion energy Eo is again site-dependent , reproducing the trends shown for the monosubstituted cases. Di­fluoro substitution at the two C2 sites (the E ~E case) increases the Ea value as compared with the unsubsti­tuted case, as occurs for the C2-monosubstituted case dealt with earlier, and may be explained by a similar rationale. Disubstitution at the two C3 sites (the H~H and 1---+1 cases) has the effect of lowering activat ion energy, as occurs for the C3 -monosubstituted case dealt with earlier, and may be simi larly rat ionali sed. Disubstitution at the two C4positions (the G---+G case) has a destabili sing effect not dealt with in the mono­substi tuted cases. It may be explained by the electron­releas ing effect of the fluorine atoms (through thei r

C-)A E-)E F-)F G-)G H-)H 1-) I

28.31 31.73 28.50 32.54 23.19 23.09

12.01 15.43 12.20 16.24 6.89 6.79

lone pairs) which should have a destabilising effect as concluded by Hoffmann and Stohrer57 in their work on similar systems.

Corrected values of activation barriers We now attempted prediction of realistic activation

energy values for all these Cope rearrangement reac­tions on the basis of the systematic error of 15-16 kcallmol suggested by Dewar and Jie4°for this type of system. Fixing this error at the value of 16.3 kcallmol proposed by them for bullvalene, Table VIII presents our AM 1 values of activation enthalpy E" along with the corrected values for all the 9 systems treated here. The corrected values (6.79 kcal/mol to 16.24 kcal/mol), are probably more in line with observed ranges of values as obtained for these kinds of sys­tems, but require experimental verification. Relati ve kinetic facility for the Cope rearrangement in the various systems studied then assumes the following order: I ~ I > H~H > C~A > F~F > unsubstituted case > C~B > A~D > E~E > G~ G.

Multiple Cope rearrangements in 1110nofluorobullva­lenes

The corrected values of Ell are used to plot an en­ergy profile diagram (Figure 6) depicting the en­thalpy changes accompanying the multiple Cope rear­rangements for the multi-step inter-conversions among monofluorosubstituted bullvalenes, which in­clude the A~B, B~D and C~D conversions, along with the three unique transiti on states TSl , TS2 and TS3. We plot Figure 6 taking the starting point corre­sponding to the lowest energy isomer C as zero. This diagram portrays energetically all the possible proc­esses (conversions) occurring in the Cope rearrange­ment reaction involvi ng monosubst ituted fluorobull­valenes. In order to ensure complete scrambling of all the carbons among these monofl uoro bullvalenes, the energy requirement would be the enthalpy difference between the lowest energy isomer C and the highest energy transition state TS3. From this diagram, it emerges that a minimum energy requirement of 19.22 kcal/mol would be required to be supplied in order for

3100 INDIAN J. CHEM ., SEC B, DECEMBER 2003

25 ---------------- - ----------------- - ,

c o

20

tl 5 '" ~ "-o

~ 0 :r:

-5

D

Reaction coordinate

Figure 6 - Reaction profile of multiple Cope rearrangement of monoflu orobullvalene systems.

complete scrambling of carbon atoms to occur among these four monofluorosubstituted bullvalenes.

Conclusions These semi-empirical AM 1 calculations predict

that a pericyclic one-step reaction is possible and feas ible for all these cases of bull va lene Cope rear­rangements , each involving a boat-shaped six­membered cyclic rearranging moiety within the tran­siti on state. While transiti on states for the degenerate cases (unsubstituted and symmetrically difluoro sub­stituted) are exactly midway along the reaction coor­dinate, those for the non-degenerate monofluorosub­stituted cases are slightly " late" in character, which observation is supported by their geometries in ac­cor'dance with the Hammond postulate. The use of sy mmetry constraints to locate transition states for the degenerate cases is found to be highly effective in reducing computational cost. The effects of fluo­rine substitution on activation energies and on transi­tion state geometries as predicted here are rational­ised qualitatively by appealing to the inductive ef­fect. A systematic correction is applied to the activa­tion energy barriers, yielding an energy profile dia­gram for all the six inter-convers ions among the monofluorosubstituted bullvalenes which predicts a minimum enthalpy requirement of 19.22 kcal/mol fo r the complete scrambling of carbon atoms in these sys tems.

References I Cope A C & Hardy E M, J Alii Chem Soc, 62, 1940, 44 1. 2 Cope A C, Hoy le K E & Heyl D, J Alii Chem Soc, 63, 1941 ,

1843. 3 Woodward R B & offman R, The conservation of orbital

sYlllmetry (Verlag Chemie, Weinheim), 1970. 4 Fleming I, FronTier orbiTals and organic chelllical reacTions

(John Wiley and Sons, New York), 1976. 5 Woodward R B & Hoffman R, Angew Chem fm Edn, 8, 1968,

78 1. 6 Owens K A & Berson J, J Am Chelll Soc. 11 2, 1990,5973. 7 Slanina Z, CollecT Czech Chem Comlllllll , 48, 1983,3027. 8 Roth W, Lennartz H oW, Doering W v E, Birladeanu L, Guy­

ton C A & Kitagawa T, JAm Chem Soc, I 12, 1990, 1722. 9 Gajewski J J, Hydrocarbon Th ennat isomeriW Tions (Aca­

demic Press, New York), 1981, p. 166- 176. 10 Doering W v E, Toscano V G & Beasley G H, TeTrahedron ,

27, 1971,5299. II Doering W v E & Wang Y, J Am Chem Soc, 12 1, 1999,

10112. 12 Doering W v E & Wang Y, J Alii Chelll Soc, 12 1. 1999,

10967. 13 Doering W v E, Birladeanu L, Sarma K, Blaschke G, Schei­

demantel U, Boese R, Benet-Bucholz J, Klamer F -G, Gehrke J oS, Zinny B U, Sustmarin R & Korth H -G, J Alii Chelll. Soc, J 22, 2000, J 93.

14 Owens K A & Berson J A, JAm Chelll Soc, I J 2, 1990, 5973. 15 Doering W v E & Roth W R, TeTrahedron, 18, 1962, 67. 16 Dewar M J S, Ford G P, Mckee M L, Rzepa H S & L E

Wade, JAm Chem Soc , 99, 1977,5069. 17 (a) 0 Wi est, K A Black & K N Houk , J Am Chem Soc, J 16.

1994, 10336. (b) H Ji ao & P v R Schleyer, Angew Chelll fill Ed Eng , 34 , 1995,334.

18 Schroder G, Angew Chem fill Edn, 2, 1963, 48 1.

SINGH et al.: COPE REARRANGEMENT IN BULLVALENE SYSTEMS 3101

19 Doering W v E, Zh Vsesoyu z Khirn Obshchestva illl D I Men­deleeva, 7,1962,308.

20 Oth J F M, Muller K, Gilles J & Schroder G, Helv Chirn Acta, 57,1974,1415.

21 Nakani shi H & Yamamoto 0, Tetrahedran Lell , 1974, 1803.

22 Gunther H & Ulmen J, Tetrahedron. 30, 1974, 3781. 23 Ault A, J Chem Educ, 78, 2001, 924. 24 Koritsanszky T, Buschmann J & Luger P, J Phys Chel1l , 100,

1996, 10547. 25 Sarma K & Schroeder G, Chern Ber, 117, 1984, 633. 26 Mansson M & Sunner S, J Chel1l Thermodynamics, 13, 1981,

671. 27 Falk B, J Chern Thermodynamics, 12,1980,967. 28 Rebsamen K, Roettele H & Schroeder G, Chern Ber, 126,

1993, 1429. 29 Rebsamen K & Schroeder G, Chem Ber, 126, 1993, 1425 . 30 Moreno PO, Suarez C, Tafazzoli M, True S & LeMaster C

B, J Phys Chel1l , 96, 1992, 10206. 31 Boeffel C, Luz Z, Poupko R & Zimmermann H, J Magn

Reson, 85, 1989, 329. 32 Luz Z, Poupko R & Alexander S, J Chem Phys, 99, 1993,

7544. 33 Titman J J, Luz Z & Spiess H W, JAm Chelll Soc, 114, 1992,

3765. 34 Schlick S, Luz Z, Poupko R & H Zimmermann, J Am Chem

Soc, 114, 1992, 4315. 35 Meier B H & Earl W L, JAm Chem Soc, 107, 1985, 5553. 36 Luz Z, Olivier L, Poupko R. Mueller K, Kri eger C &

Zimmermann H, .f Am Chem Soc, 120, 1998, 5526. 37 Moreno PO, Suarez C, Tafazzoli M, True N S & LeMaster C

B, J Phys Chem, 96,1992, 10206.

38 Schlick S, Luz Z, Poupko R & Zimmermann H, JAm Chem Soc, 114, 1992, 4315 .

39 Luger P & Roth K, J Chem Soc Perkin Trans 2, 1989,649. 40 Dewar M J S & Jie C, Tetrahedral! , 44,1988, 1351. 41 Gimarc B M & Zhao M, J Org Chem, 60, 1995, 1971 . 42 Beapark M, Bernardi F, Olivucci M & Robb M A, J Am

Chem Soc, 112, 1990, 1732. 43 Komornicki A & Mciver Jr J W, JAm Chem Soc, 98, 1976,

4553 . 44 Schroder G & Oth J F M, Angew Chem internat Ed, 6, 1967,

414. 45 Hammond G S, JAm Chem Soc, 77, 1955,334. 46 Dewar M J S, Zoebi sch E G, Healy E F & Stewart J J, J Am

Chem Soc, 107, 1985, 3902. 47 Stewart J J P, J Comput Chem, 10, 1989,22 1. 48 Stewart J, Frank J Seiler Research Laboratory, US Air Force

Academy, USA. 49 (a). Davidon W C, Comput J, 10, 1968, 406.

(b). Fletcher R & Powell M J D, Computer J, 6, 1963, 163 . 50 Singh P M & Lyngdoh R H D, Indian J Chem Sec 40A, 2001 ,

682. 51 McI ver J W & Komornicki A, J Am Chem Soc, 94, 1972,

2625. 52 Luger P, Buschmann J, Mc Mullan R K, Ruble J R, Mat ias P

& Jeffrey G A, JAm Chem Soc, 108, 1986,7825 . 53 Allerhand A & Gutowsky H S, J Am Chem Soc, 87 , 1965,

4092. 54 Dewar M J S & Jie C, J Chem Soc Chem Commun, 1987, 1451. 55 Wiberg K B & Marquez M, JAm Chem Soc, 120, 1998, 2932. 56 Dkhar P G S & Lyngdoh R H D, Indian J Chem Sec A (com-

municated). 57 Hoffman R & Stohrer W - D, JAm Chem Soc, 93,1971,6941.