Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Reactive Intermediate Chemistry

.pdf
Скачиваний:
200
Добавлен:
08.01.2014
Размер:
12.65 Mб
Скачать

284 SINGLET CARBENES

 

 

C H3O OCH 3

 

 

N

benzene

O

 

 

 

(CH 3O) 2C + (CH3 )2C=O

 

 

 

 

 

110 °C

N

–N2

H3C CH3

 

 

Scheme 7.3

attack on the carbonyl group of cyclobutanone. A zwitterionic intermediate, 10, leads to the ring-expanded, masked diketone (11, Scheme 7.4). When either of two substrate carbons can be involved in the expansion, selectivity favors migration of the more substituted carbon.

 

 

CH3O +

OCH3

 

 

O

C

_

CH3O O CH3

 

 

(CH3O )2C

benzene

O

O

 

 

+

°C

 

 

 

110

 

 

 

 

1 0

 

1 1

Scheme 7.4

Similarly, Rigby and coworkers51 used the oxadiazoline method to generate dithiacarbenes, which initiate nucleophilic attacks on isocyanates (Scheme 7.5). In this case, nucleophilic attack on the isocyanate ‘‘carbonyl’’ leads to ring closure via intermediate 12. The proximate product then undergoes an apparent N H insertion reaction with a second carbene to yield the final product. The philicity of the N H insertion is unclear; it could be initiated by ‘‘electrophilic’’ attack of the ambiphilic (C3H7S)2C on the substrates’s nitrogen lone pair, or it might originate in N H proton abstraction by a ‘‘nucleophilic’’ carbene acting as a base. Related reactions of (CH3O)2C with isocyanates are known.52

C3H7S

SC3H7

 

 

 

 

 

 

 

 

N

 

 

 

 

 

+

 

refl. be nze ne

(C3H7 S) 2C

 

 

C(SC3H7 )2

O

 

+

 

 

-N2, -(CH 3)2C=O

 

N=C

 

N

 

 

N=C=O

 

CH3

CH 3

 

 

 

 

 

 

O _

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1 2

 

C3H7S

SC3H7

 

C3H7S

SC3H7

 

 

 

 

O

(C3H7S) 2C

O

(C3H7 = propyl)

 

 

 

 

 

 

N

 

 

 

 

N

 

 

 

H

 

 

 

 

 

 

 

CH(SC3H7 )2

Scheme 7.5

MORE DETAILED TREATMENTS OF STRUCTURE AND PROTOTYPAL REACTIONS

285

Finally, addition reactions of the isolable phosphasilylcarbenes (13) to such electron-poor substrates as methyl acrylate, C4F9CH CH2, and styrene afford cyclopropanes. The additions of 13a to (E)- or (Z)-b-deuteriostyrene are stereospecific, and the competitive additions of 13b to ring-substituted styrenes exhibit nucleophilic selectivity, consistent with singlet, nucleophilic carbene addition (Fig. 7.8).53

..

(R2N)2P–C–Si(CH3)3

13a, R = i-C3H7; b, R = cyclohexyl

Figure 7.8

2.1.2. Rates and Activation Parameters. The first condensed-phase absolute rate measurement for a carbene–alkene addition was reported by Closs and Rabinow in 1976: flash lamp photolysis of diphenyldiazomethane generated

(triplet) diphenylcarbene,

which added to butadiene (in benzene) with k

¼

 

54

 

6:5 105 M 1s 1 at 25 C.

 

However, most singlet carbene additions to olefins

are faster, and their addition reactions are too rapid for the time resolution afforded by a flash lamp.

The development of laser flash photolysis (LFP), with flash times of 10–20 ns, made possible the measurement of many absolute rate constants for singlet carbene additions. The initial reports focused on chlorophenylcarbene (C6H5CCl)55 and singlet and triplet fluorenylidene.56 Consider C6H5CCl: LFP at 351 nm of chlorophenyldiazirine (14) in isooctane containing a given concentration of an alkene such as (CH3)2C C(CH3)2, affords a transient ultraviolet (UV) signal for C6H5CCl at 310 nm that decays as the carbene adds to the alkene. In the absence of an alkene, the carbene can decay by dimerization, by reaction with the diazirine precursor, or by reaction with adventitious water. When sufficient alkene trap is present, however, these pathways are suppressed. From the time dependence of the carbene signal’s decay, we obtain a pseudo-first order rate constant (kobs) for carbene–alkene addition. Variation of the alkene concentration and repetition of the LFP experiment yields additional values of kobs, and correlation of kobs with [alkene] gives a straight line, the slope of which is the second-order rate

constant for the addition of C6H5CCl

to (CH3)2C C(CH3)2, in this case, k

¼

2:8

 

108 M 1s 1 Scheme 7.6).57

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

CH3

CH 3

C6H5

N

L FP

..

( CH3)2 C=C (CH3) 2

H3C

CH 3

 

 

 

 

 

 

 

C6H5CCl

 

 

Cl

N

-N2

 

 

 

 

 

Cl

 

 

 

14

 

 

 

C6H5

 

 

 

 

 

 

 

 

Scheme 7.6

286 SINGLET CARBENES

This ‘‘direct’’ method works quite well as long as the carbene has an accessible and

sufficiently strong UV signature. Many absolute rate constants for carbene additions were measured in this way.24,26,57 Data for several of these species appear in Table 7.5.57–62

TABLE 7.5. Absolute Rate Constants for Carbene–Alkene Additionsa

Alkene

C6H5CFb

C6H5CClb

CH3OCClc C6H5COCH3d CH3OCOCH3e

 

 

 

 

 

 

 

 

 

 

 

 

 

ðCH3Þ2C CðCH3Þ2

 

8

 

8

4:2 10

3

 

 

 

 

1:6 107

2:8 108

 

 

 

 

3

ðCH3Þ2C CHCH3

5:3 10

1:3 10

 

 

3

 

4

4:7 103

ðCH3Þ2C CH2

 

6

 

6

1:8 102f

4:0 103f

4:8 103f

trans-CH3CH CHC2H5

2:4 105

5:5 106

3:3 10

 

3:8 10 3

2:2 10

n-C4H9CH CH2

9:3 106

2:2 106

 

 

3

2:8 105

 

5

CH2

CHCOOCH3

1:4

106

5:1

106

9:8

104

6:6

106

7:9

106

CH2

CHCN

2:3

108

7:0

108

1:8

105

1:7

107

1:5

107

CH2

CClCN

1:2

10

2:1

10

5:6

10

 

3:4

10

4:9

10

a Rate constants are in M 1 s 1 and were measured at 23–25 C in alkene–hydrocarbon solvents. b Data from Ref. 57–60.

c Data from Ref. 61.

d Data from Ref. 59, 61, and 62. e Data from Ref. 33 and 61.

f Data are for trans-butene.

The data reveal a very large kinetic span for carbene–alkene addition reactions with k ranging from > 108 M 1 s 1 for additions of C6H5CCl or C6H5CF to (CH3)2C C(CH3)2 to 3.3 102 M 1s 1 for the addition of CH3OCCl to transpentene. This million-fold variation testifies to the great modulating power of carbenic substituents on carbenic reactivity.

Both C6H5CCl and C6H5CF behave as ambiphiles toward the alkene set of Table 7.5. They react rapidly with both electron-rich (CH3)2C C(CH3)2 and electron-poor alkenes CH2 CClCN; their slowest reactions are with the mono-

alkylethylene, 1-hexene. Although ambiphilicity is an intrinsic property of any singlet carbene,25,26 the absolute reactivity of, for example, C6H5CCl is much

greater than that of the classic ambiphile CH3OCCl.31 The former reacts67,000 times more rapidly with (CH3)2C C(CH3)2, 375 times more rapidly with CH2 CClCN, and 17,000 times more rapidly with trans-pentene. The CH3O for C6H5 ‘‘swap’’ that formally converts C6H5CCl into CH3OCCl is responsible for very significant resonance stabilization of the carbene (see 8), and a concomitant decrease in reactivity.

Even the F for Cl change transmuting C6H5CCl into C6H5CF is accompanied by a diminution in reactivity; C6H5CF reacts about one-half as rapidly as C6H5CCl with the alkenes of Table 7.5. This lowering of reactivity can be attributed to greater stabilization of C6H5CX when X F, as opposed to X Cl. Fluorine is a better resonance electron donor to the vacant carbene 2p orbital than chlorine, and this factor (rather than the greater electronegativity of fluorine) appears to be dominant.63

MORE DETAILED TREATMENTS OF STRUCTURE AND PROTOTYPAL REACTIONS

287

What about CCl2, the paradigmatic electrophilic carbene?2,3,28 It can be generated by photoextrusion from its phenanthrene adduct (15, Scheme 7.7).64 Although CCl2 lacks a strong UV signal to follow directly, one can use Jackson and Platz’s method65a in which CCl2 addition to pyridine (affording a UVactive ylide) competes with CCl2 addition to an alkene substrate. The desired second-order rate constant for the CCl2–alkene addition can be extracted from the dependence of the apparent LFP rate constant for ylide formation as a function of alkene concentration (at a constant concentration of pyridine). For a detailed exposition of the kinetic principles underlying this methodology, the reader is referred to Chapter 18 in this volume.

Cl Cl

 

 

 

 

 

 

 

hν

 

 

 

 

 

 

 

+

Cl2C

 

15

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ð

 

 

 

Þ

 

Scheme 7.7

 

 

 

 

 

 

Rate constants

M 1s 1

9

 

 

 

 

 

 

9

 

 

 

 

were thus measured for dichlorocarbene additions to

 

 

 

 

 

 

 

 

 

 

 

 

 

(CH3)2C

C(CH3)2

(3.8

 

 

 

10 ), (CH3)2C CHCH3

(2.2

 

10 ), trans-pentene

(6.3 10

7

 

 

 

 

 

 

7

), and 1-hexene (1.1 10

7

). Dichlorocarbene

), cyclohexene (3.5 10

 

 

is again seen to be an electrophile, as we concluded from the relative reactivity data of Table 7.3, and the differential orbital energies of Table 7.4. Suprisingly, although CCl2 is somewhat more selective than C6H5CCl [e.g., krel for (CH3)2C C(CH3)2/ n-BuCH CH2 is 345 for CCl2 vs. 127 for C6H5CCl], CCl2 is also marginally more reactive than C6H5CCl. The kobs values for CCl2–alkene additions exceed those of C6H5CCl in each case by about a factor of 10.

On the other hand, the CH3O for Cl exchange that converts CCl2 into CH3OCCl is accompanied by about a 105-fold decrease in reactivity (Table 7.5). The O(2p)– C(2p) resonance electron donation of CH3O is much more effective than the analogous Cl(3p)–C(2p) interaction. Another way to approach this reactivity comparison is to note that CCl2 is a very reactive electrophile because of its low-lying, accessible LUMO orbital (0.3 eV)36 whereas the LUMO of CH3OCCl (2.5 eV)36 is raised by resonance donation from CH3O and is much less accessible. Indeed, k ¼ 3.3 102 M 1 s 1 for the addition of CH3OCCl to trans-butene is currently the lowest known bimolecular rate constant for a condensed phase carbene–alkene addition.61

Similar analyses can be made for the addition reactions of C6H5COCH3 and CH3COCH3 (Table 7.5). Both carbenes are ambiphiles/nucleophiles that are more reactive than CH3OCCl. Thus, the Cl for CH3 exchange that transforms CH3OCCH3 into CH3OCCl stabilizes the carbene and decreases its reactivity, especially toward the electron-poor alkenes of Table 7.5, where it can be seen that CH3OCCl is 80 times less reactive than CH3COCH3. A similar attenuation of reactivity accompanies the Cl for C6H5 swap that converts C6H5COCH3 into CH3OCCl.

288 SINGLET CARBENES

Moreover, we see that the nucleophilic properties of CH3COCH3 are pronounced (Tables 7.3 and 7.4): the combination of a relatively high-lying, filled HOMO donor orbital ( 9.4 eV)33 and a high-lying, poorly accessible vacant p orbital (LUMO, 4.0 eV)33 makes CH3COCH3 a good nucleophile and a poor electrophile. The computed differential orbital energies for additions of CH3COCH3 to the alkenes of Table 7.4 are dominated by the nucleophilic eN term.

The CH3O for CH3 substituent swap that transforms CH3COCH3 into (CH3O)2C results in additional stabilization of the carbene and lower reactivity. Thus, (CH3O)2C, with eLU ¼ 4:3 eV,40 no longer adds to electron-rich alkenes such as (CH3)2C CHCH3, dimerizing instead, whereas its nucleophilic additions to

electron-poor olefins, such as CH2 CClCN and CH2 CHCN, are 100–1000 times slower than those of CH3COCH3.33,40

If the resonance stabilization of singlet carbenes is increased even more, for example, by replacing the CH3O substituents of (CH3O)2C by amino groups [as in (R2N)2C] or phosphorus substituents (13), stable, isolable, nucleophilic carbenes can be obtained. These species are considered in Chapter 8 in this volume.

Given our ability to measure absolute rate constants for carbene additions, variable temperature studies readily afford activation parameters.57,63 Initial studies of C6H5CX additions gave very low Ea values, 1 kcal/mol for reactions with 1-hexene and trans-pentene.66 Most surprisingly, the Ea values for C6H5CCl additions to (CH3)2C C(CH3)2 and (CH3)2C CHCH3 were negative ( 1.7 and 0.8 kcal/ mol, respectively). The reaction rates increased as temperature decreased. The preexponential (A) factors were low (2–6 107 M 1 s 1), indicative of an unfavorable entropy of activation.

The most convincing explanation for the negative activation energies was offered by Houk et al.,67 who computed H, S, and G for additions of CBr2, CCl2, and CF2 to (CH3)2C C(CH3)2 and (CH3)2C CH2. With CBr2, a very reactive carbene, and the highly reactive (CH3)2C CHCH3, the addition reaction was so exothermic that H continually decreased all along the reaction coordinate. Therefore, Hz and Ea were negative. However, the loss of the translational, vibrational, and rotational entropy required in the transition state for addition led to an unfavorable and dominant entropy of activation ð Sz < 0Þ which, in turn, created a free energy barrier to addition ð Gz > 0Þ. With the more stable and less reactive carbene, CF2, Ea, and Hz were positive, adding to the Gz barrier.67 The key role

of entropy in carbenic additions, highlighted by Houk, was earlier noted by Skell and Cholod,35b and by Giese.68

A wide-ranging experimental study of carbene–alkene addition activation parameters examined a series of substituted arylhalocarbenes (16) in reactions with (CH3)2C C(CH3)2 and 1-hexene (Fig. 7.9).69

YC X

16

Figure 7.9

MORE DETAILED TREATMENTS OF STRUCTURE AND PROTOTYPAL REACTIONS

289

‘‘Extreme’’ cases were reactions of the least stabilized, most reactive carbene (Y ¼ CF3, X ¼ Br) with the more reactive alkene (CH3)2C C(CH3)2, and the most

 

 

 

 

 

 

alkene

stabilized, least reactive carbene (Y ¼ CH3O, X ¼ F) with the less reactive

9

 

(1-hexene). The rate

constants, as measured by LFP, were 1:7 10

 

and

5:0

 

 

67

 

an interval of 34,000. In agreement

 

104 M 1 s 1, respectively, spanning

with

Houk’s ideas,

 

the reactions were

entropy dominated ð Sz 22 to

 

 

 

Þ

 

 

 

 

 

 

29 e:u: . The Gz barriers were 5.0 kcal/mol for the faster reaction and 11 kcal/

mol for the slower reaction, mainly because of entropic contributions; the Hz components were only 1.6 and þ2.5 kcal/mol, respectively.69 Despite the dominance of entropy in these reactive carbene addition reactions, a kind of ‘‘de facto enthalpic control’’ operates. The entropies of activation are all very similar, so that in any comparison of the reactivities of alkene pairs (i.e., krel), the rate constant ratios reflect differences in Hz, which ultimately appear in Gz. Thus, carbenic philicity, which is the pattern created by carbenic reactivity, behaves in accord with our qualitative ideas about structure–reactivity relations, as modulated by substituent effects in both the carbene and alkene partners of the addition reactions.24

Finally, volumes of activation were measured for the additions of C6H5CCl to (CH3)2C C(CH3)2 and trans-pentene in both methylcyclohexane and acetonitrile.70 The measured absolute rate constants increased with increasing pressure;Vz ranged from 10 to 18 cm3/mol and were independent of solvent. These results were consistent with an early, and not very polar transition state for the addition reaction.

2.1.3. Symmetry of the Transition State. The asymmetric, ‘‘non-least-

motion’’ transition state (6) for carbene–alkene cycloaddition has been discussed in qualitative1,71 and quantitative34,36,67,72 terms for many years. As shown in Fig. 7.6, the transition state features carbene p–alkene p* orbital interactions.24–26,36,42–46

Hoffmann first presented a theoretical analysis supporting a ‘‘p-approach’’ transition state (6), which also rationalizes the electrophilic character of CH2 or CCl2 addition reactions.34 He also showed that the symmetrical ‘‘least-motion’’ approach

(17) was forbidden by orbital symmetry.34 Subsequent computational studies supported Hoffmann’s conclusions.36,67,72 Note that although the trajectory of the addi-

tion reaction begins with the geometry represented by 6, the carbene must later pivot toward geometry 17 in order to complete the cyclopropanation (Fig. 7.10).

Keating et al.73 used natural abundance isotope effects to define the ‘‘direction’’ of the asymmetry for the addition of CCl2 to 1-pentene, and presented complementary

(0.943) H

 

H (0.986)

Cl

Cl

C1

 

C2

C

 

 

H

H

(0.964) H

 

C3H7

C

C

 

H

C3 H7

(1.025)

(1.005)

 

 

 

17

18

19

 

Figure 7.10

290 SINGLET CARBENES

B3LYP/6-31G* computations. Thus, 3 mol of 1-pentene was cyclopropanated to 89–93% conversion with CCl2 (generated by the phase-transfer-catalyzed reaction of CHCl3 and NaOH). The unreacted 1-pentene was isolated and brominated, and the resulting dibromide was analyzed by 13C and 2H nuclear magentic resonance (NMR) spectroscopy. The C and H isotopic abundances were compared to those in dibromide derived from a sample of the initial 1-pentene, leading to the isotope effects shown for a typical experiment in structure 18 (Fig. 7.10).

Of course, the transition state for addition to an unsymmetrical alkene must be unsymmetrical in some fashion. Thus, the 12C/13C primary kinetic isotope effects at C1 and C2 are unequal. However, the larger isotope effect at C1 implies more advanced binding at this carbon atom in the transition state (19). Note too the inverse a–secondary (kH/kD) isotope effects at the hydrogens bonded to C1 and C2. These carbons rehybridize from sp2 toward sp3 as the addition proceeds, so that kH/kD < 0 is expected. The greater effect at H-C1 versus H-C2 is also consistent with greater carbene binding to the alkene’s terminal C in the transition state.73

The experimental results agree with electrophilic addition of CCl2 to 1-pentene. Stronger bonding at C1 in the transition state will impose most of the partial positive charge at C2, the more substituted carbon, where it can be better dispersed by inductive and hyperconjugative electron release from the alkyl substituent. The B3LYP/6-31G* (and 6-311þG*) computations for the addition of CCl2 to propene support the experimental results. The optimum computed transition state geometry

is shown

in 20

74

for a carbene–alkene separation of

˚

 

 

2.4 A (Fig. 7.11). The

computed Gz

 

for addition via 20 is

 

10.1 kcal/mol,

preferable to alternative

geometry

21,

74

 

 

 

 

 

73

 

in which Gz 12.6 kcal/mol at a separation of 2.3 A.˚

 

2.190

2.766

2.591

2.189

90.7°

 

 

99.6°

 

 

 

 

1.362

1.360

 

 

r = 2.4 Å

r = 2.3 Å

 

20

21

 

Figure 7.11. B3LYP/6-31G* geometries for CCl2 plus propene mode A (structure 20) and mode B (structure 21) addition at constrained values of r. Bond lengths are given in angstroms.

The energetically ‘‘preferred’’ direction of the asynchronous CCl2 addition transition state with propene (20) places the carbene’s s orbital closer to C1, and the substituents on the carbene directed toward C2.73,75 The asymmetry of bonding to the carbene is evident in 20, and computed carbon and hydrogen isotope effects for

20 ˚

(and for an analogous structure with a 2.5-A carbene–alkene separation) agree very well with the experimental isotope effects.73

In Section 2.1, we have not considered the vast literature devoted to the addition reactions of carbenoids.76 Typically, these compounds are organometallic molecules such as (C6H5)2CBrLi, C6H5CHBrLi, LiCCl3, and ICH2ZnI (Simmons–Smith reagent) that react with olefins to give cyclopropanes just as do the corresponding

MORE DETAILED TREATMENTS OF STRUCTURE AND PROTOTYPAL REACTIONS

291

free carbenes, (C6H5)2C, C6H5CH, CCl2, and CH2, respectively. Carbenoids usually add electrophilically to alkenes. An extensive review of carbenoid philicity has recently been published by Boche and Lohrenz.77

2.1.4. Stepwise versus Concerted Addition. One of the factors that makes carbene chemistry fascinating is that regardless of which species is first formed, reactivity may issue from either the singlet or triplet state. In this chapter (and in Chapter 9 in this volume), we will have to consider the spin state of the reacting species in essentially all reactions. It is not enough to assume that the first-formed intermediate is the reacting species. Indeed, the singlet generally has an enormous advantage over the triplet. Nearly all molecules are ground-state singlets—all electrons are paired. Thus, they can be formed directly from singlets but not from triplets. Reactions of singlet carbenes are often exceptionally exothermic, and the formation of two new bonds in a single step from a singlet implies transition states that are very reactant-like (Hammond postulate), and thus with very low barriers. Nearly every time, when a singlet is close enough in energy to be formed from a triplet, it will be the intermediate that does the chemistry and forms the products. We will point out a few exceptions in this chapter, and be sure to see Chapter 9 in this volume as well.

As noted above, singlet reactions in general, and addition reactions in particular,

must benefit greatly from the exothermicity of the formation of two sigma bonds. That idea must have been behind the early assumption, made by both the Skell27,35c

and Doering groups,78 that the observation of a stereospecific addition would be diagnostic for the singlet state. Of course, that assumption not only posits that the addition of a singlet carbene would be one step, but that the necessarily stepwise addition of the triplet would not be stereospecific. In turn, those ideas assume that in the intermediate diradical formed in the first step of a triplet addition, rotation about a single bond would be faster than spin inversion (intersystem crossing—‘‘isc’’) and closure Scheme 7.8).

 

 

intersystem

 

 

 

 

crossing =

 

 

 

(singlet)

R2C

"spin flip" = "isc"

R2C

(triplet)

 

 

 

H3C

CH3

 

H3C

 

CH3

 

R

CH3

 

R

 

CH3 CH3

 

CH3 1. "isc"

C

 

 

C

 

 

 

 

 

 

2. close

R

 

 

 

R

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

R

 

 

CH3

1. "isc"

R

CH3

 

 

 

 

 

 

 

 

C

 

C

 

 

 

 

 

2. close

 

 

 

 

 

 

 

R

 

 

 

CH3

R

CH3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Scheme 7.8

292 SINGLET CARBENES

Surprisingly, the critical experiment has been done infrequently over the last one-half of a century! The requirements for an experiment that truly speaks to the issue at hand are that one be able to see the results of addition of both spin states of a single carbene, and these requirements rarely have been met. For example, the direct irradiation of methyl diazomalonate leads to the stereospecific addition expected of a singlet carbene, whereas the photosensitized decomposition of the diazo compound leads to formation of the triplet carbene and loss of the stereochemical relationship originally present in the reacting alkene. Rotational equilibration in the intermediate seems to be complete, as it makes no significant difference whether cis or trans alkene is used as starting material (Scheme 7.9).79

 

 

 

 

 

 

CH3

CH3OOC

CH3

 

 

 

 

 

CH3OOC

CH3

C

 

 

 

 

 

 

C

 

 

 

 

 

 

 

 

CH3

 

 

 

 

 

CH3OOC

CH3OOC

 

 

 

 

 

 

 

 

 

 

 

 

H3 C CH(CH3)2

 

 

(CH3 OOC)2C

 

 

 

 

92

 

8

 

 

 

 

 

 

 

hν

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(CH3 OOC)2CN2

hν photosensitization

 

 

 

H3 C

CH(CH3)2

 

(CH3OOC)2C

 

 

 

 

10

90

 

 

 

 

 

 

(CH3 OOC)2C

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

10

90

 

 

 

 

 

 

 

 

 

 

 

 

H3 C

 

 

hν

 

 

CH(CH3)2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(CH3 OOC)2CN2

 

 

 

 

 

 

 

hν

 

 

H3 C

 

 

 

 

 

 

photosensitization

 

 

 

 

 

 

 

 

CH(CH3)2

 

(CH3OOC)2C

 

 

 

 

 

 

 

 

 

14

86

 

 

 

Scheme 7.9

MORE DETAILED TREATMENTS OF STRUCTURE AND PROTOTYPAL REACTIONS

293

The critcal experiment has also been attempted many times for the parent carbene, methylene. The difficulties of working in the gas phase, in which productscrambling hot molecule rearrangements are all-too common, and of finding a good source of the triplet carbene in solution meant that a truly convincing experiment waited until 1987. Turro et al.80 decomposed diazirine through direct and photosensitized irradiation to refine a much earlier experiment of Hammond et al.81 that used diazomethane as methylene source. The data shown in Scheme 7.10 are consistent with what the ‘‘Skell hypothesis’’ would dictate: Reactions of the singlet state are stereospecific, whereas those of the lower energy triplet scramble the stereochemical relationships present in the reactant alkene. In this case, rotation appears not to be significantly faster than intersystem crossing and closure.

 

 

H C

CH2CH3 H C

CH2CH3

 

 

3

3

 

H3 C

CH2CH3

CH2

 

CH2

 

 

 

 

 

40

 

<1

h ν

 

 

HN

H

N

 

H3 C

CH2CH3

 

 

 

 

 

 

 

 

44

51

 

 

 

h ν photosensitization

 

 

 

 

 

 

 

 

 

CH2CH3

 

 

 

 

H3 C

 

 

<1

38

 

 

 

 

h ν

 

 

 

 

 

 

H

N

 

 

CH2CH3

 

 

 

 

 

 

 

 

HN

H3 C

18

79

 

h ν photosensitization

Scheme 7.10

It is also possible to induce intersystem crossing from an initially formed singlet

state to a lower energy triplet by forcing the singlet state to suffer collisions with an inert medium (no easy task for a species as reactive as methylene!).80,82

Despite the somewhat thin underpinnings, the observation of stereospecific or nonstereospecific addition remains the most widely used experimental test of singlet or triplet reactivity.

Соседние файлы в предмете Химия