Next Article in Journal
Antibiotic Producing Potentials of Three Freshwater Actinomycetes Isolated from the Eastern Cape Province of South Africa
Previous Article in Journal
Sequential Events in the Irreversible Thermal Denaturation of Human Brain-Type Creatine Kinase by Spectroscopic Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of the Nucleophilic Solvation Effects in Isopropyl Chlorothioformate Solvolysis

by
Malcolm J. D’Souza
1,*,
Brian P. Mahon
1 and
Dennis N. Kevill
2,*
1
Department of Chemistry, Wesley College, 120 N. State Street, Dover, DE 19901-3875, USA
2
Department of Chemistry and Biochemistry, Northern Illinois University, DeKalb, IL 60115-2862, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2010, 11(7), 2597-2611; https://doi.org/10.3390/ijms11072597
Submission received: 9 June 2010 / Revised: 21 June 2010 / Accepted: 28 June 2010 / Published: 29 June 2010
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)

Abstract

:
Correlation of the solvent effects through application of the extended Grunwald-Winstein equation to the solvolysis of isopropyl chlorothioformate results in a sensitivity value of 0.38 towards changes in solvent nucleophilicity (l) and a sensitivity value of 0.72 towards changes in solvent ionizing power (m). This tangible l value coupled with the negative entropies of activation observed indicates a favorable predisposition towards a modest rear-side nucleophilic solvation of a developing carbocation. Only in 100% ethanol was the bimolecular pathway dominant. These observations are very different from those obtained for the solvolysis of isopropyl chloroformate, where dual reaction channels were proposed, with the addition-elimination reaction favored in the more nucleophilic solvents and a unimolecular fragmentation-ionization mechanism favored in the highly ionizing solvents.

1. Introduction

The application [1] of the simple [2] and extended [3] (two-term) Grunwald-Winstein Equations for the analyses of the specific rates of solvolyses of chloroformates [424], and their corresponding sulfur-for-oxygen substituted analogs (thiono-, thio-, and dithiochloroformate esters) [1,10,21,2530] has proven to be immensely beneficial [1] for the consideration of solvent effects upon reactions of these important intermediates that are found in pharmaceutical [31] and agricultural [32] products, and as derivatization agents in the flavor and fragrance industries [33]. In the simple [2] (Equation 1) and extended [3] (Equation 2) Grunwald-Winstein Equations, k and ko are the specific rates of solvolysis in a given solvent and in the standard solvent (80% ethanol), respectively, m represents the sensitivity to changes in the solvent ionizing power Y (initially set at unity for tert-butyl chloride solvolyses), and c is a constant (residual) term. In Equation 2, the additional lN term was added to include a term governed by the sensitivity l to changes in solvent nucleophilicity (N). When using Equation 1 for ionization (SN1 + E1) mechanisms, it was realized both that the ionizing scales are leaving-group dependent and that adamantyl derivatives provide better standard substrates, and now for each commonly used leaving group X a series of YX values [3436] are available. For bimolecular (SN2 and/or E2) reactions, NT scales based on the solvolyses of the S-methyldibenzothiophenium ion [37,38] have become the recognized standards for considerations of solvent nucleophilicity. To correct for dispersions in Grunwald-Winstein plots observed whenever the reaction center is adjacent to a π-system [39,40], or in α-haloalkyl aryl compounds that proceed via anchimeric assistance (kΔ) [41], Kevill and D’Souza proposed [1,4245] the addition of an additional aromatic ring parameter (hI) term to Equations 1 and 2 to give Equations 3 and 4. In Equations 3 and 4, h represents the sensitivity of solvolyses to changes in the aromatic ring parameter I. Recently, it was stressed [45] that the additional hI term can only be applied in situations where there is the presence of conjugated π-electrons adjacent to the developing carbocationic center, or in substrates that proceed with charge delocalization via a 1,2-aryl shift.
log ( k / k o ) = m Y + c
log ( k / k o ) = l N + m Y + c
log ( k / k o ) = m Y + h I + c
log ( k / k o ) = l N + m Y + h I + c
After surveying the literature [29,4654] it was pointed out [29,31], that the most stable geometries for RXCXCl (where X = S or O) exist in a configuration where the C=O or the C=S is syn with respect to R. Furthermore, studies [55] based on high resolution microwave spectra have shown that the s-isopropyl thioesters have a higher torsional barrier when compared to the s-ethyl thioesters. In Figure 1, the molecular structures for s-isopropyl chlorothioformate (1), s-isopropyl chloroformate (2), and s-phenyl chloroformate (3) are shown where the halogen atom is in a trans position with respect to the alkyl or aryl group i.e., the structures are drawn in a syn conformation.
The solvolysis of phenyl chloroformate (PhOCOCl, 3) depicted in Scheme 1 has been extensively studied [7,21] over the full range of the types of solvent usually incorporated into studies utilizing the extended Grunwald-Winstein equation (Equation 2). The observed sensitivity values of l = 1.66 and m = 0.56 are now taken as typical values [1,7,10,1216,18,2028,30,43] for attack at an acyl carbon proceeding by the addition-elimination mechanism, with the addition step being rate-determining. A sulfur-for-oxygen substitution for both oxygen atoms in 3 yields the dithio analog, phenyl dithiochloroformate (PhSCSCl). Recent application [21,25] of Equations 2 and 4 to the solvolysis of PhSCSCl generated large l values of 0.69, and 0.80, respectively, and m values of 0.95 using Equation 2, and 1.02 using 4. The appreciable l values signify a profound rear-side nucleophilic solvation occurring at the developing resonance stabilized carbocation. Also, the h value of 0.42 ± 0.15 obtained [21] using Equation 4 suggests a minimal charge delocalization into the aromatic ring. The substitution of the ether oxygen of 3 with sulfur yields phenyl chlorothioformate (PhSCOCl), and the substitution of the carbonyl oxygen of 3 with sulfur, produces phenyl chlorothionoformate (PhOCSCl). It was shown [21,25,27,28,56] that PhSCOCl and PhOCSCl solvolyze with very similar specific rates of solvolysis by dual mechanisms, that have an essential identical division [25] into the addition-elimination (A–E) and ionization (SN1) pathways as the solvent is varied. This suggests that the inductive effect of the phenoxy group in PhOCSCl is counterbalanced by the conjugative release of the thiocarbonyl group. The reverse occurs in PhSCOCl, where the mesomeric effect of the thiophenoxy substituent is redressed by the electron-withdrawing character of the carbonyl group.
On the other hand, superimposed mechanisms [1,10,12,13,16,1821,24,30,5762] were observed for the alkyl chloroformate (ROCOCl) and chlorothioformate (RSCOCl) esters and the ranges of dominance for each were found to be heavily dependent on the nature of the substituent (R), solvent nucleophilicity, and on the ionizing ability of the solvent. These solvolyses could be treated, by application of Equation 2, in terms of two correlations with very different sets of l and m values, believed to correspond to the side-by-side operation of the bimolecular addition-elimination and unimolecular ionization (SN1) pathways. Queen’s original proposal [5860] for the hydrolysis of alkyl thiochloroformates was for the operation of a dominant SN1 mechanism. This proposition was based on observations of studies of the hydrolysis of various alkyl thiochloroformates, chloroformates, and carbamoyl chlorides, including determination of their hydrolysis activation parameters and the solvent deuterium isotope effects. Lee et al. [61,62] favored an SN1 mechanism for the methyl dithiochloroformate ester (MeSCSCl), and competing SN1 and SN2 pathways for methyl chlorothioformate, (MeSCOCl), and methyl chlorothionoformate, (MeOCSCl), with the extent of operation of each mechanism being dependent on the proportion of water in the aqueous solvents. Castro, Santos and co-workers [6368] advanced a stepwise mechanism through a zwitterionic tetrahedral intermediate for a variety of thio-, dithio-, and thiono- analogs of carboxylic acids.
In the present paper, we report the first-order specific rate constants at 25.0 °C for the solvolyses of 1 in ethanol and methanol and thirteen binary mixtures of aqueous ethanol (EtOH), aqueous methanol (MeOH), aqueous acetone, aqueous 2,2,2-trifluoroethanol (TFE), and aqueous 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP). Studies were also carried out in five mixtures of TFE and EtOH. Rate constants at additional temperatures and the calculated activation parameters are also reported.

2. Results and Discussion

As reported in Table 1, the first-order specific rate constants for 1 increase as the amount of water in the binary organic mixtures increases. This upward rise in the rate coefficients coupled with the incremental water content is also observed in the highly ionizing aqueous fluoroalcohol mixtures. These observations are consistent with an ionization mechanism (SN1) with some nucleophilic participation of the solvent in the transition state. In Table 1, we report the rate data for 1 in twenty solvents together with the literature values for NT [37,38] and YCl [3436]. Additionally for 1, rates were measured at 35.0 °C and 45.0 °C for 100% MeOH, 100% EtOH, 80% EtOH, and 90% acetone, and at 15.0 °C and 20.0 °C for 97% TFE, 70% TFE, 90% HFIP, and 70% HFIP. Included in the footnotes of Table 1, are the derived Arrhenius parameters (ΔH and ΔS) at 25.0 °C for 100% MeOH, 100% EtOH, 80% EtOH, 90% acetone, 97% TFE, 90% TFE, 90% HFIP, and 70% HFIP. Also reported are the specific rate ratios for 1 and 2 and previously tabulated ratios for the corresponding ethyl [10] and phenyl [26] esters.
In Table 2, the Grunwald-Winstein l, m, and c values, the goodness-of-fit, and F-test parameters for a variety of ROCOCl, RSCSCl, ROCSCl, and RSCOCl solvolyses are reported. The l and m values obtained for MeOCOCl are very similar to those obtained for PhOCOCl, hence it was suggested [12] that, due to the strong electron-withdrawing ability of the methoxy group, MeOCOCl solvolyses by a dominant bimolecular addition-elimination mechanism in all solvents except 90% HFIP. In MeSCOCl, the mesomeric effect of the thiomethoxy group predominates in the highly ionizing aqueous fluoroalcohols and 60% acetone resulting in a unimolecular mechanism [30] with a profound rear-side nucleophilic solvation of the developing carbocation. In the more nucleophilic solvents, the mechanism switches to a bimolecular addition-elimination with the addition step being rate determining. It was also shown [30] that there was a tendency for both reaction channels to operate simultaneously in MeSCOCl. Increasing the chain length by a single carbon in ethyl chloroformate, (EtOCOCl), diminishes the alkoxy group’s inductive ability slightly, and dual reaction channels [10] are followed, with an addition-elimination in the more nucleophilic solvents and a SN1 pathway with appreciable nucleophilic solvation of the developing acylium ion, in the aqueous fluoroalcohols. In ethyl chlorothioformate, (EtSCOCl), the relative importance of the two reaction channels is reversed [10], where the unimolecular ionization pathway is followed in a majority of the solvents and an addition-elimination channel in methanol, ethanol, and 90% ethanol. The influence of a branch-chain alkyl group as in isopropyl chloroformate, (2, i-PrOCOCl) [24], induces a counterbalance between the hyperconjugative mesomeric effect and the inductive ability of the alkyl group, resulting in a bimolecular addition-elimination pathway in the more nucleophilic solvents and a unimolecular fragmentation-ionization pathway in which ionization is accompanied by loss of carbon dioxide, in the majority of the solvents studied (Scheme 2).
A correlation using all twenty data points listed in Table 1 within the simple Grunwald-Winstein equation (Equation 1) resulted in a rather poor correlation coefficient R = 0.930, m value of 0.43 ± 0.04, c value of −0.44 ± 0.09, and a F-test value of 114. The same set of solvents using Equation 2 yields, l = 0.30 ± 0.13, m = 0.63 ± 0.09, c = −0.26 ± 0.11, R = 0.947, and a F-test value of 74. Using Equation 1 and omitting the rate constant in 100 EtOH results in, m = 0.45 ± 0.04, c = −0.49 ± 0.10, R = 0.933, and F-test = 115. For the identical 19 solvents, using Equation 2, we obtained a much improved correlation coefficient R = 0.961, l = 0.38 ± 0.11, m = 0.72 ± 0.09, c = −0.28 ± 0.10, and an F-test value of 97. The improvements observed using Equation 2, in the correlation coefficient and the F-test values on omission of the 100 EtOH data point, is a strong indicator that there is a superimposed bimolecular pathway in this solvent. Utilizing the equation 0.38 NT + 0.72 YCl −0.28 for log (k/ko), we estimate the ionization rate (k) in 100% EtOH to be, 0.158 × 10−5 s−1. This value indicates that, in 100% ethanol, 87% of the reaction is occurring by the bimolecular channel. For 90% EtOH and 90% acetone, the corresponding values are 44% and 34% respectively.
With use of Equation 2, the l and m values for solvolyses of 1 imply that the stepwise ionization pathway is accompanied by a modest rear-side nucleophilic solvation of the developing SN1 transition state. As in the case of MeSCOCl [30], the entropies of activation (ΔS values) for 1 reported in the footnotes of Table 1, are negative. The negative ΔS values suggest a greater ordering of the solvent molecules in the transition state, encasing the developing carbocation and thus dispersing the electrical charge. From the data in Table 2, the observed trend for the l values as one progresses with the increasing carbon chain length, requires a necessary decrease in nucleophilic solvation requirements as one moves from MeSCOCl to 1, due to an increase in the hyperconjugative mesomeric effects. The l value of 0.79 ± 0.06 and m value of 0.85 ± 0.07 obtained for MeSCOCl [30] are similar to the l value of 0.83 ± 0.06, and m = 0.70 ± 0.04 observed for the solvolysis of acetyl chloride [69], whereas the l value of 0.38 ± 0.11 and m value of 0.72 ± 0.09 obtained for 1, are similar to l = 0.34 ± 0.04 and m = 0.84 ± 0.02 obtained for t-butyl chloride [35,45]. A plot of log (k/ko)1 versus 0.38 NT + 0.72 YCl as shown in Figure 2, is consistent with the previous discussion of the side-by-side operation of the addition-elimination and stepwise ionization (with nucleophilic solvation) mechanism in a few of the more nucleophilic solvents.
Queen et al. [59] found 2-propanethiol as a product of the hydrolysis of 1. Since they found no appreciable amounts of isopropyl alcohol, propene, or carbonyl sulfide, they suggested that there was no alkyl-sulfur bond fission in the hydrolysis of 1. These observations are consistent with the results outlined above and with the mechanistic scheme (Scheme 3) shown below.
The very different responses of the specific rate ratios for RSCOCl/ROCOCl to solvent variation depending on whether the R group is phenyl, ethyl, or isopropyl (Table 1) are nicely consistent with the conclusions drawn from the application of the extended Grunwald-Winstein equation (Equation 2). When R = Ph, the very low values (0.026–0.042) for solvolyses in the non-fluoroalcohol-containing solvents are consistent with the proposed addition-elimination mechanism, with the carbonyl carbon more prone to the rate-determining addition when the better electron-withdrawing phenoxy group is attached. In solvents with high fluoroalcohol content the ratio increases in value and reaches values of 5.9 for 97% TFE and 16 for 90% HFIP, where a superimposed SN1 mechanism becomes dominant for PhSCOCl solvolyses.
A different pattern of behavior is observed [10] for R = Et (Table 1), with ratios varying from 0.19 (100% ethanol) to 357 (90% HFIP). Both reaction channels are observed with the unimolecular (SN1) covering the larger range of solvent composition for EtSCOCl and the addition-elimination the larger range of solvent composition for EtOCOCl. The small (0.19 to 0.40) values in MeOH and EtOH and in MeOH-H2O and EtOH-H2O mixtures can be considered to reflect the dominance of the addition-elimination mechanism and the much larger (89–357) values in TFE-H2O and HFIP-H2O mixtures the dominance of the ionization (SN1) mechanism for both substrates.
The values for the i-PrSCOCl/i-PrOCOCl specific rate ratios behave in yet a different manner. There is very little variation in the specific rate ratio over the full range of solvents considered when the presently determined values for 1 are compared with the values reported earlier [24] for 2. Since both show a moderate l value and a fairly high m value over the more ionizing solvents, it would appear that both are reacting by an ionization mechanism. In this case, the values of in excess of over 100 when ionization is believed to be followed for the ethyl derivatives might be expected to be applicable also to the isopropyl derivatives. In reality, the ratio (Table 1) only reaches a value of just over ten. We tentatively propose that this is a consequence of the rate-determining step producing the carboxylium ion for 1 (Scheme 3). This view is supported by the observation of Queen and coworkers [59] of 2-propanethiol as the major product for the hydrolyses of 1, an observation that requires the retention of the isopropyl-sulfur bond throughout the solvolysis process. This process is probably a consequence of the sulfur being able to support appreciable positive charge in the resonance-stabilized carboxylium ion. If the corresponding process operated for the carboxylium ion from 2, the positive charge would be partially transferred to the oxygen (oxonium ion character). This would lead to the electronegative oxygen exerting a strong pull on the electrons of the isopropyl-oxygen bond which is apparently sufficient for the mechanism of the rate-determining step to be best described as in Scheme 2, with a rate-determining ionization-fragmentation process. To dominate, this process must be favored relative to the alternative process producing the carboxylium ion from 2 (parallel to the process of Scheme 3). This leads to a reduction the i-PrSCOCl/i-PrOCOCl ratio relative to that predicted if both proceeded with a carboxylium ion formation as the rate-determining step.

3. Conclusions

It is becoming clear that the solvolyses of alkyl chlorothioformates involve simultaneous side-by-side reaction channels where the mesomeric effect governs the dominant mechanism, which varies with the ionizing ability of the solvent. For the solvolyses of isopropyl chlorothioformate (1) in all solvents but 100% EtOH, we propose a dominant stepwise SN1 mechanism with discernible rear-side nucleophilic solvation of the resonance stabilized acylium ion (Scheme 3). Conversely in isopropyl chloroformate (2), the inductive pull of the isopropoxy group is paramount, and dual distinct channels are observed, with a bimolecular addition-elimination in the more nucleophilic solvents and a unimolecular fragmentation-ionization mechanism (Scheme 2) in the highly ionizing solvents. The extended Grunwald-Winstein equation was able to differentiate between such mechanisms and it continues to play a major role in studies using linear free energy relationships.
The very different responses of the RSCOCl/ROCOCl ratio to changes in the R group are nicely consistent with a predominantly addition-elimination process when R is Ph, perturbed for highly ionizing solvents by the incursion of an ionization component for the PhSCOCl solvolyses. When R is Et, both reaction channels are observed for each substrate but with the unimolecular dominant for EtSCOCl and the addition-elimination for EtOCOCl. The lowest values for the ratio are for the addition-elimination pathway and the highest values for the ionization pathway. Similar low values are observed when R is i-Pr but with the range where the addition-elimination mechanism is dominant being very limited for i-PrSCOCl. At first sight, a surprising aspect is that, in the highly ionizing solvents, the ratio has a much lower value than when R = Et. This is rationalized in terms of the i-PrOCOCl involving a favored ionization-fragmentation process, which must therefore be faster than what the carboxylium ion forming process would be, with a reduction in the value of the k1/k2 ratio, relative to the corresponding value when R = Et. This rationalization is nicely consistent with the observation that the s-isopropyl thioesters have a higher torsional barrier when compared to the s-ethyl thioesters [55].

4. Experimental Section

The isopropyl chlorothioformate (96%, Sigma-Aldrich) was used as received. Solvents were purified and the kinetic runs carried out as described previously [10]. A substrate concentration of approximately 0.005 M in a variety of solvents was employed. For some of the runs, calculation of the specific rates of solvolysis (first-order rate coefficients) was carried out by a process in which the conventional Guggenheim treatment [70] was modified [71] so as to give an estimate of the infinity titer, which was then used to calculate for each run a series of integrated rate coefficients. The specific rates and associated standard deviations, as presented in Table 1, are obtained by averaging all of the values from, at least, duplicate runs.
Multiple regression analyses were carried out using the Excel 2007 package from the Microsoft Corporation, and the SigmaPlot 9.0 software version from Systat Software, Inc., San Jose, CA, was used for the Guggenheim treatments.

Acknowledgments

This research was supported by grant number 2 P2O RR016472-010 from the National Center for Research Resources (NCRR), a component of the National Institutes of Health (NIH). This IDeA Network of Biomedical Research Excellence (INBRE) grant to the state of Delaware (DE) was obtained under the leadership of the University of Delaware, and the authors sincerely appreciate their efforts.

References and Notes

  1. Kevill, DN; D’Souza, MJ. Sixty Years of the Grunwald-Winstein Equation: Development and Recent Applications. J Chem Res 2008, 61–66. [Google Scholar]
  2. Winstein, S; Grunwald, E; Jones, HW. The Correlation of Solvolyses Rates and the Classification of Solvolysis Reactions into Mechanistic Categories. J. Am. Chem. Soc 1951, 73, 2700–2707. [Google Scholar]
  3. Grunwald, E; Winstein, S. The Correlation of Solvolysis Rates. J. Am. Chem. Soc 1948, 70, 846–854. [Google Scholar]
  4. Kevill, DN; Kyong, JB; Weitl, FL. Solvolysis-Decomposition of 1-Adamantyl Chloroformate: Evidence for Ion Pair Return in 1-Adamantyl Chloride Solvolysis. J. Org. Chem 1990, 55, 4304–4311. [Google Scholar]
  5. Kyong, JB; Suk, YJ; Kevill, DN. Solvolysis-Decomposition of 2-Adamantyl Chloroformate: Evidence for Two Reaction Pathways. J. Org. Chem 2003, 68, 3425–3432. [Google Scholar]
  6. Koo, IS; Yang, K; Kang, K; Oh, HK; Lee, I. Stoichiometric Solvation Effects. Product-Rate Correlation for the Solvolyses of Phenyl Chloroformate in Alcohol-Water Mixtures. Bull. Korean Chem. Soc 1996, 17, 520–524. [Google Scholar]
  7. Kevill, DN; D’Souza, MJ. Correlation of the Rates of Solvolysis of Phenyl Chloroformate. J Chem Soc., Perkin Trans 2 1997, 1721–1724. [Google Scholar]
  8. Koo, IS; Yang, K; Koo, JC; Park, J-K; Lee, I. Stoichiometric Solvation Effects. Part 4. Product–rate Correlations for Solvolyses of p-Methoxyphenyl Chloroformate in Alcohol–Water Mixtures. Bull. Korean Chem. Soc 1997, 18, 1017–1021. [Google Scholar]
  9. Koo, IS; Yang, K; Kang, K; Lee, I; Bentley, TW. Stoichiometric Solvation Effects. Part 3. Product–rate Correlations for Solvolyses of p-Nitrophenyl Chloroformate in Alcohol–Water Mixtures. J Chem Soc, Perkin Trans 2 1998, 1179–1183. [Google Scholar]
  10. Kevill, DN; D’Souza, MJ. Concerning the Two Reaction Channels for the Solvolyses of Ethyl Chloroformate and Ethyl Chlorothioformate. J. Org. Chem 1998, 63, 2120–2124. [Google Scholar]
  11. Koo, IS; Yang, K; Kang, K; Lee, I. Transition-State Variation in the Solvolyses of para-Substituted Phenyl Chloroformates in Alcohol-Water Mixtures. Bull. Korean Chem. Soc 1998, 19, 968–973. [Google Scholar]
  12. Kevill, DN; Kim, JC; Kyong, JB. Correlation of the Rates of Methyl Chloroformate with Solvent Properties. J Chem Res Synop 1999, 150–151. [Google Scholar]
  13. Kyong, JB; Kim, YG; Kim, DK; Kevill, DN. Dual Pathways in the Solvolyses of Isopropyl Chloroformate. Bull. Korean Chem. Soc 2000, 21, 662–664. [Google Scholar]
  14. Park, KH; Kyong, JB; Kevill, DN. Nucleophilic Substitution Reactions of p-Nitrobenzyl Chloroformate. Bull. Korean Chem. Soc 2000, 21, 1267–1270. [Google Scholar]
  15. Kyong, JB; Park, B-C; Kim, C-B; Kevill, DN. Rate and Product Studies with Benzyl and p-Nitrobenzyl Chloroformates under Solvolytic Conditions. J. Org. Chem 2000, 65, 8051–8058. [Google Scholar]
  16. Kevill, DN; D’Souza, MJ. Correlation of the Rates of Solvolyses of n-Octyl Fluoroformate and a Comparison with n-Octyl Chloroformate Solvolyses. J Chem Soc., Perkin Trans 2 2002, 240–243. [Google Scholar]
  17. Castro, EA; Ruiz, MG; Salinas, S; Santos, JG. Kinetics and Mechanism of the Aminolysis of Phenyl and 4-Nitrophenyl Chloroformates in Aqueous Solution. J. Org. Chem 1999, 64, 4817–4820. [Google Scholar]
  18. Kyong, JB; Won, H; Kevill, DN. Application of the Extended Grunwald-Winstein equation to Solvolyses of n-Propyl Chloroformate. Int. J. Mol. Sci 2005, 6, 87–96. [Google Scholar]
  19. Bentley, TW; Harris, HC; Ryu, ZH; Gui, TL; Dae, DS; Szajda, SR. Mechanisms of Solvolyses of Acid Chlorides and Chloroformates. Chloroacetyl and Phenylacetyl Chloride as Similarity Models. J. Org. Chem 2005, 70, 8963–8970. [Google Scholar]
  20. Ryu, ZH; Lee, YH; Oh, Y. Stoichiometric Effects. Correlation of the Rates of Solvolysis of Isopropenyl Chloroformate. Bull. Korean Chem. Soc 2005, 26, 1761–1766. [Google Scholar]
  21. Kevill, DN; Koyoshi, F; D’Souza, MJ. Correlation of the Specific Rates of Solvolysis of Aromatic Carbamoyl Chlorides, Chloroformates, Chlorothionoformates, and Chlorodithioformates Revisited. Int. J. Mol. Sci 2007, 8, 346–352. [Google Scholar]
  22. D’Souza, MJ; Reed, D; Koyoshi, F; Kevill, DN. Consideration of the Factors Influencing the Specific Rates of Solvolysis of p-Methoxyphenyl Chloroformate. Int. J. Mol. Sci 2007, 8, 788–796. [Google Scholar]
  23. D’Souza, MJ; Shuman, KE; Carter, SE; Kevill, DN. Extended Grunwald-Winstein Analysis–LFER Used to Gauge Solvent Effects in p-Nitrophenyl Chloroformate Solvolysis. Int. J. Mol. Sci 2008, 9, 2231–2242. [Google Scholar]
  24. D’Souza, MJ; Reed, DN; Erdman, KJ; Kevill, DN. Grunwald-Winstein Analysis–Isopropyl Chloroformate Solvolysis Revisited. Int. J. Mol. Sci 2009, 10, 862–879. [Google Scholar]
  25. Kevill, DN; D’Souza, MJ. Correlation of the Rates of Solvolysis of Phenyl Chlorothionoformate and Phenyl Chlorodithioformate. Can. J. Chem 1999, 77, 1118–1122. [Google Scholar]
  26. Kevill, DN; D’Souza, MJ. Dual Pathways in the Solvolyses of Phenyl Chlorothioformate. J. Org. Chem 1997, 62, 7869–7871. [Google Scholar]
  27. Koo, IS; Yang, K; Kang, DH; Park, HJ; Kang, K; Lee, I. Transition-State Variation in the Solvolyses of Phenyl Chlorothionoformate in Alcohol-Water Mixtures. Bull. Korean Chem. Soc 1999, 20, 577–580. [Google Scholar]
  28. An, SK; Yang, JS; Cho, JM; Yang, K; Lee, PL; Bentley, TW; Lee, I; Koo, IS. Correlation of the Rates of Solvolysis of Phenyl Chlorodithioformate. Bull. Korean Chem. Soc 2002, 23, 1445–1450. [Google Scholar]
  29. Bentley, TW. Structural Effects on the Solvolytic Reactivity of Carboxylic and Sulfonic Acid Chlorides. Comparisons with Gas-Phase Data for Cation Formation. J. Org. Chem 2008, 73, 6251–6257. [Google Scholar]
  30. D’Souza, MJ; Hailey, SM; Kevill, DN. Use of Empirical Correlations to Determine Solvent Effects in the Solvolysis of S-Methyl Chlorothioformate. Int. J. Mol. Sci 2010, 11, 2253–2266. [Google Scholar]
  31. Polshettiwar, V; Kaushik, MP. Recent Advances in Thionating Agents for the Synthesis of Organosulfur Compounds. J. Sulf. Chem 2006, 27, 353–386. [Google Scholar]
  32. Wakamori, S; Yoshida, Y; Tsuchidate, T; Ishii, Y. Studies on Sulfur Containing Carbonates and Their Pesticidal Activities. Agr. Biol. Chem 1969, 33, 1691–1699. [Google Scholar]
  33. Petit-Dominguez, MD; Martinez-Maganto, J. MCF Fast Derivatization Procedure for the Identification of Resinous Deposit Components from the Inner Walls of Roman Age Amphorae by GC-MS. Talanta 2000, 51, 727–734. [Google Scholar]
  34. Bentley, TW; Llewellyn, G. Yx Scales of Solvent Ionizing Power. Prog. Phys. Org. Chem 1990, 17, 121–158. [Google Scholar]
  35. Kevill, DN; D’Souza, MJ. Additional YCl Values and Correlation of the Specific Rates of Solvolysis of tert-Butyl Chloride in Terms of NT and YCl Scales. J Chem Res Synop 1993, 174–175. [Google Scholar]
  36. Kevill, DN; Ryu, ZH. Additional Solvent Ionizing Power Values for Binary Water-1,1,1,3,3,3,Hexafluoro-2-propanol Solvents. Int. J. Mol. Sci 2006, 7, 451–455. [Google Scholar]
  37. Kevill, DN; Anderson, SW. An Improved Scale of Solvent Nucleophilicity Based on the Solvolysis of the S-Methyldibenzothiophenium Ion. J. Org. Chem 1991, 56, 1845–1850. [Google Scholar]
  38. Kevill, DN. Development and Uses of Scales of Solvent Nucleophilicity. In Advances in Quantitative Structure-Property Relationships; Charton, M, Ed.; JAI Press: Greenwich, CT, USA, 1996; Volume 1, pp. 81–115. [Google Scholar]
  39. Bentley, TW; Koo, IS; Norman, SJ. Distinguishing Between Solvation Effects and Mechanistic Changes. Effects Due to Differences in Solvation of Aromatic Rings and Alkyl Groups. J. Org. Chem 1991, 56, 1604–1609. [Google Scholar]
  40. Liu, KT; Sheu, HC. Solvolysis of 2-Aryl-2-Chloroadamantanes. A New Y scale for Benzylic Chlorides. J. Org. Chem 1991, 56, 3021–3025. [Google Scholar]
  41. Fujio, M; Saeki, Y; Nakamoto, K; Yatsugi, K; Goto, N; Kim, SH; Tsuji, Y; Rappoport, Z; Tsuno, Y. Solvent Effects on Anchimerically Assisted Solvolyses. II. Solvent Effects in Solvolyses of threo-2-Aryl-1-methylpropyl-p-toluenesulfonates. Bull. Chem. Soc. Jpn 1995, 68, 2603–2617. [Google Scholar]
  42. Kevill, DN; Ismail, NHJ; D’Souza, MJ. Solvolysis of the (p-Methoxybenzyl)dimethylsulfonium Ion. Development and Use of a Scale to Correct for Dispersion in Grunwald-Winstein Plots. J. Org. Chem 1994, 59, 6303–6312. [Google Scholar]
  43. D’Souza, MJ; Darrington, AM; Kevill, DN. On the Importance of the Aromatic Ring Parameter in Studies of the Solvolyses of Cinnamyl and Cinnamoyl Halides. Org. Chem. Int 2010, 2, 13050621–13050629. [Google Scholar]
  44. Kumar, DS; Jayakumar, KP; Balachandran, S. The Leaving Group Dependence in the Rates of Solvolysis of 1,2-Phenylethyl System. J Phys Org Chem 2010. doi:10.1002/poc.1662.. [Google Scholar]
  45. Kevill, DN; D’Souza, MJ. Use of the Simple and Extended Grunwald-Winstein Equations in the Correlation of the Rates of Solvolysis of Highly Hindered Tertiary Alkyl Derivatives. Cur. Org. Chem 2010, 14, 1037–1049. [Google Scholar]
  46. Lee, I. Nucleophilic Substitution at a Carbonyl Carbon. Part II. CNDO/2 Studies on Conformation and Reactivity of the Thio-Analogues of Methyl Chloroformate. J. Korean Chem. Soc 1972, 16, 334–340. [Google Scholar]
  47. Kevill, DN. Chloroformate Esters and Related Compounds. In The Chemistry of the Functional Groups: The Chemistry of Acyl Halides; Patai, S, Ed.; Wiley: New York, NY, USA, 1972; ; Chapter 12, pp. 381–453. [Google Scholar]
  48. Silvia, CJ; True, NS; Bohn, RK. Low Resolution Microwave Spectroscopy. 13. Conformations of S-n-Propyl Thioesters. J. Phys. Chem 1978, 82, 483–488. [Google Scholar]
  49. Shen, Q; Krisak, R; Hagen, K. The Molecular Structure of Methyl Chlorothioformate by Gas-Phase Electron Diffraction and Microwave Spectroscopy Data. J. Mol. Struc 1995, 346, 13–19. [Google Scholar]
  50. Gobbato, KI; Della Védova, CO; Mack, H-G; Oberhammer, H. Structures and Conformations of (Trifluoromethyl)thioacetic Acid, CF3C(O)SH, and Derivatives CF3C(O)SCH3 and CF3C(O)SCl. Inorg. Chem 1996, 35, 6152–6157. [Google Scholar]
  51. So, SP. Structures, Relative Stabilities and Barriers to Internal Rotation of Chloroformyl Hypochlorite and Thiohypochlorite. J Mol Struc Theochem 1998, 168, 217–225. [Google Scholar]
  52. Ulic, SE; Coyanis, EM; Romano, RM; Della Védova, CO. S-Ethyl Thiochloroformate, ClC(O)CH2CH3: Unusual Conformational Properties? Spectrochimica Acta Part A: Mol. Biomol. Spec 1998, 54, 695–705. [Google Scholar]
  53. Romano, RM; Della Védova, CO; Downs, AJ; Parsons, S; Smith, S. Structural and Vibrational Properties of ClC(O)SY Compounds with Y = Cl and CH3. New J. Chem 2003, 27, 514–519. [Google Scholar]
  54. Erben, MF; Della Védova, CO; Boese, R; Willner, H; Oberhammer, H. Trifluoromethyl Chloroformate: ClC(O)OCF3: Structure, Conformation, and Vibrational Analysis Studied by Experimental and Theoretical Methods. J. Phys. Chem. A 2004, 108, 699–706. [Google Scholar]
  55. Silvia, CJ; True, NS; Bohn, RK. Low Resolution Microwave Spectroscopy: Part 14. Conformations of S-isopropyl Thioesters and a High Resolution Study of S-isopropyl Thiofluoroformate. J. Mol. Struct 1979, 51, 163–170. [Google Scholar]
  56. Kevill, DN; Bond, MW; D’Souza, MJ. Dual Pathways in the Solvolyses of Phenyl Chlorothioformate. J. Org. Chem 1997, 62, 7869–7871. [Google Scholar]
  57. Crunden, EW; Hudson, RF. The Mechanism of Hydrolysis of Acid Chlorides. Part VII. Alkyl Chloroformates. J Chem Soc 1961, 3748–3755. [Google Scholar]
  58. Queen, A. Kinetics of the Hydrolysis of Acyl Chlorides in Pure Water. Can. J. Chem 1967, 45, 1619–1629. [Google Scholar]
  59. Queen, A; Nour, TA; Paddon-Row, MN; Preston, K. Kinetics and Hydrolysis of Thiochloroformate Esters in Pure Water. Can. J. Chem 1970, 48, 522–527. [Google Scholar]
  60. McKinnon, DM; Queen, A. Kinetics and Mechanism for the Hydrolysis of Chlorothionoformate and Chlorodithioformate Esters in Water and Aqueous Acetone. Can. J. Chem 1972, 50, 1401–1406. [Google Scholar]
  61. La, S; Koh, KS; Lee, I. Nucleophilic Substitution at a Carbonyl Carbon Atom (XI). Solvolysis of Methyl Chloroformate and Its Thioanalogues in Methanol, Ethanol and Ethanol-Water Mixtures. J. Korean Chem. Soc 1980, 24, 1–7. [Google Scholar]
  62. La, S; Koh, KS; Lee, I. Nucleophilic Substitution at a Carbonyl Carbon Atom (XII). Solvolysis of Methyl Chloroformate and Its Thioanalogues in CH3CN-H2O and CH3COCH3-H2O Mixtures. J. Korean Chem. Soc 1980, 24, 8–14. [Google Scholar]
  63. Castro, EA. Kinetics and Mechanisms of Reactions of Thiol, Thiono, and Dithio Analogues of Carboxylic Esters with Nucleophilies. Chem. Rev 1999, 99, 3505–3524. [Google Scholar]
  64. Castro, EA; Cubillos, M; Santos, JG. Kinetics and Mechanisms of the Pyridinolysis of Phenyl and 4-Nitrophenyl Chlorothionoformates. Formation and Hydrolysis of 1-(Aryloxythiocarbonyl)pyridinum Cations. J. Org. Chem 2004, 69, 4802–4807. [Google Scholar]
  65. Castro, EA; Aliaga, M; Gazitúa, M; Santos, JG. Kinetics and Mechanism of the Reactions of S-Methyl Chlorothioformate with Pyridines and Secondary Alicyclic Amines. Tetrahedron 2006, 62, 4863–4869. [Google Scholar]
  66. Castro, EA. Kinetics and Mechanisms of Reactions of Thiol, Thiono, and Dithio Analogues of Carboxylic Esters with Nucleophilies. An Update. J. Sulf. Chem 2007, 28, 401–429. [Google Scholar]
  67. Castro, EA; Aliaga, M; Campodonico, PR; Leis, JR; García-Río, L; Santos, JG. Reactions of Aryl Chlorothionoformates with Quinuclidines. A Kinetic Study. J. Phys. Org. Chem 2008, 21, 102–107. [Google Scholar]
  68. Castro, EA; Gazitúa, M; Santos, JG. Kinetics and Mechanism of the Reactions of Aryl Chlorodithioformates with Pyridines and Secondary Alicyclic Amines. J. Phys. Org. Chem 2009, 22, 1030–1037. [Google Scholar]
  69. D’Souza, MJ; Ryu, ZH; Park, B-C; Kevill, DN. Correlation of the Rates of Solvolysis of Acetyl Chloride and α-Substituted Derivatives. Can. J. Chem 2008, 86, 359–367. [Google Scholar]
  70. Frost, AA; Pearson, RG. Kinetics and Mechanism-a Study of Homogeneous Chemical Reactions, 2nd ed; Wiley: New York, NY, USA, 1961; pp. 49–50. [Google Scholar]
  71. Kevill, DN; Abduljaber, MH. Correlation of the Rates of Solvolysis of Cyclopropylcarbinyl and Cyclobutyl Bromides Using the Extended Grunwald-Winstein Equation. J. Org. Chem 2000, 65, 2548–2554. [Google Scholar]
Figure 1. Molecular structures of s-isopropyl chlorothioformate (1), s-isopropyl chloroformate (2), and s-phenyl chloroformate (3).
Figure 1. Molecular structures of s-isopropyl chlorothioformate (1), s-isopropyl chloroformate (2), and s-phenyl chloroformate (3).
Ijms 11 02597f1
Figure 2. The plot of log (k/ko) for isopropyl chlorothioformate (1) against 0.38 NT + 0.72 YCl in the twenty common pure and binary solvents. The point for the 100% EtOH was not included in the correlation. It is added to show the extent of its deviation.
Figure 2. The plot of log (k/ko) for isopropyl chlorothioformate (1) against 0.38 NT + 0.72 YCl in the twenty common pure and binary solvents. The point for the 100% EtOH was not included in the correlation. It is added to show the extent of its deviation.
Ijms 11 02597f2
Scheme 1. Stepwise addition-elimination mechanism through a tetrahedral intermediate for phenyl chloroformate (3).
Scheme 1. Stepwise addition-elimination mechanism through a tetrahedral intermediate for phenyl chloroformate (3).
Ijms 11 02597f3
Scheme 2. Solvolysis-decomposition of isopropyl chloroformate (2).
Scheme 2. Solvolysis-decomposition of isopropyl chloroformate (2).
Ijms 11 02597f4
Scheme 3. Unimolecular step-wise hydrolysis of isopropyl chlorothioformate (1).
Scheme 3. Unimolecular step-wise hydrolysis of isopropyl chlorothioformate (1).
Ijms 11 02597f5
Table 1. Specific rates of solvolysis (k) of 1, in several binary solvents at 25.0 °C, literature values for NT and YCl, specific rate ratios for 1 and 2, and the corresponding ratios for the ethyl and phenyl compounds.
Table 1. Specific rates of solvolysis (k) of 1, in several binary solvents at 25.0 °C, literature values for NT and YCl, specific rate ratios for 1 and 2, and the corresponding ratios for the ethyl and phenyl compounds.
Solvent (%)a1 @ 25.0 °C; 105k, s−1bNTcYCldk1/k2eEtfPhg
100% MeOH1.99 ± 0.11h0.17−1.20.470.260.042
90% MeOH5.06 ± 0.24−0.01−0.200.61
80% MeOH24.7 ± 0.3−0.060.671.840.40
100% EtOH1.21 ± 0.06i0.37−2.501.110.190.026
90% EtOH3.32 ± 0.170.16−0.901.410.210.029
80% EtOH13.7 ± 0.7j0.000.003.490.370.029
70% EtOH42.1 ± 1.3−0.200.807.610.029
90% Acetone0.153 ± 0.004k−0.35−2.390.46
70% Acetone1.21 ± 0.06−0.420.170.47
97% TFE (w/w)49.8 ± 2.5l−3.302.834.052605.9
90% TFE (w/w)69.5 ± 2.1−2.552.855.001710.50
70% TFE (w/w)212 ± 18m−1.982.9610.7689
80T-20E14.5 ± 0.8−1.761.894.450.10
60T-40E3.75 ± 0.18−0.940.632.660.033
50T-50E2.81 ± 0.02−0.640.60
40T-60E1.55 ± 0.16−0.34−0.481.61
20T-80E1.09 ± 0.130.08−1.421.44
97%HFIP (w/w)376 ± 17−5.265.172.58253
90%HFIP (w/w)437 ± 28n−3.844.416.9135716
70%HFIP (w/w)659 ± 39o−2.943.8310.971830.22p
aSubstrate concentration of ca. 0.0052 M; binary solvents on a volume-volume basis at 25.0 °C, except for TFE-H2O and HFIP-H2O solvents which are on a weight-weight basis. T-E are TFE-ethanol mixtures.
bWith associated standard deviation.
cValues taken from [37,38].
dValues taken from [3436].
eThe k2 values from ref. 24.
fValues for k(EtSCOCl)/k(EtOCOCl) from [10].
gValues for k(PhSCOCl)/k(PhOCOCl) from [26].
hA value of 6.00 (± 0.30) × 10−5 s−1 and a value of 19.3 (± 0.15) × 10−5 s−1 was obtained at 35.0 °C and 45.0 °C respectively. ΔH = 20.8 ± 0.7 kcal/mol, ΔS = −10.4 ± 8 cal mol−1 K−1.
iA value of 2.34 (± 0.19) × 10−5 s−1 and a value of 6.11 (± 0.17) × 10−5 s−1 was obtained at 35.0 °C and 45.0 °C respectively. ΔH = 14.6 ± 1.9 kcal/mol, ΔS = −32.1 ± 6.5 cal mol−1 K−1.
jA value of 21.0 (± 0.11) × 10−5 s−1 and a value of 42.4 (± 0.18) × 10−5 s−1 was obtained at 35.0 °C and 45.0 °C respectively. ΔH = 10.4 ± 1.4 kcal/mol, ΔS = −41.4 ± 4.9 cal mol−1 K−1.
kA value of 0.248 (± 0.020) × 10−5 s−1 and a value of 0.513 (± 0.030) × 10−5 s−1 was obtained at 35.0 °C and 45.0 °C respectively. ΔH = 10.8 ± 1.6 kcal/mol, ΔS = −49.0 ± 5.7 cal mol−1 K−1.
lA value of 14.7 (± 0.1) × 10−5 s−1 and a value of 26.7 (± 0.2) × 10−5 s−1 was obtained at 15.0 °C and 20.0 °C respectively. ΔH = 20.2 ± 0.5 kcal/mol, ΔS = −5.8 ± 2.0 cal mol−1 K−1.
mA value of 81.4 (± 3.6) × 10−5 s−1 and a value of 145 (± 9) × 10−5 s−1 was obtained at 15.0 °C and 20.0 °C respectively. ΔH = 15.8 ± 1.8 kcal/mol, ΔS = −17.8 ± 6.3 cal mol−1 K−1.
nA value of 162 (± 8) × 10−5 s−1 and a value of 244 (± 11) × 10−5 s−1 was obtained at 15.0 °C and 20.0 °C respectively. ΔH = 16.3 ± 1.9 kcal/mol, ΔS = −14.6 ± 6.8 cal mol−1 K−1.
oA value of 212 (± 12) × 10−5 s−1 and a value of 358 (± 19) × 10−5 s−1 was obtained at 15.0 °C and 20.0 °C respectively. ΔH = 18.8 ± 1.0 kcal/mol, ΔS = −5.6 ± 4.2 cal mol−1 K−1.
pFor 50% HFIP (w/w).
Table 2. Correlation of the specific rates of solvolysis of a variety of ROCOCl, RSCOCl, RSCSCl, and ROCSCl substrates using the extended Grunwald-Winstein equation (Equation 2).
Table 2. Correlation of the specific rates of solvolysis of a variety of ROCOCl, RSCOCl, RSCSCl, and ROCSCl substrates using the extended Grunwald-Winstein equation (Equation 2).
SubstratenalbmbccRdFeMechanism
PhOCOClf491.66 ± 0.050.56 ± 0.030.15 ± 0.070.980568A-E
PhSCSClg310.69 ± 0.050.95 ± 0.030.18 ± 0.050.987521SN1
PhOCSClh91.88 ± 0.280.56 ± 0.150.38 ± 0.150.95028A-E
180.34 ± 0.050.93 ± 0.09−2.54 ± 0.340.95577SN1
PhSCOCli161.74 ± 0.170.48 ± 0.070.19 ± 0.230.94655A-E
60.62 ± 0.080.92 ± 0.11−2.29 ± 0.130.98344SN1
EtOCOClj281.56 ± 0.090.55 ± 0.030.19 ± 0.240.967179A-E
70.69 ± 0.130.82 ± 0.16−2.40 ± 0.270.94617SN1
EtSCOClk190.66 ± 0.080.93 ± 0.07−0.16 ± 0.310.96196SN1
MeOCOCll191.59 ± 0.090.58 ± 0.050.16 ± 0.070.977A-E
MeSCOClm121.48 ± 0.180.44 ± 0.060.08 ± 0.080.94940A-E
80.79 ± 0.060.85 ± 0.07−0.27 ± 0.180.98795SN1
i-PrOCOCln91.35 ± 0.220.40 ± 0.050.18 ± 0.070.96035A-E
160.28 ± 0.040.59 ± 0.04−0.32 ± 0.060.982176fragmentation-ionization
i-PrSCOClo190.38 ± 0.110.72 ± 0.09−0.28 ± 0.100.96197SN1
an is the number of solvents.
bWith associated standard error.
cThe earlier values are accompanied by standard error of the estimate.
dCorrelation coefficient.
eF-test value.
fValues taken from [21].
gValues taken from [21].
hValues taken from [21].
iValues taken from [21].
jValues taken from [10].
kValues taken from [10].
lValues taken from [12].
mValues taken from [30].
nValues taken from [24].
oThis work without 100% EtOH.

Share and Cite

MDPI and ACS Style

D’Souza, M.J.; Mahon, B.P.; Kevill, D.N. Analysis of the Nucleophilic Solvation Effects in Isopropyl Chlorothioformate Solvolysis. Int. J. Mol. Sci. 2010, 11, 2597-2611. https://doi.org/10.3390/ijms11072597

AMA Style

D’Souza MJ, Mahon BP, Kevill DN. Analysis of the Nucleophilic Solvation Effects in Isopropyl Chlorothioformate Solvolysis. International Journal of Molecular Sciences. 2010; 11(7):2597-2611. https://doi.org/10.3390/ijms11072597

Chicago/Turabian Style

D’Souza, Malcolm J., Brian P. Mahon, and Dennis N. Kevill. 2010. "Analysis of the Nucleophilic Solvation Effects in Isopropyl Chlorothioformate Solvolysis" International Journal of Molecular Sciences 11, no. 7: 2597-2611. https://doi.org/10.3390/ijms11072597

Article Metrics

Back to TopTop