• 沒有找到結果。

Isomerization reactions of RSNO (R=H, C n H2 n +1 n ? 4)

N/A
N/A
Protected

Academic year: 2021

Share "Isomerization reactions of RSNO (R=H, C n H2 n +1 n ? 4)"

Copied!
8
0
0

加載中.... (立即查看全文)

全文

(1)

DOI 10.1007/s00214-006-0164-7

R E G U L A R A RT I C L E

Isomerization reactions of RSNO (R

=H, C

n

H

2n

+1

n

≤ 4)

Chin-Hung Lai · Elise Y. Li · Pi-Tai Chou

Received: 21 March 2006 / Accepted: 2 June 2006 / Published online: 23 August 2006 © Springer-Verlag 2006

Abstract We have applied various theoretical meth-ods to gain detailed insights into the isomers as well as the transition states (TSs) along the corresponding reac-tion pathways for RSNO (R=H, CnH2n+1n≤ 4). On the

basis of G2 and G2MP2 results, the relative order of stability for R=H is estimated to be trans-HSNO > cis-HSNO > HNSO > cis-HONS≈ trans-HONS, while it is cis-CH3SNO ≈ trans-CH3SNO > CH3NSO >

trans-CH3ONS > cis-CH3ONS for R=CH3. A similar trend

is also obtained from the B3P86 method with consid-erably less computing effort if the nearly isoenergetic isomers cis-HONS and trans-HONS are ignored. Based on the results of B3P86, cis-RSNO is more stable than trans-RSNO when R=H is replaced by alkyl groups ex-cept for R=t-Bu. Natural bond orbital analyses allow us to explore whether the high reactivity of S-nitros-othiols is due to the strong negative hyperconjugation (nπO ↔ σ ∗N−S). The mesomeric effect of

S-nitrosothi-ols, although non-negligible, does not cause the break-age of N–O bond due to the compensation of columbic attraction between N and O.

Keywords RSNO· G2 · B3P86 · Negative

hyperconjugation· Mesomeric effect

Electronic supplementary material Supplementary material is

available to authorised users in the online version of this article at http://dx.doi.org/10.1007/s00214-006-0164-7 .

C.-H. Lai (

B

)· E. Y. Li · P.-T. Chou Department of Chemistry,

National Taiwan University, 106, Taipei, Taiwan, R.O.C. e-mail: chinhunglai@ntu.edu.tw P. -T. Chou

e-mail: chop@ntu.edu.tw

1 Introduction

S-nitrosothiols (RSNOs) have attracted much attention because species containing an -SNO functional group have been found in vivo as part of the metabolism of nitric oxide (NO), [1–3] an important biological messen-ger. For a long time, the S-nitrosothiols have been pro-posed to play a key role in transporting and storing NO within the organism [1–6]. The S-nitrosothiols also ex-hibit many biological properties similar to those of NO, including vasodilatation of arteries, inhibition of plate-let aggregation, smooth muscle cell proliferation, etc. [5, 7–11]. Furthermore, the derivatives of S-nitrosothi-ols, with a form of S-nitrosate cysteine thiS-nitrosothi-ols, were found to be involved in Zn2+ complexation to disrupt cys-teine–Zn2+ linkages via a transnitrosation mechanism [12–17]. Unfortunately, the S-nitrosothiols (RSNOs) readily release NO•via thermal agitation [18], radiation [19–21], or reactions catalyzed by certain metal ions [22– 33], superoxide [34, 35] and seleno compounds [36]. Al-though the heterolytic cleavage might as well take place, the biological activity of RSNOs has been mostly attrib-uted to the homolytic cleavage of the S–NO bond with the release of NO•[37]. Accordingly, experimental ap-proaches on the RSNOs are rare [38, 39], and our current understanding on the physical and chemical properties of this important class of biomolecules is still unsatisfac-tory.

Alternatively, computational chemistry seems to of-fer a reasonable access to investigate the reaction path-ways of RSNOs. Houk and others have performed some calculations to probe the bond dissociation energy of the N–S bond as well as the energetics of trans–cis con-formers [40–43]. In a study of primarily photolyzing cis-HSNO, forming the unexpected trans-HONS,

(2)

Nonella et al. have performed an HF calculation in or-der to extract some possible isomers that can correlate with the experimental results [44]. In another approach, Javakumar and Kolandaivel [45] have applied the max-imum hardness principle in an attempt to explain the relative order of stability for various isomers, including HSNO, HONS, HNOS and HOSN. In this paper, we re-port a comprehensive computational approach on the isomers as well as the corresponding transition states (TSs) along the reaction pathways for RSNOs (R=H, CnH2n+1 n≤ 4). On the optimization of theoretical methods, details on structures and thermodynamics of various isomers and TSs are resolved. To the best of our knowledge, this study appears to be the first approach to discuss the corresponding R-shifted isomers regarding S-nitrosothiols.

2 Theoretical methodology

All calculations are done with the Gaussian 03 program [46]. MP2 (full) and various DFT methods are used, including LSDA (SVWN), GGA (BLYP, BP86 and BPW91), and hybrid DFT (B3LYP, B3P86 and B3PW91) methods with the basis set 6-31++G**, except for tran-sition metal [47–53]. For trantran-sition metal (Cu+), we use the relativistic effective core potential (RECP) LANL2 plus DZ basis set, i.e. LANL2DZ [54]. The calculated minima and transition states (TSs) have been carefully checked by frequency analysis to examine whether the number of the imaginary frequency is zero or one. To gain more accurate relative energies among various isomers, G2 and G2MP2 calculations were further performed [55]. All mentioned energetic values are cor-rected for zero-point vibrational energy (ZPVE). Furthermore, NBO 6.0, implemented in Gaussian 03, is used to perform the natural bond orbital (NBO) analysis [56, 57].

3 Results and discussion

We would first like to discuss the optimization of theoret-ical methods in performing RSNO calculations. Tables S1 and S2 summarize the relative energies and isomerization barriers for cis-HSNO and cis-MeSNO on various theoretical methods. Note that the optimized structures are similar among different methods applied. If it is assumed that the approaches on higher levels, such as G2 and G2MP2, are more accurate, so that the results can be used as a criterion, the results of the DFT/B3P86 method seem to be better than that of the computationally more expensive method such as MP2 (full) (see Tables S1 and S2). For example, if both

results of DFT/B3P86 and MP2 (full) refer to that of the G2 method, the MP2 (full) overemphasizes the stability of RNSO by 5.2 kcal/mol for HNSO and 4.6 kcal/mol for CH3NSO.

On the above basis, the results and discussion here-after will mainly rely on B3P86 with the assistance of the NBO analysis. The energetic values and structures of each isomer and TS calculated on B3P86/6-31++G** are listed in Fig. 1 and Fig. S1. Comparing our B3P86/6-31++G** results with respect to those reported by Gauld et al. [21] based on B3P86/6-31++G(3df, 3pd) (see Table S3), it is found that the calculated bond length of N–S (r (N–S)) and N–O (r (N–O)) varies only by 0.02 and 0.01 Å, respectively, while the homolytic bond dis-sociation energy (BDE) on average differ by as little as 1 kcal/mol for both trans-HSNO and cis-MeSNO. It is thus not necessary to use 6-311+G(2df,p) or an even larger basis set, to obtain reliable structures or other properties for the cases of RSNO with R groups at least up to C4H9.

Thionitrous acid (HSNO) can be viewed as the simplest S-nitrosothiol, the results of which may also provide supplementary information for its alkyl ana-logues such as CH3SNO. The calculated results of HSNO,

along with those obtained from previous experimental and theoretical works, are collected in Table 1. Based on the G2 results, the relative energies with respect to cis-HSNO are -1.1, 4.4, 4.4 and 2.7 kcal/mol for trans-cis-HSNO, cis-HONS, trans-HONS, and HNSO, respectively. In comparison, a similar trend could also be found in pre-vious reports [21, 22, 24, 31, 32], except for Bharatam’s work [32], who predicted an opposite stability order be-tween cis- and trans-HSNO. Applying the Hartree-Fock method with various basis sets, Nonella and Huber [44] predicted the relative stability to be on the order of HNSO > HOSN > HSNO > HONS, while a relative en-ergy order of HSNO > HONS > HNOS > HOSN was estimated by Kolandaivel et al. [45] based on the HF/6-31G method. In addition, trans-HSNO was more stable than cis-HSNO by 0.6 kcal/mol. Although their results vary, depending on the basis sets applied, the predic-tion that HSNO is more stable than HONS is in mutual agreement and is also consistent with this study (see Table 1).

Table 2 summarizes the calculated results on CH3SNO.

For a fair comparison, data obtained from previous re-sults are also collected in Table 2. Based on the G2 results, the relative energies with respect to cis-MeSNO are 0.6, 14.4, 13.7 and 4.1 kcal/mol for trans-MeSNO, cis-MeONS, trans-MeONS and MeNSO, respectively. This tendency is qualitatively in agreement with previ-ous reports based on B3LYP and CBS-4M [41–43, 60]. Comparing Tables 1 and 2, it can be promptly perceived

(3)

Fig. 1 Structures and relative

energetics of various HSNO isomers and the

corresponding transition states (H in white, N in ashy, O in dark ashy and S in black

colors) calculated by

B3P86/6-31++G** (energies in kcal/mol, bond lengths in Å, and angles in◦)

that both relative energies of isomers and the asso-ciated reaction barriers increase as R=H is replaced by a methyl group, despite the relative stability orders

being slightly different in these two cases. For R=H, the order of stability is calculated to be trans-HSNO > cis-HSNO > HNSO > cis-HONS≈ trans-HONS, while

(4)

Table 1 The selected geometric parameters, Wiberg-type bond

orders and relative energies of isomers and transition states of HSNO calculated by B3P86/6-31++G** (energies in kcal/mol, bond lengths in Å)

r r BO BO Relative

(N–S) (N–O) (N–S) (N–O) energy

cis-HSNO 1.84 1.18 1.07 1.93 0.0 1.75a 1.17a trans-HSNO 1.86 1.18 1.06 1.94 −0.9 1.76a 1.17a −0.3a 1.91e 1.18e −0.6b −1.1c 0.9f cis-HONS 1.59 1.34 1.75 1.17 6.9 1.56a 1.31a 5.3a 4.8b 7.0d trans-HONS 1.58 1.37 1.80 1.12 6.9 1.55a 1.33a 5.1a 4.8b HNSO 1.63 1.23 1.37 1.45 2.6 TS1 2.00 1.16 9.8 1.85a 1.16a 10.4f TS2 1.71 1.21 47.3 TS3 1.65 1.26 28.7 33.5d TS4 1.57 1.42 20.7 1.54a 1.38a TS5 1.58 1.34 50.8 BO Bond order

aBased upon HF/basis (c), see ref. [44] bBased upon HF/6-31G, see ref. [45] cBased upon CBS-4M, see ref. [41]

dBased upon B3LYP/6-311G**, see ref. [56, 57] eBased upon UB3LYP/6-31G*, see ref. [41] fBased upon B3LYP/6-31+G*, see ref. [58, 59]

it is cis- CH3SNO ≈ trans- CH3SNO > CH3NSO >

trans- CH3ONS > cis- CH3ONS for R=CH3. This

re-sult might be due to the fact that both N–S and N–O bonds possess a double bond character in NOS, whereas the strength ofπN−S is weaker than that of πN−O. As a result, substituents R (R=H, Methyl) prefer to at-tach to the N–S bond, especially on the S center. For R=H, there might exist another important factor to influ-ence the relative order, namely the strength of hydrogen bonding from e.g. O–H. . . .S, O–H. . .N, etc. Our results indicate that the RSNO appears to be the most stable form among its isomers, regardless of the trans or cis form. Table 3 lists the results obtained from B3P86/6-31++G**, so that a fair comparison can be made of the relative energies between cis- and trans-RSNO, in which R=ethyl, n-propyl and iso-propyl, n-butyl, iso-butyl and tert-butyl. In all cases (R=H∼ t-C4H9), r (N–S) in

trans-RSNO is longer than that in cis-trans-RSNO, whereas the r (N–O) in trans-RONS is longer than that in cis-RONS.

Table 2 The selected geometric parameters, Wiberg-type bond

orders and relative energies of isomers and transition states of MeSNO calculated by B3P86/6-31++G** (energies in kcal/mol, bond lengths in Å)

r r BO BO Relative

(N–S) (N–O) (N–S) (N–O) energy

cis-MeSNO 1.82 1.19 1.10 1.86 0.0 1.87b 1.19b 1.86d 1.18d trans-MeSNO 1.82 1.19 1.09 1.89 1.0 1.86d 1.17d 0.5a 1.0c 0.8e cis-MeONS 1.60 1.34 1.73 1.17 17.5 trans-MeONS 1.58 1.35 1.77 1.13 16.9 MeNSO 1.64 1.23 1.30 1.49 5.3 TS1 1.96 1.17 14.7 13.2c TS2 1.71 1.21 67.8 TS3 1.61 1.23 53.4 TS4 1.57 1.41 31.2 TS5 1.60 1.30 76.5

aBased upon CBS-4M, see ref. [41] bBased upon UB3LYP/6-31G*, see ref. [41] cBased upon B3LYP/6-31+G*, see ref. [58, 59] dBased upon B3LYP/6-311+G*, see ref. [42, 43] e Grel in ref. [42,43]

These tendencies could be qualitatively rationalized by the repulsion of lone pair electrons between sulfur and oxygen in trans-RSNO as well as in trans-RONS (see Scheme 1, trans-MeSNO and trans-MeONS were taken as examples). Furthermore, as shown in Table 3, ex-cept for R= t-Bu, the cis form is more stable than the trans form when R=H is replaced by alkyl groups. The anomalous tendency for R=t-Bu could be attributed to the significant steric hindrance between tert-butyl and oxygen atom in the cis form.

As supported by Table 3, the quantity of the relative energy is around 0.6 kcal/mol for primary S-nitr-osothiols (R=Et, n-Pr, n-Bu), and 1.2 kcal/mol for the tertiary one (R=t-Bu), while it is wider for secondary ones, being 0.2 kcal/mol for i-PrSNO and 0.6 kcal/mol for i-BuSNO. Furthermore, the rotational barriers of cis-EtSNO, cis-i-PrSNO and trans-t-BuSNO have been calculated by B3P86/6-31++G**, and the results are summarized in Table 4. Note that these three reactions are all endothermic. Interestingly, independent of the quantity of endothermicity, the rotational barriers are close in these three reactions, being calculated to be 14.4, 13.7, 13.2 kcal/mol for cis-EtSNO, cis-i-PrSNO and trans-t-BuSNO, respectively. Based on the method developed by Wiberg [33] the bond orders for some crit-ical bonds are listed in Table 2. The bond order for N–S

(5)

Scheme 1 Simple illustration of the repulsion in tran-RONS and

trans-RSNO, respectively (H in white, C in light ashy, N in ashy,

O in dark ashy and S in black colors)

is calculated to be 1.09 and 1.10 for trans-MeSNO and cis-MeSNO, respectively, the result of which agrees with the conclusion (single bond character) made by Gauld and co-workers [40]. Furthermore, the calculated N–S bond orders (BO (N–S)) are 1.73 and 1.77 for cis-MeONS and trans-MeONS, respectively. The BO (N–O) is 1.86 and 1.89 for cis-MeSNO and trans-MeSNO, respectively. The results lead us to conclude that both N–S in MeONS and N–O in MeSNO possess double bond characters. These are in consistence with that of the previous re-ports [40, 41]. Note that a similar double bond character is found by HSNO isomers. For example, BO (N–S) is calculated to be 1.80 and 1.75 for trans- and cis-HONS. Likewise, the bond order of N–O (BO (N–O)) for trans-and cis-HSNO is estimated to be 1.94 trans-and 1.93.

As for the transition state (TS), as shown in Fig. 1 and Table 4, r (N–S) is lengthened and r (N–O) is short-ened among cis and trans forms analogues in all TS1, the structures of which are depicted in Fig. S1. The energy difference between cis and trans conformers, i.e. |E(trans)–E(cis)|, was∼0.6 kcal/mol for 1◦ S-nitrosothi-ols and 1.2 kcal/mol for 3◦ S-nitrosothiol (t-BuSNO).

Table 3 The relative energies and bond lengths of cis- and

trans-RSNO for R=Et, n-Pr, i-Pr, n-Bu, i-Bu, t-Bu (energies in kcal/mol, bond lengths in Å) R= r r Relative (N–S) (N–O) energy Et cis 1.81a 1.19a 0.0 1.86b 1.18b trans 1.82a 1.19a 0.5a 1.86b 1.18b 0.5c n-Pr cis 1.81a 1.19a 0.0 trans 1.82a 1.19a 0.6a i-Pr cis 1.81a 1.19a 0.0 trans 1.82a 1.19a 0.2a 0.1c n-Bu cis 1.81a 1.19a 0.0 trans 1.82a 1.19a 0.6a i-Bu cis 1.81a 1.19a 0.0 trans 1.82a 1.19a 0.6a t-Bu cis 1.78a 1.20a 0.0 1.82b 1.19b trans 1.80a 1.19a −1.2a 1.84b 1.18b −1.2c aBased upon B3P86/6-31++G**

bBased upon B3LYP/6-311+G*, see ref. [42, 43] c Grel in ref. [42,43]

Moreover, the rotational barrier is calculated to be 14.4 and 13.9 kcal/mol for cis-EtSNO and trans-EtSNO, respectively. Thus, trans- and cis- forms of primary S-nitrosothiols should interconvert quickly and revers-ibly at ambient temperature. Conversely, the rotational barrier is 13.2 and 12 kcal/mol for trans-t-BuSNO and cis-t-BuSNO, respectively. From both thermodynamic and kinetic viewpoints, 3◦S-nitrosothiols is less prone to undergo trans⇐⇒ cis isomerization reaction. Accord-ing to the above comparison, the relative stability order is 3◦S-nitrosothiols > 1◦S-nitrosothiols. This viewpoint is in agreement with two experimental evidences. Firstly, it has been reported that primary S-nitrosothiols are typ-ically unstable and could be characterized only by tran-sient spectroscopy [62, 63], whereas tertiary ones have been successfully isolated and have indefinite stability [64, 65]. Secondly, 1◦ S-nitrosothiols typically produce shorter-lived biological effects than 3◦[66]. The results of relative energetics based on B3P86/6-31++G** are very close to those obtained by Houk et al. based on the B3LYP/6-311+G* method [42, 43].

In yet another approach, Bharatam has proposed that the high reactivities of S-nitrosothiols are due to a negative hyperconjugation, i.e. nπO ⇐⇒ σ∗N−S (see Scheme 2, cis-MeSNO taken as example) [60, 67], and has accordingly proven this viewpoint by calculat-ing CH3SN=X (X=O, NH, CH2) systematically. Their

(6)

Table 4 Comparison of the

structures and energies of cis-,

trans- and TS of RNSO for

R = Et, i-Pr, t-Bu calculated by B3P86/6-31++G** (energies in kcal/mol, bond lengths in Å and angles in◦)

R r r r

(N–S) (N–O) (S–C) (C–S–N) (S–N–O) (C–S–N–O) Energy

Et cis 1.81 1.19 1.81 102.5 117.1 −0.70 0.0 trans 1.82 1.19 1.82 94.75 116.4 178.68 0.5 TS 1.96 1.17 1.83 94.00 113.2 88.55 14.4 i-Pr cis 1.81 1.19 1.82 103.2 117.5 −0.11 0.0 trans 1.82 1.19 1.83 95.03 116.4 −179.99 0.2 TS 1.95 1.17 1.85 93.42 114.1 83.85 13.7 t-Bu cis 1.78 1.20 1.86 109.1 118.9 −0.19 0.0 trans 1.80 1.19 1.86 97.31 115.7 −179.97 −1.2 TS 1.94 1.17 1.87 98.10 113.8 84.27 12.0

Scheme 2 The illustration of negative hyperconjugation in

S-nitrosothiols (H in white, C in light ashy, N in ashy, O in dark

ashy and S in black colors)

NH and CH2 step by step, the number of lone pair

electrons decrease along with an increase of the BDE of the N–S bond, hence the decrease in r (N–S). By NBO analysis, we also find this orbital interaction in either cis-RSNO and trans-RSNO (R=H, Me). For example, the E(2) values regarding this interaction are 44.0 and 42.8 kcal/mol for trans-MeSNO and cis-MeSNO, respec-tively. On comparing HF and B3P86 [44] as shown in Table 1, it could be promptly found that electron corre-lation is necessary to describe the negative hyperconju-gation interaction. Otherwise, a shorter r (N–S) would be gained. Note that if the bond length correlates well with BDE, a higher BDE for N–S bond is predicted without electron correlation.

From the NBO analysis, it appears that there ex-ists another strong orbital interaction between nπS and π∗N−O, namely the mesomeric effect [69] and E(2) val-ues regarding this interaction are calculated to be 25.5 and 32.3 kcal/mol for cis-HSNO and cis-MeSNO, respec-tively. Bharatam et al. [60] also noted this interaction but without discussing whether this effect facilitate or in-hibit the negative hyperconjugation [67]. To clarify this, we calculate the energetics and N–S bond length of the complexes formed by cis-HSNO and Lewis acid, BH3, at

O and S sites, namely the O complex and S complex (see Table 5). Doctorovich and coworkers synthesized and

Table 5 The comparison of O complex and S complex by BH3or Cu+of cis-HSNO (energies in kcal/mol, bond lengths in Å)

BDEa r (N–S)b Complexation energyc BH3 O complex 38.4 1.70 −8.0 S complex 13.6 1.98 −10.4 N complex 35.4 1.76 −22.0 Cu+ O complex 50.2 1.66 −33.9 S complex 30.9 2.22 −42.2 N complex 42.2 1.721.71d −35.6

aOriginal BDE of N–S = 30.6 kcal/mol bOriginal r (N–S) = 1.84 Å

cComplexation energy = E (O, S complex, or N-complex)- E

(cis-HSNO)- E (BH3or Cu+)

dBased upon B3P86/6-311+G(2df,p), see ref. [31]

calculated a stable Ir–PhSNO complex [70]. According to their reported X-ray crystal structure and theoretical result, they found that the Ir center was coordinated by the N atom. They proposed that this stabilization was probably due to the electrophilicity of the Ir center, as shown by their DFT calculations. Likewise, we also con-sider this condition, namely N complex. As revealed in Table 5, the formation of O and N complex stabilizes cis-HSNO and shortens the N–S bond, whereas the for-mation of S complex lengthens the N–S bond as well as lowers the BDE of the N–S bond. This result can be rationalized by the basicity of N atom and the rival-ing between the negative hyperconjugation effect and the mesomeric effect depicted in Scheme 3. Note that the O-atom and S-atom induce a negative hyperconju-gation (destabilization) and mesomeric (stabilization) effects, respectively, to RSNO. The addition of BH3to

the O-site, in part, removes the negative hyperconju-gation, resulting in the stabilization of the O complex. Conversely, mesomeric effect is ridden by accommo-dating the S-atom lone pair electron (via the Lewis acid BH3) destabilizes the S complex. Note that although the

(7)

Scheme 3 The possible site of RSNO attacked by Lewis acid (H in white, C in light ashy, N in ashy, O in dark ashy and S in black colors)

mesomeric effect obviously weakensπN−O, there exists a columbic attraction between N and O in the form of N+-O−. The mutual compensation does not cause the breakage of N–O bond in S-nitrosothiols. On the con-trary, the electronegativities of N and S are close to each other, so that one would expect their columbic interac-tion to be weaker than that of N and O. Once the N–S bond is weakened, decomposition takes place sponta-neously.

We believe that a similar mechanism is also appli-cable to other Lewis acids. For example, if ions, such as Cu+ or Fe2+, acting as a Lewis acid, attack the S site (see Scheme 3), the net result should cause the destabilization of S-nitrosothiols and consequently accelerate the release of NO. To prove this, we chose Cu+ ion as an instance. The results are also listed in Table 5. According to our results, we find Cu+has sim-ilar effect as BH3. This gives an explanation why some

ions could catalyze the decomposition of S-nitrosothiols on attaching RSNOs [22–33]. However, an intriguing thought then comes up to be what if the Lewis acid attacks the O-site, the net effect should increase the BDE of the N–S bond and consequently inhibit the NO formation.

4 Conclusion

In conclusion, we have carefully investigated the energetic minima and TSs on isomeric potential en-ergy surfaces of S-nitrosothiols (RSNOs). Based on the results of B3P86, cis-RSNO is more stable than trans-RSNO when R = H is replaced by alkyl groups (CnH2n+1

n≤ 4) except for R = t-Bu. Considering relative com-putational resources and accuracy, B3P86 turns out to be the most suitable method to describe the relative thermodynamics as well as the isomerization process among various theoretical methods used. We could thus summarize our results as follows:

1. The energy of TS1 of the studied S-nitrosothiols is independent of the endothermicity of the reaction. 2. The order of stability is concluded to be 3◦> 1◦

S-nitrosothiols from both thermodynamic and kinetic viewpoints.

3. The S-nitrosothiols are possibly unstable due to the negative hyperconjugation (nπO ⇐⇒ σ∗N−S) ef-fect that strongly weakens theσN−Sbond [58, 59, 67]. Accordingly, a plausible mechanism as for the role played by the metal ions in stabilizing or desta-bilizing the RSNOs is proposed and depicted in Scheme 3.

Acknowledgements We are grateful to the National Center for High-Performance Computing of Taiwan for allowing us gener-ous hours of computing time. We also thank the National Science Council for financial support.

References

1. Fujimori K, Nakajima T (2000) Rev Heteratom Chem 22:181 2. Williams DLH (1999) Acc Chem Res 32:869

3. Feldman PL, Griffith OW, Stuehr DJ (1993) Chem Eng News Dec 20:26

4. Lü JM, Wittbrodt JM, Wang K, Wen Z, Schlegel BH, Wang PG, Cheng JP (2001) J Am Chem Soc 123:2903

5. Al-Sa’Doni H, Ferro A (2000) Clin Sci 98:507

6. de Belder AJ, MacAllistr R, Radomski MW, Moncada S, Valence PJ (1994) Cardiovasc Res 28:691

7. Azuma H, Ishikawa M, Sekizaki S (1986) Br J Pharmacol 88:411

8. Furlong B, Henderson AH, Lewis MJ, Smith JA (1987) Br J Pharmacol 90:411

9. Radomski MW, Palmer RMJ, Moncada S (1987) Br J Phar-macol 92:636

10. Radomski MW, Palmer RMJ, Moncada S (1987) Biochem Biophys Res Commun 148:1482

11. Radomski MW, Palmer RMJ, Moncada S (1987) Lancet ii:1057

12. Kröncke KD (2001) FASEB J 15:2503

13. Kröncke KD, Fehsel K, Schmidt T, Zenke FT, Dasting I, Wesener JR, Bettermann H, Breunig KD, Kolb-Bachofen V (1994) Biochem Biophys Res Commun 200:1105

(8)

14. Aravindakumar CT, Ceulemans J, De Ley M (1999) Biochem J 344:253

15. Montoliu C, Monfort P, Carrasco J, Palacios Ò, Capdevila M, Hidalgo J, Felipo V (2000) J Neurochem 75:266

16. Kröncke KD, Haase H, Beyersmann D, Kolb-Bachofen V, Hayer-Hartl MK (2001) NO Biol Chem 5:289

17. Liu SX, Kawai K, Tyurin VA, Tyurina YY, Borisenko GG, Fabisiak JP, Quinn PJ, Pitt BR, Kagan VE (2001) Biochem J 354:397

18. Oae S, Shinhama K (1983) Org Prep Preced Int 15:165 19. Sexton DJ, Muruganandam A, Mckenny DJ, Mutus B (1994)

Photochem Photobiol 59:463

20. Wood PD, Mutus B, Redmond RW (1996) Photochem Pho-tobiol 64:518

21. Singh RJ, Hogg N, Joseph J, Kalyanaraman B (1995) FEBS Lett 360:47

22. Askew SC, Barnett DJ, McAninly J, Williams DLH (1995) J Chem Soc Perkin Trans 2: 741

23. Williams DLH (1996) Transition Met Chem 21:189

24. Dicks AP, Beloso PH, Williams DLH (1997) J Chem Soc Per-kin Trans 2:1430

25. Dicks AP, Swift HR, Williams DLH, Bulter AR, Al-Sa’doni H, Cox BG (1996) J Chem Soc Perkin Trans 2:481

26. Gorren ACF, Schrammel A, Schmidt K, Mayer B (1996) Arch Biochem Biophys 330:219

27. Mcaninly J, Williams DLH, Askew SC, Butler AR, Russell C (1993) J Chem Soc Chem Commun 23:758

28. Williams DLH (1996) Chem Commun 1085

29. Le M, Zhang H, Means GE (1997) Bioorg Med Chem Lett 7:1393

30. Swift HR, Williams DLH (1997) J Chem Soc Perkin Trans 2:1933

31. Baciu C, Cho K-B, Gauld JW (2005) J Phys Chem B 109:1334 32. Toubin C, Yeung DY-H, English AM, Peslherbe GH (2002)

J Am Chem Soc 124:14816

33. Vanin A F, Papin AA, Serezhenkov VA, Koppenol WH (2004) Nitric oxide 10:60

34. Trujillo M, Alvarez MN, Peluffo G, Freeman BA, Radi R (1998) J Biol Chem 273:7828

35. Aleryani S, Milo E, Rose Y, Kostka P (1998) J Biol Chem 273:6041

36. Hou YC, Guo Z M, Li J, Wang PG (1996) Biochem Biophys Res Commun 228:88

37. Bulter AR, Williams DLH (1993) Chem Rev:233

38. Lipton AJ, Johnson MA, Macdonald T, Lieberman MW, Gozal D, Gaston B (2001) Nature 413:171

39. Lipton AS (2001) Nature 413:118

40. Baciu C, Gauld JW (2003) J Phys Chem A 107:9946 41. Fu Y, Mou Y, Lin B-L, Liu L, Guo Q-X (2002) J Phys Chem

A 106:12386

42. Bartberger MD, Houk KN, Powell SC, Mannion JD, Lo KY, Stamler JS, Toone EJ (2000) J Am Chem Soc 122:5889 43. Bartberger MD, Mannion JD, Powell SC, Stamler JS, Houk

KN, Toone EJ (2001) J Am Chem Soc 123:8886

44. Nonella A, Huber JR, Ha T-K (1987) J Phys Chem 91:5203

45. Javakumar N, Kolandaivel P (2000) Int J Quant Chem 76:648 46. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery J A, Vreven Jr T, Ku-din KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Ba-rone V, Mennucci B, Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Klene M, Li X, Knox JE, Hratchian HP, Cross JB, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Ya-zyev O, Austin A J, Cammi R, Pomelli C, Octhterski JW, Ayala PY, Morokuma K, Voth GA, Salvador P, Dannenberg JJ, Zakrzewski VG, Dapprich S, Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, Foresman JB, Ortiz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P, Komaromi I, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen W, Wong MW, Gonzalez C, Pople JA (2004) Gaussian 03, Revision C.02. Gaussian, Inc., Pittsburgh

47. Head-Gordon M, Pople JA, Frisch MJ (1988) Chem Phys Lett 153:503

48. Hohenberg P, Kohn W (1964) Phys Rev 136:B864 49. Vosko SH, Wilk L, Nusair M, (1980) Can J Phys 58:1200 50. Becke AD (1988) Phys Rev 38:3098

51. Lee C, Yang W, Parr RG (1988) Phys Rev B 37:785 52. Perdew JP (1986) Phys Rev B 33:8822

53. Burke K, Perdew JP, Wang Y (1998) In: Dobson JF, Vignale G, Das MP (eds) Electronic density functional theory: recent progress and new directions. Plenum, New York 54. Hay PJ, Wadt WR (1985) J Chem Phys 82:270

55. Curtiss LA, Raghavachari K, Trucks GW, Pople JA (1991) J Chem Phys 94:7221

56. Carpenter JE, Weinhold F (1988) J Mol Struct (Theochem) 169:41

57. Reed AE, Curtiss LA, Weinhold F (1988) Chem Rev 88:899 58. Gutiérrez-Oliva S, Herrera B, Toro-Labbé A, Chermette H

(2005) J Phys Chem A 109:1748

59. Herrera B, Toro-Labbé A (2004) J Phys Chem A 108:1830 60. Bharatam PV, Amita (2002) Tetrahedron Lett 43:8289 61. Wiberg KB (1968) Tetrahedron 24:1083

62. Askew SC, Barnett DJ, McAninly J, Williams DLH (1995) J Chem Soc Perkin Trans 2:741

63. de Oliveira MG, Shishido SM, Seabra AB, Morgon NH (2002) J Phys Chem A 106:8963

64. Field L, Dilits RV, Ravichandran R, Lenhert PG, Carnahana GE (1978) J Chem Soc Chem Commun 249

65. Lenhert PG, Carnahana GE, Ravichandran R (1978) Acta Crystallogr B 34:2645

66. Stamler JS (1995) Curr Top Microbiol Immunol 196:19 67. Reed AE, Schleyer PvR (1990) J Am Chem Soc 112:1434 68. Schleyer PvR, Jiao H (1996) Pure Appl Chem 68:209 69. Murrell JN (1955) Proc Phys Soc A 68:969

70. Perissinotti LL, Estrin DA, Leitus G, Doctorovich F (2006) J Am Chem Soc 128:2512

數據

Fig. 1 Structures and relative energetics of various HSNO isomers and the
Table 1 The selected geometric parameters, Wiberg-type bond orders and relative energies of isomers and transition states of HSNO calculated by B3P86/6-31++G** (energies in kcal/mol, bond lengths in Å)
Table 3 The relative energies and bond lengths of cis- and trans- trans-RSNO for R =Et, n-Pr, i-Pr, n-Bu, i-Bu, t-Bu (energies in kcal/mol, bond lengths in Å) R= r r Relative (N–S) (N–O) energy Et cis 1.81 a 1.19 a 0.0 1.86 b 1.18 b trans 1.82 a 1.19 a 0.5
Table 4 Comparison of the structures and energies of cis-, trans- and TS of RNSO for R = Et, i-Pr, t-Bu calculated by B3P86/6-31++G**

參考文獻

相關文件

And, the value or function of the various theories of consciousness should be judged on the basis of Hua Yen's " complete teaching " in order to ascertain the status

{ As the number of dimensions d increases, the number of points n required to achieve a fair esti- mate of integral would increase dramatically, i.e., proportional to n d.. { Even

Reading Task 6: Genre Structure and Language Features. • Now let’s look at how language features (e.g. sentence patterns) are connected to the structure

we use data (random sample) to test if the data provides significant evidence to reject the null hypothesis.. If X > c reject

A subgroup N which is open in the norm topology by Theorem 3.1.3 is a group of norms N L/K L ∗ of a finite abelian extension L/K.. Then N is open in the norm topology if and only if

The Liouville CFT on C g,n describes the UV region of the gauge theory, and the Seiberg-Witten (Gaiotto) curve C SW is obtained as a ramified double cover of C g,n ... ...

We compare the results of analytical and numerical studies of lattice 2D quantum gravity, where the internal quantum metric is described by random (dynamical)

S15 Expectation value of the total spin-squared operator h ˆ S 2 i for the ground state of cationic n-PP as a function of the chain length, calculated using KS-DFT with various