Organic Electrochemistry S S 2 + S S - 2eCH3CN, NaClO4 ,2ClO4 97% Summer 2003 Historical perspective Organic Electrochemistry is a multidisciplinary science overlapping the vast fields of organic chemistry, biochemistry, physical chemistry and electrochemistry. The first electroorganic synthesis was performed by Michael Faraday (1843). It was the anodic decarboxylation of acetic acid in aqueous medium, and the formation of ethane via creation of a new carboncarbon bond: 2 CH3COO -2ePt CH3CH3 + 2CO2 Henry Kolbe (1849) electrolyzed fatty acids and half-esters of dicarboxylic acids and established the practical basis of electroorganic synthesis. Near the end of the 19th century various electrolytic industrial processes emerged, mostly reductive, stimulated by pioneering works of Kolbe, Haber, Fitcher, Tafel and other notable contemporaries who expanded the foundation of organic electrochemical technology. Great progress in the understanding of fundamental mechanisms of electrochemical reactions has been achieved during the last few decades in parallel with the spectacular advances of electroanalytical and spectroanalytical methodologies and the commercial availability of the relevant instruments. Electrochemical Techniques Previous courses: Pr. D. Thomas, Pr. M. Baker, Pr. J. Lipkowski, Dr. G. Szymanski and I. Burgos) Cyclic Voltammetry: Potential (mV) Reversible System E2 i -E E1 Time(s) Irreversible System workstation CE RE WE i -E Important characteristics: - Peak Potential - Number of electrons exchanged per molecule (n) - Transfer coefficient Bulk electrolysis: - Constant current - Constant potential Reaction Mechanisms, Kinetics and Thermodynamics k1 A + e- . A- k2 B + C k4 E +e- k3 D Initial electrochemical investigation is necessary to set up the electrosynthesis - First step: Nature of Electron Transfer (Example: dissociative ET). RX + e- R + X ? RX R eRH SH R Nu (SRN1) . RNu- R-R Dissociative Electron Transfer (Savéant’s Theory) R + X R-X + e ? R-X In electrochemical studies of organic compounds, a widely investigated process is the Dissociative ET where a chemical bond is broken as a result of a first ET. When the one electron-transfer product is an intermediate, the ET follows a stepwise mechanism and can be described by the Hush –Marcus theory when the initial electron transfer step is the rate determining step.1 On the other hand, when the ET and the bond breaking occur in a concerted manner, Savéant2 proposed a model based on Morse curve picture of bond breaking. This model gives a similar quadratic activation-free energy relationship (eq 1), the difference being the contribution of the bond dissociation energy (BDE) of the fragmented bond to the activation barrier, G ‡0 , which involves only the solvent ( 0 ) and the inner ( i ) reorganization energies for a stepwise mechanism (eq 2). G‡ (the activation free energy), G 0 (the reaction free energy) and G‡0,s and G‡0, c (i.e. the activation energy at zero driving force) represent the intrinsic barriers for a stepwise and a concerted ET respectively. G o ‡ ‡ G G0 1 4G ‡ 0 G ‡0, s i 0 4 2 and (1) G ‡0, c 0 BDER X 4 (2) (1) See for example: (a) Marcus, R. A. Theory and Applications of Electron Transfers at Electrodes and in Solution. In Special Topics in Electrochemistry; Rock, P. A., Ed.; Elsevier: New York, 1977; pp 161-179. (2) (a) Savéant, J-M. J. Am. Chem. Soc. 1987, 109, 6788. (b) Savéant, J-M. Dissociative Electron Transfer. In Advances in Electron Transfer Chemistry; Mariano, P. S., Ed.; JAI Press: New York, 1994; Vol. 4, p. 53-116. How does the driving force affect the ET mechanism? Potential Energy RX.- Stepwise E0 RX / RX . Concerted -E RX + e- . R + X- E0 RX / R. X -E RX + e- Reaction Coordinate Figure: Passage from the stepwise to the concerted mechanism upon decreasing the driving force. Potential energy profiles. E: electrode potential. Concerted vs stepwise ET mechanisms How do know, in practice, whether one or the other mechanism is followed? 1- Experimental detection of the radical anion intermediate: - High scan rate cyclic voltammetry: using ultra microelectrodes, scan rates as high as a few million V/s may be reached, corresponding to lifetimes in the sub-s range. - Homogeneous catalysis allows the extension of the lifetime range up to the ns. Example: O2N SCN O2N Figure: Cyclic Voltammetry in CH3CN/TBAF (0.1M) at a glassy carbon electrode, v = 0.2 V/s, temperature = 20 oC of 3,5-dinitroPhSCN (3) 2 mM at v = 0.2 V/s (c) and v = 2 V/s (d). 2- Transfer coefficient (), which is directly related to the intrinsic barrier (eq 3), is a sensitive probe of the mechanistic nature of the first ET in dissociative processes: 1 G 0 1 G 0 2 4G0‡ G ‡ (3) The transfer coefficient can be determined from the electrochemical peak characteristics (peak width, Ep-Ep/2), or from the slop of the Ep vs log(v) plot. RT Ep 2 F log( v) 1 Ep 29.6 log( v) RT 1.85 1 47.5 F Ep / 2 Ep Ep / 2 Ep 1 at 25o C at 25o C E p is the peak potential and E p / 2 is the half peak potential In a concerted mechanism, a value significantly lower than 0.5 is expected, whereas an value close to or higher than 0.5 is expected in the case of a stepwise mechanism. Example 1: Electrochemical reduction of alkylhalides (concerted ET) Br 250 a Ep-Ep/2 (mV) 200 Br 150 Ep = -2.30V n = 2e- /molecule = 0.3 100 log(v) 50 Ep (V vs Fc) -2 0 2 -2 b -2.2 -2.4 -2.6 log(v) -2.8 -2 Figure: CV of 1,3-dibromoadamantane (c=1.56 mM) in CH3CN + 0.1M NBu4ClO4 on a glassy carbon electrode. v = 0.1 V/s. 0 2 Figure: variation of (a) the peak with and (b) the peak potential with the scan rate. Alkylhalides are known to follow a concerted ET mechanism. Alkyl radicals are generally easier to reduce than their parent halides, the electrochemical reduction is usually a 2 electron process leading to the corresponding anion. R-X + eR + e- R + X R One can Use this chemistry for an efficient electrosynthesis. Example: The electrogenerated anion can be used as an intramolecular nucleophile (the initial molecule containing a good leaving group) for the electrosynthesis of alkenes, alkynes and cyclic compounds. X (CH2)n Y + 2en = 2, n = 3, Product + X- + Yn = 4, n = 5, Ex: Electrochemical reduction of dihaloadamantanes X H + 2eY + Y B A The yield of the intramolecular nucleophilic substitution depends on the nature of the leaving group. X I I Y Br Cl Br F H Br Cl F H Ep (V/SCE) -1.50 -1.66 -1.84 -2.38 -2.23 -2.04 -2.36 -2.80 -2.74 0.31 0.32 0.31 0.30 0.30 0.30 0.31 0.32 0.32 A(%) 3 12 56 100 100 5 30 100 100 B(%) 96 75 44 0 0 94 70 0 0 Example 2: Electrochemical reduction of arylhalides (stepwise ET) Aryl halides are reduced following a stepwise mechanism. Ar-X Ar-X + eAr-X Ar + X Here again one can target the aryl radical and uses it efficiently for the synthesis of valuable aromatic compound. An example is the SRN1 reaction (see later). How Molecular Structure Controls the ET Mechanism? Both thermodynamic and kinetic factors are involved in the competition between concerted and stepwise mechanisms. The passage from the stepwise to the concerted situation is expected to arise when the ion radical cleavage becomes faster and faster. Under these conditions, the rate determining step of the stepwise process tends to become the initial electron transfer. Then thermodynamics will favor one or the other mechanism according to equation 4. G 0 0 0 .. D R X E X . / X E RX / RX . TS . RX e R X RX e R X . (4) Thus, one passes from the stepwise to the concerted mechanism as the driving force for cleaving the ion radical becomes larger and larger. 0 A weak R-X bond, a negative value of ERX / RX and a positive value E X0 / RX of will favor the concerted mechanism and vice versa. All three factors may vary from one molecule (RX) to another. However, there are families of compounds where the passage from one mechanism to the other is mainly driven by one of them. Illustrating examples are given in the Table 1. . . Table 1: Molecular factors governing the dichotomy between concerted and stepwise mechanisms. Transition between ET Mechanisms upon Changing the Driving Force A few experimental systems have shown a transition between concerted and stepwise mechanisms as a function of the driving force which could easily be controlled in electrochemistry by varying the electrode potential. Similar mechanism transitions have been also reported in homogeneous thermal and photochemical electron transfer.3 This behavior demonstrates that the nature of the ET process is dictated by the energetic advantage of one process over another. Analyses of the variation of the transfer coefficient with either the scan rate4 or the potential5 are the main criteria for such a mechanistic transition. In this particular case, besides the conventional voltammetric analysis, the convolution approach is a powerful tool for studying intricate details of electrode processes. Convolution analysis had been initially reported many years ago 6 but has only recently been used rigorously to explore reaction mechanisms and to provide valuable kinetic and thermodynamic data for several systems.5,7 The great advantage with the convolution analysis is that all data points of the voltammetric curve are used in the kinetic analysis and that no assumptions on the ET rate law are made in the analysis of the experimental data. This differs from the conventional voltammetric method where a linear activation-driving force relationship is implicitly assumed.8 (3) J. Am. Chem. Soc. 2000, 122, 5623; 2001, 123, 4886. (4) (a) Andrieux, C. P.; Robert, M.; Saeva, F. D.; Savéant, J-M. J. Am. Chem. Soc. 1994, 116, 7864. (b) Pause, L.; Robert, M.; Savéant, J-M. J. Am. Chem. Soc. 1999, 121, 7158. (5) Antonello, S.; Maran, F. J. Am. Chem. Soc. 1997, 119, 12595. (6) (a) Imbeaux, J. C; Savéant, J-M. J. Electroanal. Chem. 1973, 44, 169. (b) Savéant, J-M.; Tessier, D. J. Electroanal. Chem. 1975, 65, 57. (7) (a) Antonello, S.; Maran, F. J. Am. Chem. Soc. 1999, 121, 9668. (b) Antonello, S.; Musumeci, M.; Wayner, D. D. M.; Maran, F. J. Am. Chem. Soc. 1997, 119, 9541. (c) Donkers, R. L.; Workentin, M. J. Phys. Chem. B 1998, 102, 401. (d) Donkers, R. L.; Maran, F.; Wayner, D. D. M.; Workentin, M. J. Am. Chem. Soc. 1999, 121, 7239. (e) Antonella, S.; Frmaggio, F.; Moretto, A.; Toniolo, C.; Maran, F. J. Am. Chem. Soc. 2001, 123, 9577. (8) (a) Nicholson, R. S.; Shain, I. Anal. Chem. 1964, 36, 706. (b) Nadjo, L.; Savéant, J-M. J. Electroanal. Chem. 1973, 48, 113. (c) Bard, A. J.; Faulkner, L. R. Electrochemical Methods, Fundamentals and Applications; Wiley: New York, 1980. Convolution Analysis The background-subtracted voltammograms are convoluted to yield convoluted current I vs E plots. I is related to the voltammetric current i through the convolution integral (eq 5). I 1 / 2 0 t i (u ) (t u )1 / 2 The limiting current Il is defined as Il = nFAD1/2C0, where n is the overall electron consumption per molecule, A the electrode area, D the diffusion coefficient, and C the bulk substrate concentration. Il is independent of scan rate and is used to calculate D. du (5) For a totally irreversible system (when the dissociative electron transfer is concerted or when the dissociation of the reduction product is fast) I can be related to the rate constant of the heterogeneous electron transfer khet through eq 6. ln k het ln D1/ 2 ln I l I (t ) i(t ) (6) Systems following a single ET mechanism show either a linear or a parabolic pattern. In mixed mechanisms neither a linear nor a parabolic pattern is obtained. Values of ln khet indicate whether the electron transfer is fast or slow. Apparent values of transfer coefficient (app) can be obtained from the ln khet vs E data by using eq 7.Derivatization is accomplished by linear regression of the experimental data within small E intervals (18 to 24 mV). app RT ln k F E (7) The app vs E plot shows that the electrode process is not ruled by a simple ET mechanism. The app vs E plot is characterized by a maximum at ca. –2.14 V, corresponding to an average value of 0.31. *app is related to through the double-layer correction whose properties are yet unknown for the glassy carbon electrode. However, it has been previously showed that uncorrected transfer coefficient values provide a reasonable representation of the process, as these values do not differ much from the true ones. Example 1: Electrochemical reduction of aryl thiocyanates* Convolution analysis was performed for scan rates ranging from 7.2 to 80 V/s for p-methyl (1) and p-methoxy (2) phenyl thiocyanates. Both compounds show a non linear variation of the coefficient transfer () with the driving force. Such nonlinear behavior reflects clearly a change of the electrode mechanism as a function of the potential. ArSCN + e ArS + CN ArSCN It is important to notice that the same maximum was obtained within only 20 mV at any of the scan rates providing data in the appropriate range. This ensures that the observed behavior reflects a real dependence of , beyond experimental error or the effect of artifacts (mainly IR compensation). 0.5 0.4 b app app a 0.5 0.3 0.2 0.1 E (V) -2.4 -2.2 -2 -1.8 0 -1.6 0.4 0.3 0.2 E (V) -2.5 0.1 -2.3 -2.1 -1.9 0 -1.7 Figure: Variation of app with E for (a) 1 (0.85 mM) at scan rate v = 7.2, 10, 20, 30, 40, 60 and 80 V/s; (b) 2 (0.69 mM) at scan rate v = 7.2, 10, 20, 30, 40, 60 and 80 V/s. For p-methyl phenyl thiocyanate (1) the maximum is observed at ca. -2.14 V for 1 corresponding to an average value of 0.31. For p-methoxy phenyl thiocyanate (2) the maximum is observed at ca. -2.08 V for 1 corresponding to an average value of 0.33. *A. Homam, E. M. Hamed and Ian Still J. Am. Chem. Soc. 2003 in press. Example 2: Electrochemical reduction of peroxides* O O O O O + e tBu + O tBu CN CN O O O tBu CN *Antonello, S.; Maran, F. J. Am. Chem. Soc. 1999, 121, 9668. Example 3: Electrochemical reduction of sulfonium salts* S CH3 S + e CH3 CH2 CH2 S CH3 + CH2 * Andrieux, C. P.; Robert, M.; Saeva, F. D.; Savéant, J-M. J. Am. Chem. Soc. 1994, 116, 7864. Electrochemical reduction of a sulfonium cation (see previous Scheme) showing the transition from the concerted to the stepwise mechanism as driving force increases upon raising the scan ate. The apparent transfer coefficient, a, is derived from the peak width according to reported equation (page 8). Does the Concerted ET Mechanism Mean that the Intermediate “Does Not Exist”? The occurrence of a concerted mechanism may result from transition state energy advantage over the stepwise mechanism or from the fact that the intermediate “does not exist”, i.e. its lifetime is shorter that one vibration. Even if examples exist of the latter alternative, there is no reason to view it as an absolute requirement for the mechanism to be concerted. What is true is that it is a sufficient condition for the concerted mechanism to occur. However it is not a necessary condition. In the examples discussed previously, the intermediate “exists”. It transpires along the reaction pathway at high driving forces. Nevertheless, a concerted mechanism is followed at low driving forces because it is then energetically more advantageous. Such competition between stepwise and concerted mechanisms is usually not easy to characterize. This is why the previous examples have a general bearing in the theory of chemical reactivity. Thermodynamics The variation of with the potential as obtained by convolution analysis can be used to estimate the value of the standard reduction potential. 00.10.20.30.40.50.60.7-2.5-2.3-2.1-1.9-1.7-1.5 E0 RX / R . X E Consequently the bond dissociation energy of the cleaved bond can be estimated (BDE). X Ex. Electrochemical reduction of 1,3-dihaloadamantanes X Y DC-X (eV) I I Br Cl Y Br F H Br Cl F H 2.07 2.18 2.20 2.29 2.32 2.70 2.82 2.91 2.94 Global Mechanisms How Molecular Structure Controls the Global Mechanism? An excellent example resides in the electrochemical reduction of benzyl thiocyanates, where electron-donating substituents induce an -cleavage as a result of an ET while electron-withdrawing substituents induce a -cleavage. CH2 Current (A) X S CN 1: X = H 2: X = NO2 -4.E-05 -3.E-05 -2.E-05 -1.E-05 2.E-06 E (V) 0.5 -0.5 -1.5 -2.5 -3.5 Current (A) -2.E-05 0.E+00 1.E-05 E (V vs SCE) 1.3 0.8 0.3 -0.2 -0.7 -1.2 2.E-05 -1.7 Figures : Cyclic Voltammetry in CH3CN/TBAF (0.1M) at a glassy carbon electrode, v = 0.2 V/s, temperature = 20 oC of (a): 1: 2.35 mM (____), 4: 1.3 mM (_ _ _) and (b): 2: 2 mM (____), 5: 1.5 mM (_ _ _). R-X + e- R + X (1) R R (2) R-R + X (3) + e- R + R-X CH2SS CH2 O2N CH2 CH2 NO2 For both compounds, the driving force is in favor of a b-cleavage (mainly due to big difference in the standard oxidation potentials of the 2 leaving groups (NC- and NCS-), whereas the intrinsic barrier is in favor of an a-cleavage, due to a weaker a-bond for both compounds. Example 1: SRN1 reaction -Consist of reducing a convenient substrate in the presence of nucleophile. -The SRN1 reaction is catalytic in electrons. Ar-X + e- Ar-X Ar-X Ar + X Ar + Nu ArNu + Ar-X k ArNu ArNu + Ar-X Arylazo Sulfides in SRN1 Aromatic Nucleophilic Substitutions (electrosynthesis of functionalized aromatic compounds) - The electrochemical reduction of aryl azosulfides yields the corresponding sulfides and decompositions products. - The mechanism of the global reaction is an SRN1 mechanism. NC H NC NN S NC S Figure: Cyclic voltammogram of paracyano phenyl azo phenylsulfide (c=1mM) in ACN + 0.1M NBu4ClO4 on a glassy carbon electrode Electrochemistry of Arylazo Sulfides in the Presence of Cyanide NC CN Figure: Cyclic voltammogram of paracyano phenyl azo phenylsulfide (c=1mM) in ACN + 0.1M NBu4ClO4 on a glassy carbon electrode (a) [CN-] = 0 mM, …; (b) [CN-] = 10 mM, - - -; (c) [CN-] = 320 mM, ____. NC N N S + e-. NC N N S . + N2 + - S NC + CN -. NC CN -. NC S +A +A . A- + NC . A- CN NC S . + A- k= 1,5 1011 M-1s-1 in CAN and 2,4 1011 M-1s-1 in DMSO Important kinetic and stereochemical aspects - Solvent effect (Noyes theory of cage effect reactions). - Higher yields of trapped radicals for trans compounds. a b Figures: Trapping by CN- of 4-cyanophenyl radicals produced by reduction of 1a (a) in CH3CN and (b) in DMSO: (A) fraction trapped by; (B) fraction trapped by the solvent or reduced; (C) sum of A and B. Generality of the reaction: R N N S R' Nu- R Nu + N2 + RS- R = CN, NO2, F, Cl, CH3, CH3O, CF3. R’ = tBu, C6H5 E and Z isomers Good substitution products yields Various nucleophiles can be used including: cyanide, thiolates, phosphides, aryloxides and carbanions. Example 2: Autocatalytic reductive processes A: Electrochemical reduction of aryl thiocyanates SCN X = H; p-CH3; p-CH3O; p-F; p-Cl; p-NO2; m-CF3; o-NO2; 2,4-diNO2; 3,5-diNO2; p-NO2-o-Cl. X ArS - + CN - ArSCN + 2e ArS - + ArSCN k (1) E1 ArSSAr + CN - (2) ArSSAr + 2e 2 ArS - ArSCN + 2e ArS - + CN - (3) E2>E1 Figure: Cyclic Voltammetry in CH3CN/TBAF (0.1M) at a glassy carbon electrode, v = 0.2 V/s, temperature = 20 oC of (a) p-CH3PhSCN (1) 3.26 mM and (b) p-CH3OPhSCN (2) 3.45 mM at v = 2 V/s. Figure: successive CVs of p-CH3PhSCN (3.26 mM), (——) 1st scan and (——) 2nd scan. In CH3CN/TBAF (0.1M) at a glassy carbon electrode; v = 0.2 V/s. While the first scan shows a CV similar to the one described earlier, the second scan shows a much broader reduction peak at a less negative potential and with no trace crossing. The CV of this second scan corresponds to the reduction of 4,4’dimethylphenyl disulfide (3) produced as a result of a nucleophilic substitution of 1 by the aryl thiocyanate (see previous Scheme). Once the disulfide is formed, it is immediately reduced since it is easier to reduce than the corresponding aryl thiocyanate and will thus induce the autocatalytic process. Figure: Cyclic voltammetry, in CH3CN/TBAF (0.1M) at glassy carbon electrode, temperature = 20 oC, of (— —) 1 (2.55 mM) and (——) 1 (2.55 mM) + 4% of 3, at (a) 0.2 V/s, (b) 2.4 V/s and (c) 10 V/s. Adding catalytic amounts of disulfide allows consumption of 1 at the reduction peak of 3 through the catalytic process and only one peak is seen located at the reduction potential of 3. When the scan rate is increased the catalytic process is diminished and two peaks appear (Figure b). At higher scan rates, the catalysis is totally eliminated and one can identify the peaks corresponding respectively to 4 and 1 (Figure c). Interesting features: Variation of the peak shape and position with the scan rate Figure: Cyclic voltammetry of 1: (a1) 0.85 mM, 0.2 V/s; (a2) 0.85 mM, 2.4 V/s; (a3) 0.85 mM, 7.2 V/s; (b1) 5 mM, 0.2 V/s; (b2) 5 mM, 2.4 V/s; (b3) 5 mM, 60 V/s. In CH3CN/TBAF (0.1M) at a glassy carbon electrode. Temperature = 20 oC. At low concentration (0.85 mM), the crossing seen at 0.2 V/s (Figure 4a 1), is readily eliminated by increasing the scan rate to 2.4 V/s (Figure 4a 2) and at 7.2 V/s (Figure 4a3). At higher concentration (5 mM) a higher scan rate is needed to eliminate the crossing and to allow consumption of all the thiocyanate directly at the electrode. At 2.4 V/s (Figure 4b2) trace crossing is still observed and the autocatalytic process is thus efficient. Moreover, one can clearly see the positive shift of the reduction potential of 1 as C0 increases confirming again the existence of an autocatalytic mechanism. The difference in the reduction potential of 1 between 0.5 V/s and 80 V/s is about 700 mV which is too large to be due only to the type of ET mechanism. For a well behaving system reduced following a concerted mechanism with a value of 0.3, the difference would be of the order of 300 mV. This is also a consequence of the autocatalytic process. -2 a b -2.1 Ep (V/Fc) -2.8 -2.5 -2.9 log (v) log (v) -3.2 -1.5 -0.5 0.5 1.5 Ep (V/Fc) -2.4 2.5 -3.3 -1.5 -0.5 0.5 1.5 2.5 Figure: Variation of the reduction peak potential with the log of the scan rate (log (v)) for (a) 1, (3.45 mM), (0.69 mM); (b) 2, (5 mM), (1.25 mM). An important feature is that the variation of the reduction peak potential with the log of the scan rate does not have the same slope over the entire studied scan rate range and three major regions can be discerned in all cases. For the high concentrations one can distinguish the following: (i) at low scan rates, the variation of the peak potential is linear with a relatively low slope (120 and 122 mV per unit log (v) for 1 and 2 respectively). Within this range the crossing exists and the autocatalytic process is very efficient and the reduction peak is consequently pushed towards more positive potentials. The intermediate region shows a much larger slope since it corresponds to the competition between autocatalysis and diffusion (520 and 565 mV per unit log (v) for 1 and 2 respectively): diffusion being favored as v increases. The peak potential shifts dramatically to more negative values and the higher the value of v, the less efficient the autocatalytic process. At higher scan rates (third region), where the crossing is no longer observed, the slope is much smaller (84 and 86 mV) and the autocatalysis is totally eliminated. At lower concentrations, similar behavior is seen, but competition between the autocatalytic process and direct electrochemical consumption (intermediate region) of the aryl thiocyanate at the electrode takes place at lower scan rates. An important factor is the total independence of peak potential of concentration at these high scan rates, for both compounds 1 and 2, indicating clearly that the autocatalytic process is not involved anymore. 0.15 a 0.2 0.15 0.1 Ep-Ep/2 (V) 0.2 0.25 Ep-Ep/2 (V) 0.25 b 0.1 0.05 0.05 log (v) log (v) 0 -1.5 -0.5 0.5 1.5 2.5 0 -1.5 -0.5 0.5 1.5 2.5 Figure: Variation of the reduction peak width with the scan rate v of (a) 1, (3.45 mM), (0.69 mM); (b) 2, (5 mM), (1.25 mM). At high concentrations, the autocatalytic process is very efficient and the peak is very sharp (peak width around 50 mV for both for 1 and 2) at low scan rates. At higher scan rates, corresponding to more competition between diffusion and autocatalysis, the peak width increases dramatically (to higher values than 150 mV for both 1 and 2), before settling on some more or less stable values (around 150 mV) at much higher scan rates where the autocatalysis is eliminated. At lower concentrations, similar phenomenon are seen, with the difference that the competition (increase of the peak width) takes place at lower scan rates. It is important to notice here again that at high scan rates, the peak width is independent of concentration, proving again that these values correspond to the intrinsic characteristics of the direct electrochemical reduction of 1 and 2 and not to secondary chemical reactions, i.e. the autocatalysis process. Organic Electrosynthesis Agrochemicals (fluorinated compounds) Pharmaceuticals (Chiral products) Fine chemicals (free base of cysteine) Organic electrosynthesis Complex natural products (b-hydroxyalkonates) Precursors intermediates Ex.: chiral drugs, market: $ 50 billion USD, annual growth > 25% Example: Reaction of Levoglucosenone in the presence of electrogenerated superoxide anion Possibility to control the stereoselectivity by changing current density or potential. Example: Electrosynthesis of sulfonium and bisulfonium salts* - Radical source (polymer initiators, surfaces modifications,…). - Good starting materials for synthesis (C-glycosides). - Understanding electron transfer to organic molecules and factors controlling it. - The electrosynthesis consists of oxidizing thianthrene in the presence of a convenient “nucleophile” (alkene, alkyne, ketone). S S O S + S Cyclic voltammogram in CH3CN + 0.1 NBu4ClO4, on a glassy carbon electrode, at v = 0.1 V/s of (a) thianthrene (c = 2 mM) and (b) thianthrene (2mM) + acetophenone (c = 320 mM) Upon Adding the ketone, the CV of thianthrene becomes irreversible and the oxidation peak increases (corresponds to 2e-/molecule). * A. Houmam; D. Shukla; D. D. M. Wayner. J. Org. Chem. 1999, 64, 3342 Electrolysis of thianthrene in the presence of ketones R O O ClO4 S S + S - 2e- +H CH3CN, NaClO4 S R = H, 95% R = CH3, 97% . S The chemical method: CH3CN + e- NaClO4 S O - Making a solution of the radical cation by adding a strong acid to thianthrene (explosive). - The reaction is possible only in carbon tetrachloride (toxicity). - The yield <50%. - Longer time. - Difficulties to isolate the compound. - Mechanism cannot be solved. S+ S O . HO HO S S -e- S S ,ClO4 O S + H S Electrosynthesis of bis-sulfonium salts S S 2 S - 2e- + CH3CN, NaClO4 S ,2ClO4 95-100% . S+ S CH3CN NaClO4 S + eS . Th+ . S S S S ,2ClO4 S . Th+ - In the presence of an alkyne, The CV of thianthrene becomes irreversible but does not increase to two electrons per molecule. - Excellent yields are obtained. S X-ray crystal structure of the electrosynthesized cyclohexyl bis-thianthrenium salt. Inability to synthesize and isolate this bissulfonium salt lead chemists to believe it was very unstable. Electrochemistry allowed quantitative synthesis of such salt. Simulation of the CV of thianthrene at different concentration of the nucleophile allows determination of the rate constant of the reaction of thianthrene radical cation with the “nucleophile” Alkene or ketone k (M-1s-1) 10-2 10 OEt 1.6 1.2 Ph OMe 50 Me Ph Ph 5 Ph 90 Ph 2.7 4 4 0.1 O Ph Me 0.9 O p-MePh Me Generality of the process: Electrosynthesis of sulfonium and ammonium Possibility to control the reaction product by monitoring the reaction potential and the concentration of the “nucleophile”. Electrochemical Oxidation and Reduction of Organic Compounds Oxidation of alkanes Alkanes are functionalized by anodic oxidation in acetonitrile, methanol, acetic acid and more acidic solvents such as trifluoroacetic acid and fluorosulphuric acid. Reaction requires very positive electrode potentials and platinum has generally been used as anode in laboratory scale experiments. The first stage of these reactions involves the removal of an electron from either a carbon-hydrogen or a carbon-carbon -bond, with simultaneous bond cleavage to yield the most stable carbon radical and carbonium anion. These are dissociative processes where the radical cation cannot be detected as an intermediate. In acetonitrile, carbon ions combine with the solvent to form a nitrilium ion. The latter reacts with added water to form the Nsubstituted acetamide, often in good yields (introduction of an aminosubstituent. R 1) Pt anode, CH3CN, LiClO4 2) H2O R Et iPr tBu CO2CH3 NHCOCH3 R NHCOCH3 %yield 0 9 62 64 %yield 77 75 7 0 In acetic acid and trifluoroacetic acid, the carbonium ion is quenched by reaction with the carboxylate anion. Electrochemical oxidation in these solvents is a route for the introduction of a hydroxyl substituent. Electrochemical oxidation of haloalkanes Oxidation of alkyl bromides and iodides leads to loss of a nonbonding electron from the halogen substituent, followed by cleavage of the carbon-halogen bond to form a carbonium ion and a halogen atom. The products isolated are formed by further reactions of the carbonium ion while two halogen atoms combine to form the halogen molecule. In acetonitrile, the carbonium ion reacts with a solvent molecule to form a nitrollium ion. The latter is quenched with water to give the N-alkylacetamide. Ex1: CH2I CH3 Pt anode CH3CN, LiClO4 NHCOCH3 Ex2: CH2NHCOCH3 10% CH2I Pt anode CH3CN, LiClO4 + NHCOCH3 90% Pt anode CH3CN, LiClO4 I Electrochemical oxidation of alkenes Alkenes are electrochemically oxidized in nucleophilic solvents such as alcohols, acetic acid or acetonitrile using an undivided cell and the products isolated result from quenching of intermediate carbonium ions by nucleophiles. The radical cation reacts with the nucleophile to form a radical intermediate, which is then oxidized to the carbonium ion, usually at a potential less positive potential. Ex. 1: Electrochemical oxidation of cyclohexene. OMe + MeOH - e- - e- -H + MeOH OMe - e- CH(OMe)2 CHOMe + MeOH Ex. 2: The oxidation potential for the alkene bond is closed to that for a carboxylate ion. In the styrene derivative below, the alkene moiety is preferentially oxidized and intramolecular capture of a carbocation leads to a lactone product. Ph PhCH CHCH2CH2CO2 Pt anode MeOH, NaOMe OMe O O H Ex. 3: Electrochemical oxidation of alkyl substituted butadienes in the presence of dimethylurea as a 1,3-dibentate nucleophile, leads to the formation of a five membered ring heterocycle. O Me + MeNHCONHMe C anode CH3CN, NaClO4 N N Me Electrochemical oxidation of alcohols and esters The direct oxidation of alcohols, 1,3- or higher 1,n-diols is carried out in absence of solvents or in acetonitrile and presence of perchlorates or tetrafluoroborates as supporting electrolytes. Depending on the structures of the alcohols and on the electrolytic systems the electrooxidation of alcohols may lead to the formation of aldehydes, acetals, acids and esters. Ex. 1: Anodic oxidation of 2-butin-1,4-diol to acetylene dicarboxylic acid. OH HO -8e-, H2SO4 - H2O (PbO2) HOOC COOH 70-76% Ex. 2: Anodic oxidation of enolethers and enolacetates. OR R = Me, Ac -2eMeOH OMe OMe OMe -2eMeOH COOMe O 74 - 94 % Oxidation of organic compounds of sulfur and selenium The anodic oxidation of thiols and dialkyl or diaryl sulfides makes it possible to prepare sulfur compounds at different degrees of oxidation in dependence potential, electrode material (carbon, platinum, stainless steel) and composition of electrolyte. (-II) -4e- R SH (II) R SO2O OSO2 R R S OR O -H ,-e(-I) -R ,-e-2eR S R (-II) (IV) -4e- R S S R R SO2O O (0) -2e- R S R (II) R S R O O Ex. 1: thiophenol oxidation to diphenyl disulfide and to esters of benzenesulfonic acid. Ph S OR O 95 % AcOH-ROH AcONa, (Pt) PhSH MeOH-H2O-KOH Ph S S Ph KH2PO4, (Pt) 100% Ex. 2: Oxidation of diphenyl sulfide to diphenyl sulfoxide, diphenyl sulfone and diphenyl-4-(phenylmercaptophenyl)sulfonium. O Ph S Ph O AcOH-H2O PhSPh AcOH-H2O (Pt) Ph S Ph 21,30V, (Pt) Cl , Br , SO4 , 0.94V O 96% MeCN-LiCLO 4 (Pt) PhS SPh2,ClO4 71% Electrochemical oxidation of amines Simple aliphatic amines have relatively low oxidation potentials. The course of oxidation depends on the structure of the amine, on the anode material and on the composition of the electrolyte. Ex.: amidoalkylation: -methoxyamides and –carbamates generate iminium ions under the influence of an acid which react then with sufficiently “nucleophilic” substrates (aromates, alkenes, C-acids etc.) by an electrophilic substitution. N MeOH, Et4NTsO -2e- COOMe N OMe H+ - MeOH COOMe Si(CH3)3 TiCl4 N O N COOMe , TsOH 89% COOMe N O COOMe OAc O N , TiCl4 (MeCO) CH , HCl 2 2 84% 69% COOMe N Ph2PCl, MeCO2H 89% P(O)(Ph)2 COOMe Me3SiCN, SnCl4 90% COMe COOMe COMe N N CN COOMe Electrochemical oxidation of aromatic systems Simple aromatic compounds (benzene, naphthalene, toluene, xylene) have relatively very positive oxidation potentials. A radical cation is usually formed which, in general, dimerizes or reacts with a nucleophile to yield a substitution product. Ex. 1: Electrooxidation of sufficiently “nucleophilic” aromatics. -eBu4NBF4 CH2Cl2 -e-2 H+ Ex. 2: Electrochemical nitration and cyanation of aromatics. NO2 -2e-, MeCN N2O4, Bu4NPF6 91% CN -2e-, MeOH NaCN 74% Ex. 3: ipso substitution in aromatics. OMe NC O CH2 H - CH2O - H+ -2eEt4NCN MeCN OMe OMe CN (95%) OMe Electrochemical oxidation of carboxylate anions In the electrochemical oxidation of carboxylate anions two types of reactive intermediates can be obtained: radicals which dimerize (the so called Kolbe reaction) and carbocations which can eliminate a proton, transform to isomeric carbocations and react with nucleophiles (the so-called Hofer-Moest reaction). R COO -e-CO2 R Rads -2e-CO2 -e- -CO2 R COO -e-CO2 MeCN R H alkene R R Rads R YO - R NHAc R OY -H+ alkene Direct Reductive Reactions Electrochemical reduction of functional group Ex. 1: Hydrogenation of multiple C-C bonds can proceed indirectly, electrocatalytically, or by a direct electron transfer to the multiple bond. The reaction can proceed with good selectivity. COOH COOH + 2e- Ex. 2: Reductive dehalogenation (also very selective). O O Cl + 2e-, Et NBr 4 DMF Br Cl 96% H Ex. 3: Electrochemical reduction of nitroaromatics. High yields of amines are obtained in the reduction of nitroaromatics carrying a group in the para position which prevents a rearrangement. R CO2H Me R = Me, + 6e(Pb), H2SO4 R = CO2H, + 6e(Pb), HCl NO2 NH2 92% NH2 100% Ex. 4: Reductive transformation of acrylonitrile to allylamine in a cell with a diaphragma. CN (Pb), 30% H3PO4 4 e- NH2 Electrochemically induced substitution The electrochemically generated anionic intermediates can react with “electrophiles” to generate substitution products. Ex. 1: Electrochemical alkylation and acylation can be achieved when alkylating or acylating agents (usually: anhydrides, halogenides of acids, their nitriles and DMF) are present in the reductive process of a convenient substrate. a) O S S S 2 e-, DMF Ac2O 90% b) CCl3 + Br H Cl 2 e-, LiClO4 DMF Cl O Br O Cl - Br Cl 40% O c) Remember SRN1 reactions? Pinacolizations and hydrodimerizations By electroreduction of aliphatic and aromatic aldehydes and ketones at cathodes with a different hydrogen over voltage and within a broad range of potentials, vicinal diols (pinacols) are formed. The new C-C bond is a result of radicals or of radical ions formed by a “single electron” reduction of the carbonyl group. The aromatic carbonyl compounds give higher yields of pinacols than aliphatic ones. Ex.: synthesis of 1,2-cyclopropanediols. O O OH OH + 2 eTHF, (MeCO)2O The intermolecular electroreductive hydrodimerization of a-bunsaturated compounds (aldehydes, ketones, esters and nitriles) is one of the most powerful electroorganic synthesis reactions. The best results were achieved in the case of commercially realized acrylonitrile dimerization to adiponitrile. CN + eEt4NOTs 1. CN 2. + e- H2C CH C N 2 H 2O 2 HO NC NC CN CN Initially performed in divided cells, at lead cathodes and in the presence of tosylate or ethylsulfate as supporting electrolyte. This has been improved later by using undivided cells. This has been made possible by applying highly selective cadmium cathodes and an optimal composition of the electrolyte; the supporting electrolyte consists is hexamethylene bis(dibutylethylammonium)phosphate jointly with corrosion inhibitors (NaHPO4, Na2B4O7, Na4-EDTA). Reductive Eliminations The cathodic reduction of the geminal polyhalogeno derivatives in he absence of both electrophiles and protons donors results in the formation of a carbanion that is stabilized by splitting off a further halogenide ions which in the presence of a sufficiently “nucleophilic” alkene adds on to the double bond and forms gemdihalogenocyclopropanes. Ex. 1: Electrosynthesis of 1,1-dichlorotrimethylcyclopropane from tetrachloromethane. CCl4 +2 e-, -2Cl CH3Cl CCl2 Cl -Cl CCl2 Me Me Me Me H Me Me 82% Cl Cl Ex. 2: Electrosynthesis of gem-difluorocyclopropane derivatives from dibromodifluoromethane. F Me CF2Br2 +2e-, -2Br CH2Cl2 CF2 Ph F Ph 57% Me Ex. 3: Electroreduction of vic-dihalogeno derivatives undergo a stereospecific trans-elimination to yield an alkene. This can be used for the synthesis of alkenes with strong internal stress, usually isolated as cycloaddition products of a reaction with dienes. Br +2e-, Et4NBF4 DMF Cl 100% Reductive Removal of protecting groups The electrochemical method for the removal of the protecting groups has a parallel in other cathodic reactions (electrohydrogenolysis and 1,2-elimantion of halogeno-derivatives). An advantage of this method in comparison with the chemical procedures is its high selectivity and mild reaction conditions. A high selectivity is achieved by combination of suitable electrophores-protecting groups, which sufficiently differ in reduction potentials (∆E1/2 200mV). Ex. 1: The difference in the values of reduction potentials of benzenesulfonyl and formethoxybenzoyl group enables their selective cleavage and hence also their application in the protection of the respective amino group, e.g. in the molecule of L-lysine below. O COOH N H MeO N H -2.4V SO2 -2.15V + e-, Me4NCl EtOH O COOH N H MeO NH2 90% Ex. 2: Synthesis of of 1,2-dibromo-1(3-cyclohexenyl)ethane from 4vinylcyclohexene. Br Br 2Br2 Br Br Br Py.HBr3 Br + e-, -1.2V(SCE) DMF Br Bu4NBF4 Br 62% Industrial Applications Advantages of electrochemical processes: - Fast and give generally good yields - Application to compounds thermally unstable - Simple and low cost (electricity and equipment) - No (or less) hazardous and toxic reagents - Large number of known laboratory processes - Better selectivity - Mechanistic, kinetic and thermodynamic data Limitations: - Industrial cells - Electrode material, Poisoning: Passivation Loss of selectivity - Solvent and supporting electrolyte Industrial Processes Product/Precursor Company Anode Cathode Yield Reduction of aromatic C=C bonds Asahi Chemical Pt 1,4-cyclohexadiene / Gulf Res. & Dev. C benzene Mitsubishi Chem. Pt Monsanto S.S 1,4-dihydronaphtalene / Standard Oil C naphtalene pipyridine / pyridine Robinson Bros. Pb 3,5-cyclohexadiene-1,2dicarboxilic acid / ophtalic BASF Pb acid Dihydro-2methoxynaphtalene / Hoechst C methoxynaphtalene Reduction of nitro group p-aminobenzoic acid / CECRI Pb p-nitrobenzoic acid 3-amino-p-cresol / CECRI Pb 3nitro-p-cresl CECRI Pb p-aminophenol / U of Southampton Pt p-nitrobenzene Constructors J. B. Pt Miles Laboratories Pb o-aminophenol / CECRI Pb o-nitrobenzene Oxidation of aliphatics and alcohols BASF PbO2 Acetylene dicarboxylic acid General aniline Pb Victor Wolf PbO2 decylaldehyde/ 1-decene ETHYL Pb isobutyric acid Kryschenko, et al PbO2/Ni methylethylcetone ESSO C BAYER Pt/Ti Propylene oxide INTERNE Pt/Ti KELLOGG graphite Hg C Hg Hg Pb 93 92 96 90 92 Pb 90 Pb 94 Hg/Cu 94 Sn/Cu 75 Cu 75 Hg/Cu Cu Cu C Zn 70 82 84 85 80 Pb Cu Cu Pb Cu Pt S.S S.S S.S 70 75 75 75 50 92 90 99 86 Industrial Cells In class Useful Bibliography 1) Marcus, R. A. Theory and Applications of Electron Transfers at Electrodes and in Solution. In Special Topics in Electrochemistry; Rock, P. A., Ed.; Elsevier: New York, 1977; pp 161-179. 2) Savéant, J-M. J. Am. Chem. Soc. 1987, 109, 6788. 3) Savéant, J-M. Dissociative Electron Transfer. In Advances in Electron Transfer Chemistry; Mariano, P. S., Ed.; JAI Press: New York, 1994; Vol. 4, p. 53-116. 4) Savéant, J-M.; Electron Transfer, Bond Breaking and Bond Formation. In Advances in Physical Organic Chemistry; Tidwell, T. T. Ed.; Academic Press: New York, 2000; Vol. 35, pp 177-192. 5) Volke, J. and Liska, F; Electrochemistry in Organic Synthesis. Springer-Verlag; Berlin Heidenberg, 1994. 6) Grimshaw, J.; Electrochemical Reactions and Mechanisms in Organic Chemistry. Elsevier: New York, 2000.