You are on page 1of 9
Journal of MOLECULAR STRUCTURE ELSEVIER Journal of Molecular Structure 378 (1996) 189-197 The infrared spectra and structure of acetylsalicylic acid (aspirin) and its oxyanion: an ab initio force field treatment LG. Binev*, B.A. Stamboliyska, Y.I. Binev Instcate of Organic Chemisty, Bulgarian Academy of Sciences, 1113 Sofa Bulgaria Received 25 September 1995; acepted 16 October 1995 Abstract ‘The structures of acetylsalicylic acid (aspirin) (I) and its oxyanion (If) have been studied by means of infrared spectra and ab initio 3-21 G force field calculations. The 3100-1100 cm region bands of both the aspirin molecule and its oxyanion have been assigned. The theoretical infrared data for the free aspirin anion are in good agreement with the experimental data for aspirin alkali-metal salts in dimethyl sulfoxide-d, The theoretical geometrical parameters for the isolated aspirin molecule are close to the literature X-ray diffraction data for its dimer in the solid state, except for those of the carboxy group, which participates directly in hydrogen bond formation. The changes in both the spectral and geometrical parameters, caused by the conversion of the aspirin molecule into the anion, are essential, but they are localized mainly within the carboxy group and the adjacent C-Ph bond. This is also true for the changes in the corresponding bond indices and electronic charges. Keywords: Infrared spectroscopy; Ab initio calculation; Acetylsalicylic acid 1, Introduction Raman studies on aspirin [4-24]. Most of them deal with methods of its identification Acetylsalicylic acid (aspirin) (1) was first pre- pared as a pure substance by A. Bichengriin and F, Hoffmann in 1897 [1,2] in the laboratories of F. Bayer and Co. in Elberfeld, Germany (2; its first pharmacological testing was performed in 1898 (2] by F. Hoffmann and H. Dreser [3]. There are thousands of publications on aspirin; most of them deal with its various (analgesic, antipyretic, antithrombotic, antirheumatic, etc.) physiological actions. Over 1300 studies on aspirin are cited in a recently published book [3]. ‘We have found in the literature 21 infrared or * Corresponding author. [6-9,13,14,18,24] and quantitative determination [9,14-18,20--22,24] as an individual substance or in multicomponent mixtures (medicines, etc.) (9,14,20-22,24], Other studies refer to the quanti- tative determination of salicylic acid [15] and moisture [19] in aspirin samples. The hydrogen bonds formed by aspirin molecules have been studied in detail [5,10-12,23]. We were suprised to have found in the literature neither a detailed vibrational assignment nor a normal coordinate analysis of the aspirin molecular vibrations. The vibrational spectra of the aspirin anion (II) have not been studied at all; a series of complexes of some transition metals with mixed ligands, (0022-2860/96/$15.00 © 1996 Elsevier Science B.V. All rights reserved SSDI 0022-2860(95)09170-X. 190 1G. Binev et al{Journal of Molecular Structure 378 (1996) 189-197 14 o7 ABSORBANCE UNITS 00 2000 1800 1600 1400 1200 WAVENUMBER (cm Fig. 1. Infrared spectrum (2000-1100 em“) of aspirin, 0.12 mol I~! in DMSO-d including aspirin, have been characterized by their fraction methods [26,27]. No similar study on infrared spectra [25] aspirin salts has been performed The crystal and molecular structure of aspirin Quantum chemical calculations of the aspirin has been determined by three-dimensional X-ray molecule have been carried out with regard to its 1a g gz 3 Sor z g z q oo A SNS 2000 va «00 Ta00 Tato WAVENUMBER (em“}) Fig 2. Infrared spectrum (2000-1100 em“) of the sodium salt of aspirin, 0.12 mol I~! in DMSO-d, 1G. Binev et al.jJournal of Molecular Structure 378 (1996) 189-197 191 electronic spectrum [28] and its solvolysis [29]. The electronic structure of salicylaldehyde [30] and the stability and acidity of salicylic acid [31] have recently been studied at the ab initio 6-31G level. There are in the literature no ab initio force field studies of either aspirin or its anion. Alcolea Palafox et al. [32] have recently published their ‘AMI and ab initio force field study on 4-methoxy- benzoic acid, which is structurally somewhat similar to aspirin. However, ab initio force field calculations of certain molecules [33-35] and their carbanions [33,34] and nitranions (35] have proved useful in the elucidation of the structure of those species. The purpose of the present study is to elucidate the structure and the force field of both aspirin and its anion by means of infrared spectra and ab initio force field calculations. The structure of the aspirin anion is also of general interest, as in living organisms part of the aspirin molecules is dissociated in the body liquids, so that certain physiological actions are assumed to be due to its solvated anions. The equilibrium concentrations in a few solvents [36-38] are the only information available at present for the aspirin anion. 2. Experimental Commercial aspirin (Medexport, Russia) was recrystallized from water and from diethyl ether. No differences between cither the melting points 4 H Roba yeh 0. =0 H -H ‘The aspirin molecule (133°C) or the infrared spectra of the two samples have been detected. Aspirin was converted into the corresponding salts by adding its solution in hexadeutero-dimethyl sulfoxide (DMSO-d,) to excess of dry sodium or potassium carbonate. The suspensions obtained were stirred for 1 min and then filtered. 0-CHyCOO—CgH,-COOH + M,CO; Aspirin + o-CH;COO—C,H,—COO-M* + MHCO; (M=Na and K) The metalation is practically complete: no bands of the parent molecule are seen in the spectrum after metalation (see Figs. | and 2). Carbonate or hydro- carbonate bands are also not observed in the spec- trum (Fig. 2). The spectra of the sodium and potassium salts of aspirin in DMSO-d, are identical. The infrared spectra were recorded with a Bruker IFS-113v FTIR spectrophotometer in CaF, cells of path length 0.13 mm, at a resolution of 1 cm“! and 50 scans, using a DTGS detector. 3. Computations The ab initio force field computations were performed using the GAMess software [39] at the HF 3-12G level. No scaling was done for the ab initio force field obtained. ‘The aspirin anion 192 1G. Binev et al.j Journal of Molecular Structure 378 (1996) 189-197 4, Results and discussion 4.1. Infrared spectra 4.1.1. The aspirin molecule (I) ‘A fragment of the infrared spectrum of aspirin in DMSO-d, is shown in Fig. 1. The numerical data for the frequencies and intensities of the bands are listed in Table 1 together with the corresponding theoretical values. No experimental data for the Table 1 vou band were given in Table 1, as it is very broad because of the formation of strong hydrogen bonds mainly with the solvent. Table 1 demon- strates the good agreement between the scaled theoretical and the experimental frequencies. The mean deviation between them is 14 cm”! (0.7%), this value being close to those for other molecules studied by us: 0.8% for the isomeric (cyanophenyl)acetonitriles [34] and 0.9% for o- sulfobenzimide (saccharin) [35]. ‘Theoretical (ab initio 3-21G) and experimental (DMSO-dg) infrared data for the aspirin molecule Ne Abinto forge Geld Experimental* : : 4 “Approximate 7 4 (em) (om (km mot) description? (em) (km mot!) 1 3862 467 1382 100 vox very broed > M7 3068 2 100 ven ine) son 17 3 391 304s 92 100 re 8) sore to 4 3386 3041 Ld 99 voy (ring) 5 3359 3017 35 100 vex (ring) $ sas 3008 3a 100 v8, aoe ° , 07 ool i 100 vt 2x a 8 yaa ria 16 100 v8, 2912 31 1767 36 204 1304 sea en e cursom {Oe 10 1936 1743 408.1 82 veuo (carboxy group) 1706 172 " 1780 1604 1027 6 v6c 23 8B 12 boce 1606 180 2 iat 180 a2 68 vg 16 fe 12 Soce 1578 mM B 1663 149 7 83 by 26 100, 20 nascco re an 4 1657 1494 321 59 muyeco, 15 688, 15 1635 1474 288 15 mcco, 21 6, ass 22 6 162s tas m3 5 dee 16 feces vec 138 ira " 1866 i312 40 onan vail 26 1B 1502 1385 2068 25 byoes 21 Ye-ons 15 Boces 4 omce Il Bn 1370 sal i336 326 » 1430 1250 ni 33 6a 20 1387 sz 189.1 39 vege, 29 5, 26 Yee 1250 n3 21 1347 1216 499 69 6 24 vee 1220 487 » 1308 list 402 26 bela, 22 vBocs 12 vec 1197 167 aes 23 1299 us 280 ee 13.6 net 79 24 1258 1136 542 7B vee, 14 68 1133 16 * Scaled, according to the correlation equation (:experimental)] em (0.89521(3-216) + 10.2} om“ (344 ° Vibrational modes: v, stretching; é, deformation (all kinds of); 7, torsion; superscript s, symmetric; superscript as, asymmetric; superscript ip, in-plane. The numbers before the mode symbols indicate per cent contribution (10 or more) of a given mode to the corresponding normal vibration, according to the energy distribution matrix [3]. © Measured after having decomposed the complex bands into components 4 Followed by 33 lower-frequency normal vibrations, 1G. Biney et al,{Journal of Molecular Structure 378 (1996) 189-197 193 We correlated the unscaled theoretical frequen- cies with the experimental ones for the vy (ring), Yeu,» Ueno (carboxy), vg and vig bands (11 frequen- cies altogether), whose assignment is found to be entirely unambiguous for 4-methoxybenzoic acid (4-21G theory, solid-state spectra, Ref. (32) and for aspirin (3-21G theory, DMSO-d, solution spec- tra, Table 1). The two linear correlations obtained show correlation coefficients R > 0.999, so the results seem close in reliability. The calculations resolve and locate well the two carbonyl stretching vibrations, those of the ester and the carboxy group. The ester vcuo band in the experimental spectrum has a clearly seen shoulder (Fig. 1, Table 1). Its appearance is probably due to Fermi resonance; the alternative Table 2 explanation, i.e. a conformational equilibrium, is much less likely, as this band preserves its doublet character in the spectrum of the anion (see below), the carboxy vc. band is a singlet, etc. (see also 42). The theory predicts qualitatively well the high integrated intensities for the strongest bands in the experimental spectrum, namely vc_o (ester), Veo (carboxy) and vég¢ (Table 1). 4.1.2. The aspirin anion (II) The spectra of the sodium and potassium salts of aspirin in DMSO-d, are identical; hence, there are no strong anion/counter ion interactions, and we can assume that in the solutions studied the aspirin anion is either solvent separated or is a kinetically Theoretical (ab initio 3-21G) and experimental (OMSO-ds) infrared date for the aspirin anion No [Ab initio force field Experimental® v Pa A Approximate . 4 (em) (em («mm mot!) description” (en) (km mot!) 1 3400 3054 198 99 vew (cing) 2 3395 3049 91 100 ven (ring) } ms “ 3 3313 3030 409 100 ven (ring) 4 3351 3010 10. 100 ve (ring) jo a 5 3u21 2983 81 100 v2, 2992 sa 6 3298 2963 259 100 v2, 2933, 12 7 3223, 2895 248 100 v8, 2912 28 8 1978 W781 4306 85 ve-0 { ney aoe 1750 912 9 1853 1669 6140 83 4, 1609 154 0 165 1590 246 62 Hee, 24 5g 1598 48.2 " 1137 1365 23 11 wees 16 8 1365 36.7 2 1672 1507 266 83 na,coos 16.8%, B 1640.2 478s 90.3 52 By. 45 bcoc 74 90 tem test 0 roreco 2 8, } 1s 1607 1499 124 65 8B. 29 vec Mat 320 16 1533 1383 100.0 91 8, 1349 442 7 ass 1306 606.3 S4 uo, I wdc 1370 7 18 1397 1261 303 45 6840 17 bdo, 19 1363 1230 2187 45 vebes 17 BB, 15 vec. 1230 10s 10 deco ey 1332 1203 n2 $6 €y. 24 vee 21 1325, 1196 129.5 22.68, 20 voce lugs 106 13 GR 10 Yee 22 1262 1140 16 63 vec, 16 AB. 12 Ecce 1158 18 * See Footnotes to Table 1 4 Followed by 32 lower-frequency normal vibrations. 194 1G. Binev et al,{Journal of Molecular Structure 378 (1996) 189-197 free ion. This conclusion makes it possible to com- pare, in this work, the experimental infrared data for the aspirin alkali-metal salts with the theoretical data calculated for the free aspirin anion. ‘A fragment of the infrared spectrum of the aspirin anion is shown in Fig. 2. The numerical frequency and intensity data for the experimental bands are listed in Table 2 together with the corre- sponding theoretical values. As seen in Table 2, the agreement between the scaled theoretical and the experimental frequencies is almost as good as that in the case of the aspirin molecule (see above). The mean deviation between them is 17 cm”! (0.9%); this value coincides with that for the isomeric (cyanophenylacetonitrile carbanions (0.9%) (34] and is slightly lower than that for the o-sulfo- benzimide nitranion (1.3%) (35]. The calculations predict well the essential changes in the spectrum of the aspirin molecule, which accompany its conver- sion into the anion (see Figs. 1 and 2, and Tables | and 2). Removing the proton (and hence the disappear- ance of bands | and 18, Table 1) after the conver- sion of the carboxy group into the carboxylate anion leads to equalization of the two C=O bonds in the anion (see 4.2). The two bands, ve. and véo., appear instead of the vc_o (carboxy) and Yc_o bands. The frequency undergoes a strong, decrease: predicted 256 cm™', measured 216 cm (calculated by the mean vco- value). The integrated intensity shows a moderate increase: predicted 3.0-fold, measured 1.9-fold (calculated for the sum dé; + Ato;). Let us follow the frequency and intensity varia- tions in the other analytically important bands, which accompany the conversion of the aspirin molecule into the corresponding oxyanion. (i) The frequencies of the series of C-H stretch- ing bands, both ring vc and ycy,, do not change essentially. The theory predicts a six-fold increase in their overall intensity while the experiment shows that a slight decrease in this value accom- panies the conversion of the aspirin molecule into the anion. We are not surprised by this result, as the same result has been obtained qualitatively for the phenylacetonitrile and (cyanophenylacetonitrile carbanions [34] and the o-sulfobenzimide nitranion BS} (ii) The theory predicts and the experimental confirms both a slight frequency decrease and a slight intensity increase for the veo (ester) band. This band preserves its doublet character (a basic band with a shoulder). Because of the essential change in the intensity ratio, however, the lower- frequency component (the shoulder in the spectrum of the molecule) becomes the basic band in the spectrum of the oxyanion, (iii) the vg and v9 skeletal bands of the phenylene ring undergo moderate frequency decreases: predicted 12 cm~', measured 11 cm™! (mean values). Instead of the predicted 1.6-fold decrease, the overall intensity of these bands shows a 2.4-fold increase. This result is surprising, as we have found in previous studies [34] good agreement between the predicted and found strong intensity changes of the phenyl and phenylene ring skeletal bands, caused by the conversion of the series of nitriles into carbanions. (iv) The conversion of the aspirin molecule into the oxyanion causes moderate frequency decreases of the vac and vég¢ bands: predicted 8 cm™', measured 11 cm7' (mean values). The sum of the corresponding integrated intensities of the anion is also higher than that of the neutral molecule: pre- dicted 1.7-fold, measured 1.2-fold. The vag¢ band of the aspirin molecule is more intense than the vec band (less frequently found among the spectra of esters; as a rule the vac bands are very strong, comparable in intensity to the corre- sponding vc_o bands (40,41]), and in the spectrum of the aspirin anion these bands are equally intense. Moderate or slight variations in both frequencies and intensities have been found in the bands of the other normal vibrations. As seen above, however, the strong spectral changes are localized within the carboxy/carboxylato groups. This result is qualita- tively different from those obtained for a series of conjugated carbanions [34,41], nitranions and oxyanions [35], where the conversion of the molecules into anions causes strong changes in the whole spectra, One of the causes for this difference is the localization of the anionic charge mainly within the carboxylato group (see 4.2) whereas the anionic charges of the above- mentioned conjugated anions are strongly delocalized [34,35,41] 1G. Biney et al,|Journal of Molecular Structure 378 (1996) 189-197 195 4.2, Structure of the aspirin molecule and its oxyanion According to both the X-ray diffraction experi- ment [26] and the ab initio theory (this work), except for the hydrogen atoms of the methyl group, the aspirin molecule is composed of atoms lying approximately in two planes: one of the phenylene ring and the carboxy group, and the other of the acetoxy group. The angle between these planes has been experimentally found to be 84°45’ (for the aspirin dimer in the solid state) (26; the same angle for the isolated aspirin molecule has been theoretically estimated at 77°. The corresponding value for the free aspirin anion is 62°. The twisting angles of the carboxy/carboxylato groups with respect to the phenylene plane have been estimated to be 16 and 12° respectively. The total energy of the “open” conformation of the aspirin molecule (I) is —641.355951 h. The corresponding “closed” conformation (planar, with an intramolecular hydrogen bond) is, 13.3 kcal mol”! less stable, ie. in this case the Table 3 Certain internuclear distances (A) in the aspirin molecule and its anion Internuclear Molecule ‘Anion distances* = ——__—________ —___ (Experimental)®* (Ab initio)’ (Ab initio)’ HOH) 1.1207) 0.968 HOO) 1.28714) 1353 1.255 nO") 1.235(4) 1.205 1283 HCC) 1.48965) L474 1354 HEC) 1.39215) 1.390 1386 nce) 13875) 1376 1379 (CC) 1.381(6) 1.386 1,386 HEC) 13876) 1379 1379 HCC) 138508) 1.385 1.388 CO}, 1.40218) 1.376 1386 HCHO") 1.183(5) 1.190 1.197 HCO") 1.364(8) 1387 1376 nc"c") —_1.496(4) 1.508 1.305 * For atom numbering see I and I ° X-ray diffraction data for the dimer in the solid state (Ref. (26). © Standard deviations in parentheses, as unit in the last place * Isolated particles. ortho repulsion is more effective than the hydrogen bond formation. The total energy of the aspirin anion is —640.792126 h. This value corresponds to 1479.9 kJ mol! deprotonation energy. The latter lies in the 1469-1497 kJ mol interval of deprotonation energies of cyanoacetic acid, o-sulfobenzimide and the isomeric hydroxybenz~ aldehydes [35,41]. All these compounds are moderately strong (or moderately weak, but acidic to litmus) O-H or N-H acids. Certain internuclear distances (both lengths) of the particles studied are listed in Table 3. As seen, there is good agreement between the experimental and the theoretical values. Exceptions (deviations of 0.03 A or more) are typical only for the carboxy group itself (the first three data sets in Table 3), whose atoms participate directly in hydrogen bond formation. This result leads us to believe that the theoretical bond lengths for the aspirin anion (Table 3) are also reliable. The data in Table 3 make it possible to conclude that the con- version of the aspirin molecule into the oxyanion causes essential (more than 0.04 A) bond length changes only in the carboxy/carboxylato groups (the first three data sets in Table 3) and in the adjacent C-Ph bond (the next data set). The latter result can be explained qualitatively as follows: the carboxylato group is much less con- jugated with the aromatic ring than the carboxy group (see also Table 5). Tabled Certain ab initio bond angles (deg) of the aspirin molecule and its anion Bond angles* Molecule Anion H'O"C 1123 : Octo” 1215 1293 occ 138 uss ec 125.1 1252 ec uss 1173 ccc i194 120.2 ccot 123.6 128.1 ec’ 120.7 1208 CCH" N81 1180 orc! 1268 1278 ofcl'o" 1182 ng. ofctic? 1186 1163 ccc 1202 1196 * For atom numbering see I and I 196 1G. Binev et al,{Journal of Molecular Structure 378 (1996) 189-197 Table $ Certain ab initio 3-21G bond indices ofthe aspirin molecule and its anion Bonds* Bond indices Molecule Anion OF-Ht 0.777 Co 0.864 1.428 co" 173 1526 e-ct 0.956 0.208 ccs 1.299 1312 cc! 1365 1364 ca 0920 0938 ec 1.405 Lala on" 0.929 0.952 Co! 0.803 0.766 Hof 0.775 0,802 og 0.863 0.868 cho" 1871 1.826 * For atom numbering see I and II. Certain bond angles of the particles studied are listed in Table 4. It is seen there that the conversion of the aspirin molecule into the oxyanion causes an essential change in the OCO angle of the carboxy/ carboxylato group only. Certain bond indices of the particles studied are listed in Table 4. It is seen there that the bond index changes which accompany the conversion of the aspirin molecule into the oxyanion are localized mainly in the COOH/COO™ groups and in the (molecule) Charges: q= 0.434 (oxyanion) Charges: q= 0.510 (molecule / oxyanion) Charge changes: Aq * adjacent C-Ph bond. This is in agreement with the bond length variations (Table 3). The comparison of the net charge distribution over fragments of the species studied is given in Scheme 1. The values of the net charge changes which accompany the conversion of the aspirin molecule into the oxyanion indicate that the oxy- anionic charge remains localized within the car- boxylato group, like the other structural changes (see above). This conclusion is in contrast to the data for conjugated carbanions (34,41), nitranions and oxyanions [35], where the structural changes accompanying the conversion of the neutral molecules into anions are spread over the whole particles. Conclusions A comparison of calculated with measured infra- red data can be used as a test for the reliability of the structural ab initio predictions for various molecules and anions. These predictions can be very useful in cases of molecules and ions for which experimental structural parameters are inaccessible or unknown. Acknowledgments ‘This work has been supported by the Bulgarian CHsCOO— C,H, — COOH, 0368 0.066 CHyCOO—— _Celty ——_COO™ 0.085 0575 0.283 0681 ‘Scheme 1, Ab initio net charges of the aspirin molecule and its oxyanion. 1G. Binev et al,|Journal of Molecular Structure 378 (1996) 189-197 197 National Fund of Scientific Researches, Project X-442. References 11] GY. Bykov, The Organic Chemistry History (Russ.), Vol. 2, Nauka, Moskva, 1978, p. 108, 1 R, Ellmer, Chem. Labor. Betr., 29 (1978) 82; Chem. Abstr, 89: 162372n, (3) LR. Vane and R.M. Bottling (Eds.), Aspirin and Other Salicylates, Chapman and Hall, London, 1992. [4] L. Kahovee and K.W.F. Kohlrausch, Monatsh., 74 (1943) 333. [5] GV.LN. Murty and TR. Seshardi, Proc. Indian Acad. Sci., Sect. A, 19 (1944) 17. [6] OR. Sammul, W.L. Brannon and A.L. Hayden, J. Assoc. Off. Agric, Chem, 47 (1964) 918 (7M. Tatsuzawa, Eisei Shikenjo Hokoku, 83 (1965) 58; ‘Chem. Abstr., 65: 19929h. [8] SIM. Reprintseva, N.V. Fedorovich and V.L. Dragua, Vestsi. Akad. Nauk B. SSR, Ser. Fiz.Energ, Nauk, (1972) 124; Chem. Abstr, 78: 101903n. (9] BB. Lal and S.P. Varma, Indian J. Phys., SIA (1977) 324 (10) AF. Mynka and NM. Turkevich, Farm. Zh.(Kiev), 1 (1979) 27, (111 Y. Nakai, 8. Nakajima, K. Yamamoto, K. Terada and T. Konno, Chem. Pharm. Bull, 28 (1980) 652. [12] MJ. Wojcik, Chem. Phys. Lett, 83 (1981) 503, [13] A.F. Mynka and M.L. Lyutaya, Farm. Zh. (Kiev), 4 (1982) 38. (14] G.E, Zuber, RJ. Warren, P.P. Begosh and E.L. O'Donnell, ‘Anal. Chem., 56 (1984) 2935, 15] ASR. Krishnamurthy, R. Shailaja and 8. Husain, Indian Drugs, 23 (1986) 513 (16] EM. Suzuki and W.R. Gresham, J. Forensic Sci. 31 (1986) 1292, [17] W.N. Hansen, JF. Andrew and GJ. Hansen, Appl. Spectrosc., 41 (1987) 562. [18] R.A. Lodder and G.M. Hiefije, Appl. Spectrose., 42 (1988) 556. (19] RM. Issa, M.K. El-Marsafy and M.M. Gohar, An. Quim., ‘Ser. B, 84 (1988) 312. (20) S. Shah and LT. Taylor, LC-GC, 7 (1989) 340, 342, 346; Chem. Abstr. 110: 199318e. (21) F. Ahmad, M. Akbar, S. Aminuddin and M. Siddia, Pak, J. Sci. Ind. Res., 32 (1989) 155. 22] MJ. Walters, RJ. Ayers and D.J, Brown, Anal. Chem.,73 (1990) 904, 23) R. Dilshard, LR. Dixon, W.O. George, R. Levis, B. Minty and R. Upton, Proc. Soc. Photo-Opt. Instrum. Eng., 2089 (1993) 150; Chem. Abstr., 121: 923984, [24 CG. Kontoyannis and M. Orkoula, Talanta, 41 (1994) 1981 [25] AE. Shvelashvili, M.G. Tskitishvili, LI. Mikadze and MN. Chrelashvii, Izy. Akad, Nauk Gruz., Ser. Khim., 17 (1991) 167; Chem. Abst, 118: 182157. [26] PJ. Wheatley, . Chem. Soc-, Suppl, (1964) 6036. [27] D. Clement, A. Verain and A. Durf, Ann. Pharm. Fr. 36 (1978) 10. [28] AS. Bl-Shahawy, Spectrochim. Acta, Part A, 44 (1988) 903. [29] H. Umeyama and S. Nakagawa, Chem. Pharm. Bull, 25 (977) 1671 {30} K. Inuzuka, Nippon Kagaku Kaishi, 9 (1994) 771 [BI] G. Alagona and C. Ghio, J. Mol. Lig, 61 (1994) | [52] M. Alcolea Palafox, M. Gil and J.L. Nunez, Appl Spectrose., 48 (1994) 27 [3] G. Raabe, E. Zobel, J. Fleischhauer, P. Gerdes, D. Mannes, E. Miller and D. Endes, Z. Naturforsch., Teil ‘A, 47 (1990) 275. (34) LG. Binev, LA. Tsenov, EA. Velcheva and LN, Juchnovski, J. Mol. Struct, 344 (1995) 205; J. Mol. Struct, 378 (1996) 133-146. [35] 1.G. Binev, B.A. Stamboliyska and E.A. Velcheva, Spectro- chim, Acta, Part A, 52 (1996) in press. 6] K.1. Evstratova, N.A. E] Rabbat and N.A. Kupina, Zh. Fiz. Khim., 48 (1974) 1751 (37) K.1 Bvstratova, A.A. Kochegina, N.A. Kupina and V.V, Belozerskaya, Elektrokhimiya, 12 (1976) 677. {38} A. Fini, P. de Maria, A. Guarnieri and L. Varoli, J. Pharm, Sci, 76 (1987) 4. [39] M.W. Schmidt, K-K. Baldridge, J.A. Boatz, $.T. Elbert, MS. Gordon, JJ. Jensen, S. Koseki, N. Matsunaga, K.A. Nguyen, S. Su, T.L, Windus, M. Dupuis and J.A. Montgomery, J. Comput. Chem., 14 (1993) 1347. {40} G. Andreev, Molecular Spectroscopy (Bulg), _P. Khilendarski University of Plovdiv, Plovdiv, 1989, p. 116. [41] B. Stamboliyska, E. Velcheva, Y. Binev and J. Juchnovski, CR Acad, Bulg. Sei, 47 (1994) 69.

You might also like