Computational modelling for the clustering degree in the ... - CiteSeerX

9 downloads 4506 Views 186KB Size Report
Recent computational findings of temperature increase of clustering degree in several quite ... It has been found in the model studies that the clustering degree.
ARTICLE IN PRESS

Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423 www.elsevier.com/locate/jqsrt

Computational modelling for the clustering degree in the saturated steam and the water-containing complexes in the atmosphere Zdeneˇk Slaninaa,, Filip Uhlı´ kb, Shyi-Long Leec, Shigeru Nagasea a

Department of Theoretical Molecular Science, Institute for Molecular Science, Myodaiji, Okazaki 444-8585, Aichi, Japan b School of Science, Charles University, 128 43 Prague 2, Czech Republic c Department of Chemistry, National Chung-Cheng University, Ming-Hsiung, Chia-Yi 621, Taiwan Received 7 September 2004; accepted 2 May 2005

Abstract Recent computational findings of temperature increase of clustering degree in several quite different saturated vapors are analyzed further. A thermodynamic proof is presented, showing that this event should be rather common, if not general. Illustrations are based on the saturated steam and consequences for the atmosphere are discussed with both homo- and hetero-clustering. MP2 ¼ FC=6-311G  results for the H2 O  N2 (the best stabilization energy 1:62 kcal=mol, which corresponds to 1:44 kcal=mol with the G3 theory) and H2 O  O2 (the best stabilization energy 0:96 kcal=mol) hetero-dimers, and the G1-theory ðO2 Þ2 data (the best stabilization energy 0:62 kcal=mol) are reported. The water-dimer thermodynamics is recomputed in an anharmonic regime and a remarkable agreement with experiment is found. The results have some significance for the atmospheric greenhouse effect. r 2005 Elsevier Ltd. All rights reserved. Keywords: Atmospheric complexes; Water clusters; Hydrogen-bonded heteroclusters; Greenhouse effect; Atmospheric cluster populations; Anharmonic water-dimer thermodynamics

Corresponding author.

E-mail address: [email protected] (Z. Slanina). 0022-4073/$ - see front matter r 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.jqsrt.2005.05.065

ARTICLE IN PRESS 416

Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

1. Introduction Cluster studies have developed into a considerably broad field (see, e.g., [1–5]), covering aggregates of very different types. Although their experimental generation is frequently rather far from thermodynamic equilibrium, computational modelling of clusters has mostly worked— owing to obvious reasons—with dimers under the equilibrium conditions. Water clusters have been studied very intensively [6–13], including water hetero-clusters [3,14–19]. In order to estimate the possible effects of water clusters of any dimension, a thermodynamic extrapolation scheme has been applied [20,21]. Relative populations of water clusters in saturated steam (over ice or liquid water) have also been computed [22–25] based on the observed saturation pressure. It has been found in the model studies that the clustering degree increases with increasing temperature. It may be a surprising result but in fact it can be easily rationalized. While the equilibrium constants for cluster formation decrease with temperature, the saturated pressure increases. It is just the competition between these two terms which decides the final temperature behavior. As several computational approximations have to be used with the water clusters, some simple systems have also been investigated like dimerization in the argon-saturated vapor [26] with the dimerization equilibrium constant from advanced evaluations [27]. A temperature increase of the dimeric fraction has again been found. Similar results were obtained for Mg [28] and carbon [29] vapors. In this report, this interesting computational finding is further analyzed without any reference to a particular potential function or approximation, i.e., from a thermodynamic point of view. A simple thermodynamic criterion for the temperature enhancement of clustering degree is suggested and relation to the atmosphere is discussed ((H2 OÞ2 , H2 O  N2 , and H2 O  O2 ).

2. Clusters in saturated vapor A substance A exists in condensed (liquid or solid) and gas phases and they are in thermodynamic gas–liquid or gas–solid equilibrium. Moreover, in the gas phase clusters Ai are present as follows: iAðgÞ ¼ Ai ðgÞ

ði ¼ 2; . . .Þ,

(1)

described by the equilibrium constants: K p;i ¼

pAi piA

ði ¼ 2; . . .Þ.

(2)

The temperature dependency of K p;i is given by the van’t Hoff equation: d ln K p;i DH 0i ¼ dT RT 2

ði ¼ 2; . . .Þ.

(3)

where DH 0i denotes the (negative) standard change of enthalpy upon the respective i-merization.

ARTICLE IN PRESS Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

417

The total pressure P is the sum of all the partial pressures derived from Eqs. (2): P ¼ pA þ

1 X

piA K p;i .

(4)

i¼2

Eq. (4) can also be expressed using the monomeric molar fraction x1 : 1 ¼ x1 þ

1 X

xi1 Pi1 K p;i .

(5)

i¼2

Generally speaking, this algebraic equation cannot be solved in a closed form. However, we do not need the algebraic solution—it is just enough for our purpose to express the temperature derivative dx1 =dT. Moreover, we are not interested in a general pressure but in the saturated pressure. Let us suppose that the saturated pressure obeys the Clausius–Clapeyron equation: d ln P DH vap , ¼ dT RT 2

(6)

where DH vap is the (positive) enthalpy of vaporization. The temperature derivative dx1 =dT can be expressed as ! 1 1 X X dx1 DH 0i þ ði  1ÞDH vap ixi ¼ x1 xi (7)  1þ dT RT 2 i¼2 i¼2 where the xi terms are given as i1 xi ¼ xi1 K p;i . 1 P

(8)

Eq. (7) suggests a simple sufficient (although unnecessary) condition for the temperature decrease of x1 in the saturated vapor (i.e., for temperature increase of the clustering degree 1  x1 ). As long as DH vap 4jDH 0i j=ði  1Þ for every iX2, the temperature derivative dx1 =dT is negative and thus, the clustering degree increases with temperature in the saturated vapor. Table 1 illustrates the rule on the water dimer and other water oligomers as steam is a pertinent system [32,33] for such observations.

3. H2 O  N2 and H2 O  O2 hetero-dimers The importance of molecular complexes for spectral records and for thermophysical properties of the atmosphere has clearly been established [34–36]. There is a particular sub-group in the set of molecular complexes relevant to the atmosphere—the complexes with water, both water clusters and hetero-clusters with water as one component. According to the previous section, the contents of molecular complexes in saturated (or nearly saturated vapors) increase with temperature. The atmosphere can also be close to its saturation with water. Hence, the contents of water dimers or even of water-containing hetero-dimers can actually increase with warming up of the atmosphere. Strictly speaking, the contents of water clusters should increase under near-saturation conditions—the case of hetero-clusters is not covered by the above treatment and should be given a special numerical simulation. Let us consider a second component, B, that is not in a

ARTICLE IN PRESS Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

418

Table 1 Comparison of the DH 0i =ði  1Þ and DH vap terms (kcal/mol) for water System

i

T (K)

DH 0i =ði  1Þ

DH vap a

H2 O

2

298.15 400 500 298.15

4.3b 3.8b 3.5b 3.5c 6.0c 7.5c 7.5c 7.7c

10.5 9.4 7.8 10.5 10.5 10.5 10.5 10.5

2 3 4 5 6 a

Refs. [30,31]. Ref. [25]. c The G3 values from Ref. [60]. b

contact with its condensed form. For the total pressure P, two simplified components can be then considered, P ¼ PA þ P0 , the saturated (water) pressure PA (heat of vaporization DH A;vap ) and the pressure P0 as a constant pressure of the B component. It can be shown that a more general sufficient (though not necessary) condition for the temperature increase of the clustering degree can again be formulated, if some further simplifications are accepted: DH A;vap 4

jDH 0Ai j P i  1 PA

for every iX2,

(9)

DH A;vap 4

jDH 0Bi j P i  1 PA

for every iX2,

(10)

and finally DH A;vap 4

jDH 0Ai Bj j P i þ j  1 PA

for every pair of iX1 and jX1

(11)

(where DH 0Ai , DH 0Bi , and DH 0Ai Bj have an analogous meaning as the DH 0i terms in Eq. (3)). As for the Earth’s atmosphere, the P=PA scaling factor makes the conditions less convenient than in the homo-cluster case and a full numerical modelling should be applied instead. Once the B species is eliminated, which also implies P ¼ PA , the new rule is reduced back to the rule derived for the one-component saturated vapor. Clearly enough, those water-containing homo- and hetero-dimers are also greenhouse-effect agents, somewhat different from free water. Hence, it make sense to clarify if the water-containing complexes themselves could, if increased, contribute to some additional atmospheric warming up. This hypothesis can only be studied computationally at present. Anyhow, it would be interesting to estimate what kind of contribution to the greenhouse effect can be expected from watercontaining homo- and hetero-dimers. There are two particularly interesting water-containing hetero-dimers, viz. H2 O  N2 and H2 O  O2 . In order to investigate their temperature development in a mixture of air and saturated

ARTICLE IN PRESS Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

419

steam, we have been performing ab initio computations with the Gaussian program package [37]. The computations are carried out at the second-order Møller–Plesset (MP2) perturbation treatment with the frozen-core option ðMP2 ¼ FCÞ in the standard 6-311G basis set, i.e., the MP2 ¼ FC=6-311G or UMP2 ¼ FC=6-311G treatment for the H2 O  N2 and H2 O  O2 species, respectively. In both cases, three stationary points have been located. If we, somewhat arbitrarily, define hydrogen bonds here simply as H–O or H–N long-range linkages shorter than 3 A˚, the stationary points can be described as exhibiting two or one such hydrogen bond. In the H2 O  N2 case, the lowest-energy structure is non-planar with two hydrogen bonds 2.76 and 2.80 A˚ long. Its MP2 ¼ FC=6-311G dimerization potential energy is DE ¼ 1:62 kcal=mol. It is followed by a planar structure with one hydrogen bond of 2.49 A˚ and dimerization energy of DE ¼ 1:59 kcal=mol. The third lowest structure is a planar species with two hydrogen bonds of 2.68 and 2.87 A˚ and a stabilization energy slightly less than 1:59 kcal=mol. The decrease in the potential energy upon dimerization is by about 0.5 kcal/mol smaller for the H2 O  O2 species. The lowest-energy structure is also non-planar in this case; it has two hydrogen bonds of 2.75 and 2.87 A˚, and the related UMP2 ¼ FC=6-311G dimerization potential energy is DE ¼ 0:96 kcal=mol. The second lowest isomer is a planar structure with two hydrogen bonds 2.65 and 2.68 A˚ long, stabilized by 0:91 kcal=mol. Finally, the third species is a C 2v planar form with two hydrogen bonds 2.72 A˚ long and the dimerization energy DE ¼ 0:81 kcal=mol. The listed dimerization potential energy terms are without the BSSE correction [38]. Although the physical nature of the counterpoise approach has been clarified [39], its accuracy is not always satisfactory. In order to avoid the BSSE correction, one can consider an application of some version of the developing G-theories [40–42]. The G1 theory [40] applies the full fourth-order Møller–Plesset (MP4) treatment with the 6-311G , 6-311 þ G , and 6-311G(2df,p) basis sets. In addition, a quadratic configuration interaction with all singles and doubles (QCISD) and with a quasi-perturbative treatment of higher excitations (QCISD(T)) is applied with the 6-311G basis set. The effects of polarization space extension and diffuse functions are then assumed to be additive at the MP4 level, and a correction formula for further basis set incompleteness is applied. The G-theories are aiming [40–42] at a precise description of thermochemistry; however, they have not been tested for the description of weak molecular complexes. If the H2 O  N2 case is recomputed at the G3 theory [42], we get the dimerization energy of 1:44 kcal=mol (the potential-energy change only, without the zero-point energy), i.e., not very different from the MP2 ¼ FC=6-311G term without the BSSE correction.

4. ðO2 Þ2 homo-dimers In order to observe the performance of the new tool, we have applied the G1-theory [40] to the three structures considered for ðO2 Þ2 in our previous evaluations [43]. The T-shape C 2v , rhomboid C 2h , and linear D1h exhibit the G1 dimerization energies of 0:62, 0:51, and 0:48 kcal=mol, respectively. In order to carry out through the computations, the frozen-core approach had to be applied, although not used in the original G1 scheme [40]. The potential-energy terms should, however, be further corrected for the zero-point vibrational energy and for temperature effects in order to be compared with the experiment [44].

ARTICLE IN PRESS 420

Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

Table 2 The standarda Gibbs-energy change DG02 for the gas-phase water dimerization evaluatedb in an anharmonic approach and the related observed valuesc T (K)

372.4 373.0 423.0 573.15

DG02 (kcal/mol) Calculated

Observed

3.24 3.25 4.94 6.51

3.33 3.08 4.00 6.51

a

The standard state—ideal gas phase at 101 325 Pa pressure. The potential-energy change evaluated with the G3 theory and the anharmonic partition functions with the MP2 ¼ FC=6-311 þ þG approach, using the Gaussian 03 program package [54]. c Ref. [25] and therein quoted data. b

Orlando et al. studied [44] the temperature dependence of collision-induced absorption by oxygen over the temperature range 225–356 K. They obtained a heat of formation ðDH f Þ for ðO2 Þ2 as 1:1  0:5 kcal=mol. Long and Ewing reported [45,46] a DE f term of 0:53 kcal=mol near 100 K which corresponds [44] to a DH f of about 0:75 kcal=mol (for a wider, diversified perspective of tetraoxygen, cf. also papers [47–53]). Overall, the G1 results seem to be in an encouraging agreement with the observed values (interestingly enough, a much older treatment [47] is also in a reasonable agreement). However, from a practical point of view, the H2 O  N2 system should be given preference in submission to still higher levels of theory (also because it is free of open-shell difficulties). As we are aiming at concentration terms, the quality of the partition functions is important, too. In order to have an insight into the computational reliability, the standard Gibbs-energy change DG02 for the gas-phase water dimerization is evaluated here in anharmonic rather then traditional harmonic approach (Table 2). While the potential-energy change is extracted from the G3 theory [42] computations, all the terms related to partition functions are evaluated within the MP2 ¼ FC=6-311 þ þG approach, using the anharmonic treatment included in the Gaussian 03 program package [54]. The agreement with the observed terms is considerably good, basically within possible experimental errors. The computational modelling of the atmospheric clustering will require further efforts before it can culminate with the simultaneous simulation of the temperature development of ðH2 OÞ2 , H2 O  N2 , and H2 O  O2 under the atmospheric conditions. Complexation with larger water clusters [55–62] can then represent another topic of a broader environmental interest.

Acknowledgements The reported research has been supported by a Grant-in-aid for NAREGI Nanoscience Project, and for Scientific Research on Priority Area (A) from the Ministry of Education, Culture, Sports, Science and Technology of Japan, by the Czech National Research Program ‘Information

ARTICLE IN PRESS Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

421

Society’ (Czech Acad. Sci. 1ET401110505), and by the National Science Council, Taiwan, Republic of China. Initial phase of the research line was supported by the Alexander von Humboldt-Stiftung and the Max-Planck-Institut fu¨r Chemie (Otto-Hahn-Institut). Last but not the least, the referee comments are highly appreciated, too.

References [1] Slanina Z. Contemporary theory of chemical isomerism. Dordrecht: D. Reidel Publishing Company; 1986. [2] Castleman AW, Wei S. Cluster reactions. Annu Rev Phys Chem 1994;45:685–719. [3] Franken KA, Dykstra CE. Quantum Monte Carlo vibrational analysis of the nitrogen-water complex. Chem Phys Lett 1994;220:161–6. [4] Castleman AW, Bowen KH. Clusters: structure, energetics, and dynamics of intermediate states of matter. J Phys Chem 1996;100:12911–44. [5] Bondybey VE, Beyer MK. How many molecules make a solution? Int Rev Phys Chem 2002;21:277–306. [6] Curtiss LA, Pople JA. Ab-initio calculation of vibrational force-field of water dimer. J Mol Spectrosc 1975;55:1–14. [7] Clementi E. Determination of liquid water structure, coordination numbers for ions and solvation for biological molecules. Berlin: Springer; 1976. [8] Moszynski R, Jeziorski B, Szalewicz K. Møller–Plesset expansion of the dispersion energy in the ring approximation. Int J Quant Chem 1993;45:409–31. [9] Mas EM, Szalewicz K, Bukowski R, Jeziorski B. Pair potential for water from symmetry-adapted perturbation theory. J Chem Phys 1997;107:4207–18. [10] van Duijneveldt-van de Rijdt JGCM, van Duijneveldt FB. Interaction optimized basis sets for correlated ab initio calculations on the water dimer. J Chem Phys 1999;111:3812–9. [11] Mas EM, Bukowski R, Szalewicz K, Groenenboom GC, Wormer PES, van der Avoird A. Water pair potential of near spectroscopic accuracy. I. Analysis of potential surface and virial coefficients. J Chem Phys 2000;113:6687–701. [12] Groenenboom GC, Wormer PES, van der Avoird A, Mas EM, Bukowski R, Szalewicz K. Water pair potential of near spectroscopic accuracy. II. Vibration-rotation-tunneling levels of the water dimer. J Chem Phys 2000;113:6702–15. [13] Smit MJ, Groenenboom GC, Wormer PES, van der Avoird A, Bukowski R, Szalewicz K. Vibrations, tunneling, and transition dipole moments in the water dimer. J Phys Chem A 2001;105:6212–25. [14] Curtiss LA, Melenders CA. A theoretical study of the solvation of O2 in water. J Phys Chem 1984;88:1325–9. [15] Curtiss LA, Eisgruber CL. A theoretical study of the interaction of N2 with water molecules—ðH2 OÞN 2N2 , N ¼ 128. J Chem Phys 1984;80:2022–8. [16] Bauer A, Godon M, Kheddar M, Hartmann JH, Bonamy J, Robert D. Temperature and perturber dependences of water-vapor 380 GHz-line broadening. JQSRT 1987;37:531–9. [17] Bonamy J, Robert D, Hartmann JM, Gonze ML, Saintloup R, Berger H. Line broadening, line shifting, and line coupling effects on N2 2H2 O stimulated Raman spectra. J Chem Phys 1989;91:5916–25. [18] Bonamy J, Bonamy L, Robert D, Gonze ML, Millot G, Lavorel B, et al. Rotational relaxation of nitrogen in ternary mixtures N2 2CO2 2H2 O—consequences in coherent anti-Stokes–Raman spectroscopy thermometry. J Chem Phys 1991;94:6584–9. [19] Seminario JM, Concha MC, Murray JS, Politzer P. Theoretical analyses of O2 =H2 O systems under normal and supercritical conditions. Chem Phys Lett 1994;222:25–32. [20] Slanina Z. A cluster model of the gas-phase water. Thermochim Acta 1987;114:273–80. [21] Slanina Z. Computational studies of water clusters: temperature, pressure and saturation effects on cluster fractions within the RRHO MCY-B/EPEN steam. J Mol Struct 1990;273:81–92. [22] Slanina Z. Pressure enhancement of gas-phase water-oligomer populations: a RRHO MCY-B/EPEN computational study. Z Phys D 1987;5:181–6.

ARTICLE IN PRESS 422

Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

[23] Slanina Z. Temperature enhancement of water-oligomer populations in saturated aqueous vapor. Thermochim Acta 1987;116:161–9. [24] Slanina Z. Computational studies of water clusters: temperature, pressure and saturation effects on cluster fractions within the RRHO MCY-B/EPEN steam. J Mol Struct 1990;273:81–92. [25] Slanina Z. A comparative study of the water-dimer gas-phase thermodynamics in the BJH- and MCYL-type flexible potentials. Chem Phys 1991;150:321–9. [26] Slanina Z. A useful test of temperature dependency of cluster populations under saturation conditions: the Ar2 case. Thermochim Acta 1992;209:1–6. [27] Dardi PS, Dahler JS. Equilibrium constants for the formation of van der Waals dimers—calculations for Ar-Ar and Mg-Mg. J Chem Phys 1990;93:3562–72. [28] Slanina Z. On temperature increase of the Mg2 mole fraction in saturated magnesium vapor. Thermochim Acta 1992;207:9–13. [29] Slanina Z, Zhao X, Kurita N, Gotoh H, Uhlı´ k F, Rudzin´ski JM, et al. Computing the relative gas-phase populations of C60 and C70 : beyond the traditional DH 0f ;298 scale. J Mol Graph Mod 2001;19:216–21. [30] JANAF thermochemical tables. Natl. Bur. Stand., Washington, DC, 1971, 1985. [31] Lide DR, editor. CRC handbook of chemistry and physics, 76th ed. Boca Raton: CRC Press; 1995. [32] Walrafen G, Fisher MR, Hokmabadi MS, Yang WH. Temperature dependence of low- and high-frequency Raman scattering from liquid water. J Chem Phys 1986;85:6970–1. [33] Fujita Y, Ikawa S. Effect of temperature on the hydrogen bond distribution in water as studied by infrared spectra. Chem Phys Lett 1989;159:184–8. [34] Vigasin AA. Water vapor continuous absorption in various mixtures: possible role of weakly bound complexes. JQSRT 2000;64:25–40. [35] Vaida V, Headrick JE. Physicochemical properties of hydrated complexes in the Earth’s atmosphere. J Phys Chem A 2000;104:5401–12. [36] Slanina Z, Uhlı´ k F, Saito AT, Osawa E. Computing molecular species of atmospheric significancy in Earth’s and other atmospheres. Phys Chem Earth C 2001;26:505–11. [37] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, et al. Gaussian 98, Revision A7. Pittsburgh, PA: Gaussian Inc.; 1998. [38] Boys SF, Bernardi F. The calculation of small molecular interactions by the difference of separate total energies. Some procedures with reduced errors. Mol Phys 1970;19:553–68. [39] van Duijneveldt FB, van Duijneveldt-van de Rijdt JGCM, van Lenthe JH. State-of-the-art in counterpoise theory. Chem Rev 1994;94:1873–85. [40] Curtiss LA, Jones C, Trucks GW, Raghavachari K, Pople JA. Gaussian-1 theory of molecular-energies for 2ndrow compounds. J Chem Phys 1990;93:2537–45. [41] Curtiss LA, Raghavachari K, Trucks GW, Pople JA. Gaussian-2 theory for molecular-energies of 1st-row and 2nd-row compounds. J Chem Phys 1991;94:7221–30. [42] Curtiss LA, Raghavachari K, Redfern PC, Rassolov V, Pople JA. Gaussian-3 (G3) theory for molecules containing first and second-row atoms. J Chem Phys 1998;109:7764–76. [43] Uhlı´ k F, Slanina Z, Hinchliffe A. An ab initio correlated study of structure, energetics and vibrations of ðO2 Þ2 . J Mol Struct (THEOCHEM) 1993;285:273–6. [44] Orlando JJ, Tyndall GS, Nickerson KE, Calvert JG. The temperature-dependence of collision-induced absorption by oxygen near 6 mm. J Geophys Res D 1991;96:20755–60. [45] Long CA, Ewing GE. The infrared spectrum of bound state oxygen dimers. Chem Phys Lett 1971;9: 225–9. [46] Long CA, Ewing GE. Spectroscopic investigation of van der Waals molecules. I. The infrared and visible spectra of ðO2 Þ2 . J Chem Phys 1973;58:4824–34. [47] Slanina Z. Dimerization of homonuclear biatomic molecules. Collect Czech Chem Commun 1974;39:228–35. [48] Uhlı´ k F, Slanina Z, Hinchliffe A. Gas-phase association of O2 : a computational thermodynamic study. Thermochim Acta 1993;228:9–14. [49] Slanina Z, Uhlı´ k F, De Almeida WB, Hinchliffe A. A computational thermodynamic evaluation of the altitude profiles of ðN2 Þ2 , N2 2O2 , and ðO2 Þ2 in the Earth’s atmosphere. Thermochim Acta 1994;231:55–60.

ARTICLE IN PRESS Z. Slanina et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 97 (2006) 415–423

423

[50] Aquilanti V, Ascenzi D, Bartolomei M, Cappelletti D, Cavalli S, Vitores MD, et al. Molecular beam scattering of aligned oxygen molecules. The nature of the bond in the O2 –O2 dimer. J Am Chem Soc 1999;121:10794–802. [51] Cacace F, de Petris G, Troiani A. Experimental detection of tetraoxygen. Angew Chem Int Ed Engl 2001;40:4062. [52] Cacace F, de Petris G, Troiani A. Experimental detection of tetranitrogen. Science 2002;295:480–1. [53] Cacace F. From N2 and O2 to N4 and O4 : pneumatic chemistry in the 21st century. Chem Eur J 2002;8:3839–47. [54] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, et al. Gaussian 03, Revision of C.01. Wallingford CT: Gaussian Inc.; 2004. [55] Khan A. Ab initio studies of ðH2 OÞ20 Hþ and ðH2 OÞ21 Hþ prismic, fused cubic and dodecahedral clusters: can H3 Oþ ion remain in cage cavity? Chem Phys Lett 2000;319:440–50. [56] Goldman N, Fellers RS, Leforestier C, Saykally RJ. Water dimers in the atmosphere: equilibrium constant for water dimerization from the VRT (ASP-W) potential surface. J Phys Chem A 2001;105:515–9. [57] Staikova M, Donaldson DJ. Ab initio investigation of water complexes of some atmospherically important acids: HONO, HNO3 and HO2NO2. Phys Chem Phys 2001;3:1999–2006. [58] Khan A. Theoretical studies of CO2 ðH2 OÞ20;24;28 clusters: stabilization of cages in hydrates by CO2 guest molecules. J Mol Struct (THEOCHEM) 2003;664–5:237–45. [59] Vaida V, Kjaergaard HG, Feierabend KJ. Hydrated complexes: relevance to atmospheric chemistry and climate. Int Rev Phys Chem 2003;22:203–19. [60] Dunn ME, Pokon EK, Shields GC. Thermodynamics of forming water clusters at various temperatures and pressures by Gaussian-2, Gaussian-3, complete basis set-QB3, and complete basis set-APNO model chemistries; implications for atmospheric chemistry. J Am Chem Soc 2004;126:2647–53. [61] Goldman N, Leforestier C, Saykally RJ. Water dimers in the atmosphere II: results from the VRT(ASP-W)III potential surface. J Phys Chem A 2004;108:787–94. [62] Daniel JS, Solomon S, Kjaergaard HG, Schofield DP. Atmospheric water vapor complexes and the continuum. Geophys Res Lett 2004;31:L06118.

Suggest Documents