Active Thermochemical Tables: The Adiabatic Ionization Energy of

Sep 7, 2017 - JILA, National Institute of Standards and Technology and University of Colorado Boulder, Boulder, CO 80309, United States. ‡ Quantum T...
0 downloads 8 Views 387KB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Active Thermochemical Tables: The Adiabatic Ionization Energy of Hydrogen Peroxide P. Bryan Changala, Thanh Lam Nguyen, Joshua H. Baraban, G. Barney Ellison, John F. Stanton, David H. Bross, and Branko Ruscic J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.7b06221 • Publication Date (Web): 07 Sep 2017 Downloaded from http://pubs.acs.org on September 7, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry A is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Active Thermochemical Tables: The Adiabatic Ionization Energy of Hydrogen Peroxide P. Bryan Changala,† T. Lam Nguyen,‡ Joshua H. Baraban,¶ G. Barney Ellison,¶ John F. Stanton,∗,‡ David H. Bross,§ and Branko Ruscic∗,§,k JILA, National Institute of Standards and Technology and University of Colorado Boulder, Boulder, CO 80309 (USA), Quantum Theory Project, Departments of Chemistry and Physics, University of Florida, Gainesville, FL 32611 (USA), Department of Chemistry, University of Colorado, Boulder, CO 80302 (USA), Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, IL 60439 (USA), and Computation Institute, The University of Chicago, Chicago, IL 60637 (USA) E-mail: [email protected]; [email protected]



To whom correspondence should be addressed JILA, National Institute of Standards and Technology and University of Colorado Boulder, Boulder, CO 80309 (USA) ‡ Quantum Theory Project, Departments of Chemistry and Physics, University of Florida, Gainesville, FL 32611 (USA) ¶ Department of Chemistry, University of Colorado, Boulder, CO 80302 (USA) § Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, IL 60439 (USA) k Computation Institute, The University of Chicago, Chicago, IL 60637 (USA) †

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 31

Abstract The adiabatic ionization energy of hydrogen peroxide (HOOH) is investigated, both by means of theoretical calculations and theoretically assisted reanalysis of previous experimental data. Values obtained by three different approaches: 10.638±0.012 eV (purely theoretical determination), 10.649±0.005 eV (reanalysis of photoelectron spectrum) and 10.645±0.010 eV (reanalysis of photoionization spectrum) are in excellent mutual agreement. Further refinement of the latter two values to account for asymmetry of the rotational profile of the photoionization origin band leads to a reduction of 0.007±0.006 eV, which tends to bring them into even closer alignment with the purely theoretical value. Detailed analysis of this fundamental quantity by the Active Thermochemical Tables (ATcT) approach, using the present results and extant literature, gives a final estimate of 10.641±0.006 eV.

Introduction In the past decade, the accuracy of fundamental thermochemical quantities (such as bond energies and enthalpies of formation) for many molecules have increased by at least an order of magnitude relative to values found in 20th century data compilations. This dramatic advance is made possible by the concept of Active Thermochemical Tables (ATcT), 1,2 which represents a departure from the sequential manner in which thermochemical data have traditionally been determined and updated. ATcT relies on the concept of a thermochemical network (TN), in which the interdependence of thermochemical quantities that can exist between very different molecules (for example, the phenyl radical and ammonia

3

is made

explicit. By design, ATcT automatically accounts for these dependencies and thus guarantees an internally consistent set of data. All available information (in practice, both experimental and theoretical) relevant to the species involved is contained in the TN, and this provides a set of conditional thermochemical constraints that must be satisfied. The TN is then statistically analyzed, iteratively corrected for possible inconsistencies, and solved to 2

ACS Paragon Plus Environment

Page 3 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

obtain the thermochemical parameters, a process that involves detailed consideration of the uncertainties associated with the input data and which has been discussed elsewhere in the literature. 1,2,4–9 As indicated above, the use of thermochemical networks means that isolated experimental measurements impact not just the species that are being probed in the laboratory, but potentially many other similar as well as chemically unrelated molecules. This provides considerable motivation for highly accurate and precise determinations of fundamental quantities such as bond energies, adiabatic electron affinities and adiabatic ionization energies; any improvement in such information will propagate through the network and potentially lead to refined values and error bounds for many other species in the network. In the course of developing ATcT to its current state, effort has also been made to reinvestigate and reanalyze extant data. An excellent example of this is the dissociation energy of carbon monoxide, which ultimately plays a role in determining the enthalpy of formation for the carbon atom, clearly a quantity of fundamental importance. It transpires that the longquoted standard enthalpy of formation of the carbon atom (711.49±0.45 kJ mol−1 at 0 K; 716.68±0.45 kJ mol−1 at T=298 K from the CODATA compilation 10 ) relies solely on the bond dissociation energy of carbon monoxide (CO) (as well as the thermochemistry of CO - which was determined by considering experimental calorimetry, 11–14 the study of high-temperature equilibrium of CO, CO2 and O2 , 15,16 as well as the Boudouard reaction involving the equilibrium of CO, CO2 and graphite 15,17 - and that of the oxygen atom). The CODATA value for D0 (CO) is taken from a 1955 study of predissociation of CO (B1 Σ+ ) by Douglas and Møller, 18 which was originally intended to distinguish between the contending high (≈11 eV) or low (≈9 eV) estimates of the dissociation energy favored by, among others, Pauling 19 and Herzberg 20 (low values) as well as Kistiakowsky 21 and Brewer 22 (high values). Perhaps as a consequence of the motivation for that study, which was essentially qualitative in nature, the value was determined by a method of data analysis that was not particularly rigorous, using simple straight lines instead of the so called limiting curves of

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

dissociation. Nevertheless, this rather imprecisely determined value of ∆f H (C(g) ) attracted essentially no attention for fifty years. This, despite its considerable importance in computational chemistry, where thermochemical parameters are often determined by calculating the “total atomization energy” (the entire molecular binding energy with respect to constituent atoms in their ground state) and then determining ∆f H of the target species by removing the contributions from the unbound atoms, which are based on literature values. Focusing on this issue, the ATcT project has motivated highly accurate calculations of this quantity (D0 (CO)) 23,24 and a new photoionization study of CO in which the appearance energy of C+ was precisely determined . 25 In addition, the original data from Ref. 18 were analyzed using limiting curves of dissociation and the value of D0 (CO) was revised upwards by some 25 cm−1 (0.30 kJ mol−1 ). These recent estimates, along with other relevant data, have led to the current ATcT estimates of ∆f H (C(g) ) of 711.401±0.050 (0 K) and 716.886±0.050 (298.15 K) kJ mol−1 9 , which are consistent with, but nearly an order of magnitude more accurate than, those from the CODATA compilation. It is worthwhile to explicitly mention that the thermochemical networks in ATcT contain not just original experimental data, but also high-quality (and calibrated to some degree) ab initio data that come from computational thermochemistry. The carbon atom example is a case in point: the quoted enthalpies of formation above come from the solution of a thermochemical network, and of the ten most important determinations contributing to the provenance of ∆f H(C(g) ), four come from high level quantum chemical calculations. This combination of theoretical and experimental information is an important feature of ATcT; high-quality experiments and theory provide the raw materials that lead to accurate thermochemical parameters. Beyond this, and again exemplified by the example of C(g) , ATcT sometimes also uses published data which has been reanalyzed with appropriate, and often sophisticated, techniques. The present paper is an example of such a contribution, and focuses on the adiabatic ionization energy of hydrogen peroxide (HOOH). This quantity has been measured by one of us (BR) using photoionization mass spectrometry (10.631±0.007

4

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

eV), 26 a value that is consistent with a previous determination of 10.62±0.02 eV. 27 The implied ATcT value , 28 which is given by the difference in the ATcT ∆f H values of the ion and neutral at 0 K, is 10.637±0.006 eV. However, there is a recent study 29 that reports an adiabatic ionization energy of 10.685±0.005 eV, which is well outside the range of values consistent with ATcT. Given the stoichiometry of this molecule, this fundamental quantity – the ionization energy of HOOH – has potentially significant impact on ions of compounds involved in hydrogen combustion, and via their ionization energies, several key chemical species. The authors of Ref. 29 investigated two plausible assignments for the origin band in the photoelectron spectrum: their preferred assignment (10.685±0.005 eV), and one at a lower value of 10.649±0.005 eV. The latter overlaps the ATcT error range, although just barely so, but is outside the ranges given by Refs. 26 and. 27 However, the authors of Ref. 29 state that their determination is far from certain, and hope that their study will “stimulate ... interest in this problem to test the assignments presented and the derived AIE”. This work takes up that challenge. In the following, the ionization energy of HOOH is determined using three distinct approaches. First, we apply the HEAT approach from computational thermochemistry 24,30,31 to this problem. In addition, we carry out an elaborate temperature-dependent FranckCondon simulation of the photoelectron spectrum, which is a nontrivial undertaking since the electronic structure of the cation is complex, and is further complicated by the large amplitude motion associated with the torsional mode in this molecule. Moreover, the dihedral angle differs by roughly π/2 in the two electronic states studied, so it is imperative to treat the large amplitude motion appropriately. This calculation allows us to make a direct comparison with the spectrum of Ref., 29 as well as to analyze the photoionization efficiency curve produced in Ref. 26 It will be shown conclusively that the alternate assignment by the authors of Ref. 29 was indeed the correct one, and that the three determinations: direct calculation, and reanalysis of the experimental data from Refs. 26 and 29 are in mutual agreement. Finally, all of these results are added to the ATcT database and a new value for the adiabatic

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 31

ionization energy of HOOH is determined.

Results Quantum Chemical Determination The 1 A electronic ground state of HOOH and the ground 2 Bg state of its cation

32

were

treated with the composite thermochemical methods known as HEAT-345(Q) and HEAT456QP, 24 albeit with modifications motivated by both practical and necessary considerations. First, the equilibrium geometries of the two species (which have C2 and C2h symmetry, respectively), were optimized at the CCSD(T) 33 level of theory in the frozen core (fc) approximation using a triple-zeta atomic natural orbital basis designated as ANO1. 34,35 The standard HEAT protocol, which has been extensively benchmarked, 24,30,31 is based on geometries calculated at the CCSD(T) level of theory, but with the correlation-consistent cc-pVQZ basis 36 in an all-electron (ae) calculation. 30 The choice of the ANO1 basis is perhaps preferable for two reasons: the geometry optimization and - especially - the evaluation of the zero-point energy are significantly less expensive; and the ANO1 basis has been shown to provide excellent estimates of the positions of vibrational fundamentals 35 (and therefore, logically, the zero-point energy) when used in conjunction with CCSD(T). The effect on the energy of using the different geometry (fc-ANO1 v. ae-cc-pVQZ) is likely to be negligible, as was discussed in Ref. 24 At the fc-CCSD(T)/ANO1 geometries, the electronic energy contributions to HEAT were evaluated according to the standard protocols of the HEAT-345(Q) and HEAT-456QP approaches. That is, the total electronic energy is partitioned as follows:

Eelec = ESCF + ECCSD(T ) + ET −(T ) + EHLC + EREL + ESO + EDBOC ,

(1)

where ESCF and ECCSD(T ) are obtained from ae calculations using the aug-cc-pCVXZ ba-

6

ACS Paragon Plus Environment

Page 7 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

sis sets (X=T,Q and 5 or Q,5 and 6, for the two methods, respectively) and extrapolating both results to the basis set limit according to the prescriptions given in Ref. 30 ET −(T ) and EHLC (high-level correction) are intended to remedy remaining deficiencies of the CCSD(T) treatment of correlation; the first accounts for the difference between a full treatment of triple excitations (CCSDT 37–39 ) and the perturbative treatment of CCSD(T), while EHLC endeavors to account for remaining “high-level” correlation effects. The triples correction involves basis set extrapolation from fc calculations done with the cc-pVTZ and cc-pVQZ basis sets, and the final correction is obtained (again in the fc approximation) with CCSDT(Q) 40 (HEAT-345(Q)) or CCSDTQP (HEAT-456QP) 41 with the economical cc-pVDZ basis. The remaining corrections for relativistic, spin-orbit and the diagonal Born-Oppenheimer term (the adiabatic correction) are calculated according to the prescriptions given in Refs. 30 and. 31 The spin-orbit term does not enter into the current calculations; HEAT treats these interactions only to leading order, which vanishes for non-degenerate electronic states. The remaining contribution to the HEAT energy is the zero-point vibrational correction (EZP E ), viz. EHEAT = Eelec + EZP E ,

(2)

a matter that is deserving of special attention. Most of the molecules in the standard HEAT calibration set (in which the accuracy obtained for enthalpies of formation with HEAT345(Q) and HEAT-456QP is better than 0.5 kJ mol−1 ) are relatively rigid systems, and the treatment of anharmonicity provided by second order vibrational perturbation theory (VPT2) 42 is both appropriate and adequate. However, in HOOH, there is a large amplitude torsional motion in the ground state with a relatively small barrier to interconversion of the two enantiomers and a sizable (11 cm−1 ) tunneling splitting of the ground vibrational state; the propriety of VPT2 can certainly be questioned here. Thus, the determination of the ZPE used in the original benchmark calculations (done with ae-CCSD(T)/cc-pVQZ) will be abandoned here. For HOOH, we have chosen to use a variational zero-point energy of 5726 cm−1 obtained with the fc-CCSD(T)/cc-pVQZ potential energy surface of Koput et al., 43 a 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

result that has been calculated with the nitrogen program 44 as part of the present research. Although this value compares favorably with the fc-CCSD(T)/ANO1 VPT2 determination (which of course neglects the tunneling splitting) that gives 5711 cm−1 , it behooves us to use the variational result that more properly treats the large amplitude motion. For the cation, which is a considerably less problematic case in terms of nuclear motion, other issues arise due to artifactual symmetry breaking of the wavefunction. While the use of analytic second derivatives 45 provides a means to obtain the harmonic zero-point energy without interference from symmetry breaking, the evaluation of cubic and quartic force constants necessary for VPT2 necessarily requires calculations to be done at lower-symmetry geometries. Moreover, unphysical harmonic frequencies are obtained for this species, for reasons that are wellknown and discussed, for example, in Refs. 46 and. 47 In order to obtain a suitable potential energy surface for calculations, we used EOMIP-CCSD(T)(a), 48 which is free of artifacts from symmetry breaking. 47 At the fc-EOMIP-CCSD/ANO0 level used to construct the potential energy surface used for the simulation described in the next subsection, variational and VPT2 calculations gave essentially identical zero-point energies (within 3 cm−1 ), so the strategy taken here was to evaluate the ZPE of the ion using VPT2. Accordingly, the final cation ZPE of 5844 cm−1 was obtained at the fc-EOMIP-CCSD(T)(a)/cc-pVQZ level of theory with VPT2, the basis set chosen to be consistent with the surface used in the variational calculation for the neutral species. Despite the issues described above, we are confident that the differential zero-point energy between the neutral and the cation (118 cm−1 ) is in error by no more than 40 cm−1 (0.005 eV). The HEAT-345(Q) and HEAT-456QP determinations of the ionization energy are fully documented in Table 1. Given the fairly significant spin contamination of the cation unrestricted wavefunction, it was deemed worthwhile to carry out calculations with both unrestricted Hartree-Fock (UHF, which is the standard approach for HEAT-based treatments of open-shell systems) and restricted open-shell Hartree-Fock (ROHF) reference functions. While significant differences are indeed seen at the SCF level, the characteristic near-

8

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

invariance of coupled-cluster energies obtained with methods that include single excitations is apparent: the corresponding ionization energies differ by