Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

How reduced are nucleophilic gold complexes?

Isaac F. Leach ab, Diego Sorbelli cd, Leonardo Belpassi d, Paola Belanzoni cd, Remco W. A. Havenith abe and Johannes E. M. N. Klein *a
aMolecular Inorganic Chemistry, Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands. E-mail: j.e.m.n.klein@rug.nl
bZernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
cDepartment of Chemistry, Biology and Biotechnology, University of Perugia, Via Elce di Sotto, 8, 06123 Perugia, Italy
dCNR Institute of Chemical Science and Technologies “Giulio Natta” (CNR-SCITEC), Via Elce di Sotto, 8, 06123 Perugia, Italy
eGhent Quantum Chemistry Group, Department of Chemistry, Ghent University, 9000 Gent, Belgium

Received 31st May 2022 , Accepted 8th July 2022

First published on 25th July 2022


Abstract

Nucleophilic formal gold(-I) and gold(I) complexes are investigated via Intrinsic Bond Orbital analysis and Energy Decomposition Analysis, based on density functional theory calculations. The results indicate gold(0) centres engaging in electron-sharing bonding with Al- and B- based ligands. Multiconfigurational (CASSCF) calculations corroborate the findings, highlighting the gap between the electonic structures and the oxidation state formalism.


The remarkable two-coordinate gold complex 1 (Scheme 1), first synthesised by Aldridge and co-workers in 2019, is capable of inserting CO2 into its Au–Al bond, providing a nucleophilic source of gold.1 This “umpolung reactivity” is contrasted by gold's well studied electrophilicity,2 typical of gold(I) and gold(III) species.3–9 Indeed, a wide variety of transformations are now known to be catalysed by gold complexes, including carbon bond-formation10,11 and oxygen atom transfer,12–15 so potential applications of 1 include waste valorisation and CO2 conversion.16 The aluminyl ligand [Al(NON)], where NON is 4,5-bis(2,6-diisopropylanilido)-2,7-di-tert-butyl-9,9-dimethylxanthene, is known to act as a strong σ-donor, and a polarised Auδ–Alδ+ bond was reported,1 consistent with the (atomic) electronegativity (EN) difference: EN(Al, Au) = (1.61, 1.92) on the Allen scale.17,18 It follows that 1 is formally a gold(–I) species,19 which may at first seem unlikely but auride salts are stable in the solid state,20–23 in ionic liquids,24 and in liquid ammonia.25 Furthermore, theoretical work has proposed lithium aurides may be stable at very high (GPa) pressures, with the metal oxidation state (OS) tuneable down to gold(–III).26 Gold's unique ability to accept electron density is well-known, ultimately due to relativistic lowering of the 6s orbital,27 making it more accessible for bonding. More generally, gold is now known experimentally to exist in a wide range of formal OSs (up to +V).28–31
image file: d2dt01694j-s1.tif
Scheme 1 The nucleophilic complexes studied in this work (IPr = N,N′-bis(dipp)imidazole-2-ylidene, dipp = 2,6-iPr2C6H3), and the CO2-insertion of 1a.

In 2021, Suzuki et al. reported 2, a diarylboryl analogue of the aluminyl complex 1 (Scheme 1), which also exhibits nucleophilic reactivity.32 In particular, 2b can perform insertion reactions with methyl-substituted carbodiimide, forming a species just like the CO2-insertion product of 1a. We might expect this similar reactivity to be reflected in equivalent OS assignment of the gold centres in 1 and 2. Curiously though, the EN of gold (1.92) lies between that of aluminium (1.61) and boron (2.05) on the Allen scale (the EN scale recommended by IUPAC).19,331 and 2 are therefore formally gold(–I) and gold(I) complexes, respectively, a fact that seems hard to square with their chemical and structural parallels.

Typical two-coordinate gold complexes have the metal centre in the +I OS, with a formal 6s05d10 configuration at the metal and a classical ligand-polarised dative covalent bond (Fig. 1a) for each of the two ligands. However, the strongly σ-donating [Al(NON)] ligand in 1 and EN(Au)>EN(Al), may lead us to expect more electron sharing (Fig. 1b) or even inverted (Fig. 1c) metal–ligand bonding scenarios. An inverted σ-bond was proposed,1 based on the observed nucleophilic reactivity and a calculated negative partial charge on the gold centre, but subsequent computational analysis by some of us points towards an electron-sharing covalent Au–Al bond as the source of nucleophilicity.34–37 The inverted scenario implies nucleophilic action of the gold centre, as the bonding pair of electrons is more closely associated to the metal, able of performing e.g. nucleophilic reduction of CO2 as in a proposed mechanism.2 Herein, we report a detailed computational analysis of the nucleophilic gold complexes 1 and 2 (Scheme 1), focused on the OS of the metal centre.


image file: d2dt01694j-f1.tif
Fig. 1 Possible bonding scenarios, their effective gold configurations and corresponding metal oxidation states for the Au–X bond, where X is [Al(NON)] (1) or [B(o-tol)2] (2) (see Scheme 1).

At the (B97-3c)38 optimized geometry, we performed single point calculations with the PBE039/def2-TZVPP40 functional and basis set. Immediate insight into the nature of the Au–P and Al–Au σ-bonds of 1a can be gained via inspection of the Intrinsic Bonding Orbitals (IBOs)41,42 in Fig. 2A and B, respectively. The intrinsic atomic orbital (IAO)41 partial charge distribution of the Au–P bond, qσ-IBO(Au, P) = (0.19, 1.68), is typical for a dative covalent σ-interaction between gold and the L-type phosphine ligand.43,44 By contrast, the partial charge distribution of the Al–Au bond, qσ-IBO(Al, Au) = (1.17, 0.81), lies much closer towards the ideal electron-sharing bonding scenario, qσ-IBO(Au, X) = (1.00, 1.00). Similar results (Table S5) were obtained using the B3LYP45–48-D349(BJ)50/def2-SVP,40 PBEh-3c51 and GFN2-xTB52 methods, which performed well in our recent benchmark for efficient computation of geometries for gold complexes.53 This gold–aluminyl bond is significantly more electron-sharing than e.g., gold–alkyl bonds in analogous complexes (Table S6). The electron-sharing covalent bonding motif suggests an effective s1d10 configuration of gold(0) (Fig. 1b). The IBO analysis of the other X–Au–L bonds in 1 and 2 shows a maximum variation in qσ-IBO(Au) of only 6% (1avs.2b in Table 1), consistent with the reduced ancillary ligand and aluminyl effects reported in ref. 33 and 34. Since the IBO localization procedure conveniently condenses the Au–X σ-interactions into a single orbital, examination of the IAO partial charge distributions provides a robust interpretation of the relevant bond polarity. This is contrasted by inspection of calculated atomic charges, which sum over all interactions and are not uniquely defined. In fact, one can always choose a partial charge definition and calculate either classical or inverted bond polarities in complexes 1 and 2 (Table S2, see also ref. 32). This is just one example of the tenuous link between partial atomic charges and chemical OSs.54–56


image file: d2dt01694j-f2.tif
Fig. 2 IBOs of 1a, calculated with PBE0/def2-TZVPP//B97-3c in ORCA 5.0.2,57,58 rendered in IboView.41,42 IAO partial charge distributions indicated in parenthesis.
Table 1 IAO partial charge distributions of the IBOs corresponding to the Au–X and Au–L σ-bonds in X–Au–L (Scheme 1)
Species σ(IBO)2 σ(IBO)2
Au X Au L
Calculated with PBE0/def2-TZVPP//B97-3c.
1a (X = Al, L = P) 0.81 1.17 0.19 1.68
1b (X = Al, L = C) 0.75 1.22 0.20 1.68
2a (X = B, L = P) 0.75 1.18 0.21 1.67
2b (X = B, L = C) 0.69 1.23 0.23 1.67


Energy decomposition analysis (EDA)59–61 is another approach to interpret DFT calculations that can complement the application of IBO analysis to probe bonding, oxidation states and metal configurations. EDA quantifies the various interaction energies between user-defined fragments and was previously applied in ref. 32 and 33 using [LAu] and X neutral radical fragments. Here we instead choose to fragment the molecule into the metal centre Aun+ and a united ligand fragment Fn. The charge, n = (−1, 0, 1), is varied to prepare the gold fragment in the (s2d10, s1d10, s0d10) configurations. We can judge which set of fragment orbitals are most similar to the combined molecule's orbitals by identifying the fragments with the smallest orbital interaction energy (ΔEorb).62,63 The results show that the s1d10 configuration is most favourable for 1 and 2 (Table 2), consistent with a gold(0) centre. The s0d10 configuration is in fairly close competition, particularly for 2b.

Table 2 Orbital interaction energies (ΔEorb) of all species in kcal mol−1, with the Au fragment prepared in the [Xe] 4f145d106sn state with n = (0, 1, 2) corresponding to gold(I), gold(0) and gold(–I), respectively
Species s0d10 s1d10 s2d10
EDA calculated with PBE0-ZORA/TZ2P//B97-3c in AMS2020.64
1a (X = Al, L = P) −166 −123 −273
1b (X = Al, L = C) −173 −137 −294
2a (X = B, L = P) −185 −151 −388
2b (X = B, L = C) −196 −172 −370


To further probe the X–Au–L σ-bonding frameworks, CASSCF(4,4) calculations were performed at the optimized B97-3c geometries using Pipek-Mezey (PM) localization65 of the active space (see ESI for details). Significantly populated active orbitals localized well onto each of the three atomic centres (A–C in Fig. 3). The most delocalized of the three, A, is 77% centred on its most contributing atom, Al. The σ-antibonding orbital D is spread over the three centres but is minimally occupied (0.09e). These fractional occupations numbers are obtained as a weighted sum of the integer occupation numbers in each configuration. In order to chemically interpret the results, we perform a Valence Bond (VB)-like reading of the CASSCF wavefunction, in the spirit of the work by Angeli, Malrieu, and co-workers.66,67 To find the portion of the wavefunction that corresponds to gold(0), we simply sum the weights of the configurations with the gold-centred orbital (Fig. 3B) singly occupied, here yielding a 51% contribution from gold(0). Additional contributions from gold(–I) and gold(I) are 33% and 15%, respectively. Very similar gold(0) contributions were found for the other complexes (53–54%, Table S1). These results are consistent with both the IBOs and the EDA, which point towards the electron-sharing bonding of a s1d10 gold(0) centre (Fig. 1b). Validation of the VB-like reading of the CASSCF wavefunctions was obtained via VB-SCF calculations on two model complexes of 1a and 2a (see ESI for more details). Although we note some variation in the minor gold(I) and gold(–I) contributions (Table S7),68 the VB-SCF calculations similarly find the largest weight (>60%) for the gold(0) structure.


image file: d2dt01694j-f3.tif
Fig. 3 Active PM localized CASSCF(4,4) orbitals (A–D) of the Al–Au–P σ-bond in 1a, with their IAO partial charge distributions and occupation numbers shown in curved and square brackets, respectively. Calculated at the optimized B97-3c geometry in ORCA 5.0.2,57,58 rendered in IboView.41,42

A computational analysis of several nucleophilic gold complexes with aluminyl and boryl ligands is presented. We investigated the bonding scenario, using localized orbitals and energy decomposition analyses based on DFT calculations, and found that the electronic structures are consistent with 6s15d10 gold(0) centres participating in electron-sharing covalent bonding with the Al- and B- based ligands. A valence-bond-like interpretation of CASSCF(4,4) calculations supports this assignment, indicating >50% gold(0) character for all species investigated, with some additional contributions from gold(–I) and gold(I). These results are in line with previous theoretical investigations,34–37 and lead us to echo recommendations to avoid using atomic partial charges for OS assignment.54–56 While the gold(0) OS assignment conveniently summarizes the electronic structures in chemical terms, it lies in stark contrast to the formal gold(–I) and gold(I) assignments of 1 and 2, highlighting the pitfalls associated with determining OSs based on atomic negativity differences alone. Similar conclusions have been reached by Salvador and co-workers when investigating transition metal complexes with Fischer and Schrock carbenes.69–72 Future efforts to pin down the elusive OS concept may benefit from EN definitions capable of accounting for the molecular environment, such as the charge-dependent EN concept introduced by Sanderson73–75 and pioneered by Pritchard76–80 – which provided the basis for more recent treatments of EN within conceptual DFT.81,82

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We wish to thank the Center for Information Technology of the University of Groningen for their support and for providing access to the Peregrine high performance computing cluster. IFL thanks the Dutch Ministry of Education, Culture, and Science (OCW) for his PhD scholarship. JEMNK acknowledges funding from the Netherlands Organisation for Scientific Research (NWO START-UP grant). This research used resources of the Oak Ridge Leadership Computing Facility at the Oak Ridge National Laboratory, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC05-00OR22725. D.S., L.B. and P.B. thank the Ministero dell'Università e della Ricerca (MUR, project AMIS, through the program “Dipartimenti di Eccellenza −2018-2022′′) for financial support.

Notes and references

  1. J. Hicks, A. Mansikkamäki, P. Vasko, J. M. Goicoechea and S. Aldridge, Nat. Chem., 2019, 11, 237–241 CrossRef CAS.
  2. D. Bourissou, Nat. Chem., 2019, 11, 199–200 CrossRef CAS PubMed.
  3. L.-P. Liu and G. B. Hammond, Chem. Soc. Rev., 2012, 41, 3129–3139 RSC.
  4. H. Ohno, Isr. J. Chem., 2013, 53, 869–882 CrossRef CAS.
  5. H. G. Raubenheimer and H. Schmidbaur, J. Chem. Educ., 2014, 91, 2024–2036 CrossRef CAS.
  6. L. Zhang, Acc. Chem. Res., 2014, 47, 877–888 CrossRef CAS PubMed.
  7. R. Dorel and A. M. Echavarren, Chem. Rev., 2015, 115, 9028–9072 CrossRef CAS.
  8. D. Pflästerer and A. S. K. Hashmi, Chem. Soc. Rev., 2016, 45, 1331–1367 RSC.
  9. M. Mato, A. Franchino, C. GarcíA-Morales and A. M. Echavarren, Chem. Rev., 2021, 121, 8613–8684 CrossRef CAS.
  10. W. E. Brenzovich Jr., D. Benitez, A. D. Lackner, H. P. Shunatona, E. Tkatchouk, W. A. Goddard III and F. D. Toste, Angew. Chem., Int. Ed., 2010, 49, 5519–5522 CrossRef.
  11. G. Zhang, L. Cui, Y. Wang and L. Zhang, J. Am. Chem. Soc., 2010, 132, 1474–1475 CrossRef CAS.
  12. D. P. Zimin, D. V. Dar'In, V. Y. Kukushkin and A. Y. Dubovtsev, J. Org. Chem., 2021, 86, 1748–1757 CrossRef CAS PubMed.
  13. J. Xiao and X. Li, Angew. Chem., Int. Ed., 2011, 50, 7226–7236 CrossRef CAS.
  14. M. Rudolph and A. S. K. Hashmi, Chem. Commun., 2011, 47, 6536–6544 RSC.
  15. C. A. Witham, P. Mauleón, N. D. Shapiro, B. D. Sherry and F. D. Toste, J. Am. Chem. Soc., 2007, 129, 5838–5839 CrossRef CAS.
  16. C. McManus, J. Hicks, X. Cui, L. Zhao, G. Frenking, J. M. Goicoechea and S. Aldridge, Chem. Sci., 2021, 12, 13458–13468 RSC.
  17. L. C. Allen, J. Am. Chem. Soc., 1989, 111, 9003–9014 CrossRef CAS.
  18. J. B. Mann, T. L. Meek, E. T. Knight, J. F. Capitani and L. C. Allen, J. Am. Chem. Soc., 2000, 122, 5132–5137 CrossRef CAS.
  19. P. Karen, P. Mcardle and J. Takats, Pure Appl. Chem., 2016, 88, 831–839 CrossRef CAS.
  20. A. Pantelouris, G. Kueper, J. Hormes, C. Feldmann and M. Jansen, J. Am. Chem. Soc., 1995, 117, 11749–11753 CrossRef CAS.
  21. M. Jansen, Solid State Sci., 2005, 7, 1464–1474 CrossRef CAS.
  22. H. Nuss and M. Jansen, Angew. Chem., Int. Ed., 2006, 45, 4369–4371 CrossRef CAS PubMed.
  23. M. Jansen, Chem. Soc. Rev., 2008, 37, 1826–1835 RSC.
  24. J. Ghilane, O. Fontaine, P. Martin, J.-C. Lacroix and H. Randriamahazaka, Electrochem. Commun., 2008, 10, 1205–1209 CrossRef CAS.
  25. W. J. Peer and J. J. Lagowski, J. Am. Chem. Soc., 1978, 100, 6260–6261 CrossRef CAS.
  26. G. Yang, Y. Wang, F. Peng, A. Bergara and Y. Ma, J. Am. Chem. Soc., 2016, 138, 4046–4052 CrossRef CAS PubMed.
  27. P. Pyykkö, Angew. Chem., Int. Ed., 2004, 43, 4412–4456 CrossRef PubMed.
  28. K. Leary, A. Zalkin and N. Bartlett, J. Chem. Soc., Chem. Commun., 1973, 131–132 RSC.
  29. G. Kaindl, K. Leary and N. Bartlett, J. Chem. Phys., 1973, 59, 5050–5054 CrossRef CAS.
  30. K. Leary, Lawrence Berkeley Lab., [Rep.] LBL, 1972.
  31. M. C. Gimeno and A. Laguna, Gold Bull., 2003, 36, 83–92 CrossRef CAS.
  32. A. Suzuki, X. Guo, Z. Lin and M. Yamashita, Chem. Sci., 2021, 12, 917–928 RSC.
  33. Note that the ordering of the Al, Au and B electronegativites (and therefore the formal OS assignment) may differ when using alternative EN scales.
  34. D. Sorbelli, L. Belpassi and P. Belanzoni, J. Am. Chem. Soc., 2021, 143, 14433–14437 CrossRef CAS.
  35. D. Sorbelli, L. Belpassi and P. Belanzoni, Inorg. Chem., 2022, 61, 1704–1716 CrossRef CAS PubMed.
  36. D. Sorbelli, L. Belpassi and P. Belanzoni, Chem. Sci., 2022, 13, 4623–4634 RSC.
  37. D. Sorbelli, E. Rossi, R. W. A. Havenith, J. E. M. N. Klein, L. Belpassi and P. Belanzoni, Inorg. Chem., 2022, 61, 7327–7337 CrossRef CAS PubMed.
  38. J. G. Brandenburg, C. Bannwarth, A. Hansen and S. Grimme, J. Chem. Phys., 2018, 148, 064104 CrossRef PubMed.
  39. C. Adamo, M. Cossi and V. Barone, J. Mol. Struct.: THEOCHEM, 1999, 493, 145–157 CrossRef CAS.
  40. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  41. G. Knizia, J. Chem. Theory Comput., 2013, 9, 4834–4843 CrossRef CAS PubMed.
  42. G. Knizia and J. E. M. N. Klein, Angew. Chem., Int. Ed., 2015, 54, 5518–5522 CrossRef CAS PubMed.
  43. L. Nunes dos Santos Comprido, J. E. M. N. Klein, G. Knizia, J. Kästner and A. S. K. Hashmi, Chem. – Eur. J., 2017, 23, 10901–10905 CrossRef PubMed.
  44. L. Nunes dos Santos Comprido, J. E. M. N. Klein, G. Knizia, J. Kästner and A. S. K. Hashmi, Angew. Chem., Int. Ed., 2015, 54, 10336–10340 CrossRef CAS.
  45. P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627 CrossRef CAS.
  46. A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  47. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
  48. A. D. Becke, Phys. Rev. A, 1988, 38, 3098–3100 CrossRef CAS.
  49. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  50. S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
  51. S. Grimme, J. G. Brandenburg, C. Bannwarth and A. Hansen, J. Chem. Phys., 2015, 143, 054107 CrossRef.
  52. C. Bannwarth, S. Ehlert and S. Grimme, J. Chem. Theory Comput., 2019, 15, 1652–1671 CrossRef CAS PubMed.
  53. I. F. Leach, L. Belpassi, P. Belanzoni, R. W. A. Havenith and J. E. M. N. Klein, ChemPhysChem, 2021, 22, 1262–1268 CrossRef CAS PubMed.
  54. G. Aullón and S. Alvarez, Theor. Chem. Acc., 2009, 123, 67–73 Search PubMed.
  55. R. Resta, Nature, 2008, 453, 735–735 CrossRef CAS PubMed.
  56. H. Raebiger, S. Lany and A. Zunger, Nature, 2008, 453, 763–766 CrossRef CAS PubMed.
  57. F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2018, 8, e1327 Search PubMed.
  58. F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 73–78 CAS.
  59. K. Kitaura and K. Morokuma, Int. J. Quantum Chem., 1976, 10, 325–340 CrossRef CAS.
  60. T. Ziegler and A. Rauk, Inorg. Chem., 1979, 18, 1755–1759 CrossRef CAS.
  61. T. Ziegler and A. Rauk, Inorg. Chem., 1979, 18, 1558–1565 CrossRef CAS.
  62. F. M. Bickelhaupt and E. J. Baerends, in Rev. Comput. Chem, ed. K. B. Lipkowitz and D. B. Boyd, John Wiley & Sons, Ltd, 2000, pp. 1–86 Search PubMed.
  63. M. V. Hopffgarten and G. Frenking, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 43–62 Search PubMed.
  64. G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra, S. J. A. van Gisbergen, J. G. Snijders and T. Ziegler, J. Comput. Chem., 2001, 22, 931–967 CrossRef CAS.
  65. J. Pipek and P. G. Mezey, J. Chem. Phys., 1989, 90, 4916–4926 CrossRef CAS.
  66. J.-P. Malrieu, C. Angeli and R. Cimiraglia, J. Chem. Educ., 2008, 85, 150–158 CrossRef.
  67. J.-P. Malrieu, N. Guihéry, C. J. Calzado and C. Angeli, J. Comput. Chem., 2007, 28, 35–50 CrossRef CAS.
  68. For further discussion of the effect of (ligand) truncation, please refer to the ESI of ref. 34.
  69. M. Gimferrer, A. Aldossary, P. Salvador and M. Head-Gordon, J. Chem. Theory Comput., 2022, 18, 309–322 CrossRef CAS.
  70. M. Gimferrer, J. Van Der Mynsbrugge, A. T. Bell, P. Salvador and M. Head-Gordon, Inorg. Chem., 2020, 59, 15410–15420 CrossRef CAS PubMed.
  71. M. Gimferrer, G. Comas-Vilà and P. Salvador, Molecules, 2020, 25, 234 CrossRef CAS.
  72. V. Postils, C. Delgado–Alonso, J. M. Luis and P. Salvador, Angew. Chem., Int. Ed., 2018, 57, 10525–10529 CrossRef CAS PubMed.
  73. R. Sanderson, Chemical bonds and bonds energy, Academic Press, New York, 1976 Search PubMed.
  74. R. Sanderson, Polar covalence, Elsevier, 2012 Search PubMed.
  75. R. T. Sanderson, Science, 1951, 114, 670–672 CrossRef CAS.
  76. H. Pritchard and F. Sumner, Proc. R. Soc. London, Ser. A, 1956, 235, 136–143 CAS.
  77. H. Pritchard and H. Skinner, Chem. Rev., 1955, 55, 745–786 CrossRef CAS.
  78. H. Pritchard and F. Sumner, Proc. R. Soc. London, Ser. A, 1954, 226, 128–140 CAS.
  79. H. Skinner and H. Pritchard, Trans. Faraday Soc., 1953, 49, 1254–1262 RSC.
  80. H. Pritchard, Chem. Rev., 1953, 52, 529–563 CrossRef CAS.
  81. L. von Szentpály, Quantum Matter, 2015, 4, 47–55 CrossRef.
  82. L. von Szentpály, J. Phys. Chem. A, 2015, 119, 1715–1722 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2dt01694j

This journal is © The Royal Society of Chemistry 2023
Click here to see how this site uses Cookies. View our privacy policy here.