research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Solid energy calibration standards for P K-edge XANES: electronic structure analysis of PPh4Br

CROSSMARK_Color_square_no_text.svg

aThe University of Iowa, Department of Chemistry, E331 Chemistry Building, Iowa City, IA 52242, USA, and bDepartment of Chemistry, Colgate University, 13 Oak Drive, Hamilton, NY 13346, USA
*Correspondence e-mail: jkeith@colgate.edu, scott-daly@uiowa.edu

Edited by R. W. Strange, University of Essex, UK (Received 10 August 2017; accepted 8 January 2018; online 20 February 2018)

P K-edge X-ray absorption near-edge structure (XANES) spectroscopy is a powerful method for analyzing the electronic structure of organic and inorganic phosphorus compounds. Like all XANES experiments, P K-edge XANES requires well defined and readily accessible calibration standards for energy referencing so that spectra collected at different beamlines or under different conditions can be compared. This is especially true for ligand K-edge X-ray absorption spectroscopy, which has well established energy calibration standards for Cl (Cs2CuCl4) and S (Na2S2O3·5H2O), but not neighboring P. This paper presents a review of common P K-edge XANES energy calibration standards and analysis of PPh4Br as a potential alternative. The P K-edge XANES region of commercially available PPh4Br revealed a single, highly resolved pre-edge feature with a maximum at 2146.96 eV. PPh4Br also showed no evidence of photodecomposition when repeatedly scanned over the course of several days. In contrast, we found that PPh3 rapidly decomposes under identical conditions. Density functional theory calculations performed on PPh3 and PPh4+ revealed large differences in the molecular orbital energies that were ascribed to differences in the phospho­rus oxidation state (III versus V) and molecular charge (neutral versus +1). Time-dependent density functional theory calculations corroborated the experimental data and allowed the spectral features to be assigned. The first pre-edge feature in the P K-edge XANES spectrum of PPh4Br was assigned to P 1s → P-C π* transitions, whereas those at higher energy were P 1s → P-C σ*. Overall, the analysis suggests that PPh4Br is an excellent alternative to other solid energy calibration standards commonly used in P K-edge XANES experiments.

1. Introduction

P K-edge X-ray absorption near-edge structure (XANES) spectroscopy is a powerful method for analyzing organic and inorganic compounds containing phosphorus. Electronic transitions at the P K-edge occur in the `tender' X-ray region (P 1s binding energy = 2145.5 eV), and the resulting spectra have sufficient resolution for quantitative electronic structure comparisons and utility in diverse applications (Northrup et al., 2016[Northrup, P., Leri, A. & Tappero, R. (2016). Protein Pept. Lett. 23, 300-308.]). For example, P K-edge XANES has been used to identify phospho­rus-containing chemicals in soil and other complex mixtures, especially those encountered in agricultural, environmental and tribological-related samples (Ingall et al., 2011[Ingall, E. D., Brandes, J. A., Diaz, J. M., de Jonge, M. D., Paterson, D., McNulty, I., Elliott, W. C. & Northrup, P. (2011). J. Synchrotron Rad. 18, 189-197.]; Yoshida et al., 2016[Yoshida, M., Mineo, T., Mitsutomi, Y., Yamamoto, F., Kurosu, H., Takakusagi, S., Asakura, K. & Kondoh, H. (2016). Chem. Lett. 45, 277-279.]; Morel et al., 2015[Morel, F. L., Pin, S., Huthwelker, T., Ranocchiari, M. & van Bokhoven, J. A. (2015). Phys. Chem. Chem. Phys. 17, 3326-3331.]; Tan et al., 2014[Tan, Y., Sun, D., Yu, H., Yang, B., Gong, Y., Yan, S., Chen, Z., Cai, Q. & Wu, Z. (2014). CrystEngComm, 16, 9657-9668.]; Kasrai et al., 1999[Kasrai, M., Vasiga, M., Suominen Fuller, M., Bancroft, G. M. & Fyfe, K. (1999). J. Synchrotron Rad. 6, 719-721.]; Zhang et al., 2005[Zhang, Z., Yamaguchi, E. S., Kasrai, M. & Bancroft, G. M. (2005). Tribol. Lett. 19, 211-220.]; Pereira et al., 2006[Pereira, G., Lachenwitzer, A., Munoz-Paniagua, D., Kasrai, M., Norton, P. R., Abrecht, M. & Gilbert, P. U. P. A. (2006). Tribol. Lett. 23, 109-119.], 2007[Pereira, G., Munoz-Paniagua, D., Lachenwitzer, A., Kasrai, M., Norton, P. R., Capehart, T. W., Perry, T. A. & Cheng, Y.-T. (2007). Wear, 262, 461-470.]; Sharma et al., 2016[Sharma, V., Doerr, N. & Aswath, P. B. (2016). RSC Adv. 6, 22341-22356.]). It has also been used to quantify relative differences in covalent metal–phospho­rus bonding and electronic structure in coordination compounds containing phospho­rus ligands (Adhikari et al., 2008[Adhikari, D., Mossin, S., Basuli, F., Huffman, J. C., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2008). J. Am. Chem. Soc. 130, 3676-3682.]; Harkins et al., 2008[Harkins, S. B., Mankad, N. P., Miller, A. J. M., Szilagyi, R. K. & Peters, J. C. (2008). J. Am. Chem. Soc. 130, 3478-3485.]; Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]; Mossin et al., 2012[Mossin, S., Tran, B. L., Adhikari, D., Pink, M., Heinemann, F. W., Sutter, J., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2012). J. Am. Chem. Soc. 134, 13651-13661.]; Shearer et al., 2012[Shearer, J., Callan, P. E., Masitas, C. A. & Grapperhaus, C. A. (2012). Inorg. Chem. 51, 6032-6045.]; Soma et al., 2016[Soma, S., Van Stappen, C., Kiss, M., Szilagyi, R. K., Lehnert, N. & Fujisawa, K. (2016). J. Biol. Inorg. Chem. 21, 757-775.]; Hu et al., 2015[Hu, Y., Maclennan, A. & Sham, T. K. (2015). J. Lumin. 166, 143-147.]; Donahue et al., 2015[Donahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646-5659.]; Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]). This latter type of metal–ligand bonding analysis is more generally referred to as ligand K-edge X-ray absorption spectroscopy (XAS) (Solomon et al., 2005[Solomon, E. I., Hedman, B., Hodgson, K. O., Dey, A. & Szilagyi, R. K. (2005). Coord. Chem. Rev. 249, 97-129.]; Glaser et al., 2000[Glaser, T., Hedman, B., Hodgson, K. O. & Solomon, E. I. (2000). Acc. Chem. Res. 33, 859-868.]; Queen et al., 2013[Queen, M. S., Towey, B. D., Murray, K. A., Veldkamp, B. S., Byker, H. J. & Szilagyi, R. K. (2013). Coord. Chem. Rev. 257, 564-578.]).

As with all XANES measurements, ligand K-edge XAS requires a convenient and well defined energy calibration standard so that accurate comparisons can be made between different data sets collected at different synchrotron facilities. Photodecomposition-resistant energy standards with intense, highly resolved pre-edge features have long been established for referencing ligand K-edge XAS data with the third row elements chlorine (Cs2CuCl4) and sulfur (Na2S2O3·5H2O) (Solomon et al., 2005[Solomon, E. I., Hedman, B., Hodgson, K. O., Dey, A. & Szilagyi, R. K. (2005). Coord. Chem. Rev. 249, 97-129.]). At the beginning of our electronic structure investigations of metal-phosphine complexes (Donahue et al., 2015[Donahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646-5659.]), we sought a well defined solid energy standard for K-edge XANES measurements with phosphorus (which neighbors sulfur and chlorine) and found, unexpectedly, that a single reference had yet to be established for ligand K-edge XAS studies. Numerous phosphorus compounds have been used as energy standards, and much of this data comes from researchers using P K-edge XANES in applications other than ligand K-edge XAS.

Here we provide a brief, but critical, review of commonly used P K-edge energy standards, and report the P K-edge XANES spectrum and density functional theory (DFT) calculations of PPh4Br. This compound has worked extremely well as a P K-edge energy calibration standard for ligand K-edge XAS studies (Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]), and offers several advantages over other calibration standards.

2. Experimental

Na4P2O7, OPPh3 and PPh4Br were purchased from commercial vendors in their highest available purity and used as received. PPh3 was recrystallized from distilled EtOH and analyzed by 31P NMR spectroscopy before use to ensure that no OPPh3 impurity was present.

2.1. XAS data collection and analysis

P K-edge XAS data were collected at beamline 14-3 at the Stanford Synchrotron Radiation Lightsource under dedicated operating conditions of 3.0 GeV and 500 mA using beamline optics and equipment described previously (Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]). Beamline 14-3 has a bending-magnet source and a water-cooled double Si(111) crystal monochromator. The sample chamber was maintained under a He atmosphere and sample fluorescence was measured using a PIPS or Vortex detector with the sample placed at a 45° angle with respect to the incident beam. The beam spot size was ≤3 mm2 and the flux was ca. 2 × 1010 photons s−1. No filters were used.

Samples were finely ground using a Wig-L-Bug grinder for two minutes and thinly painted on Kapton tape with a size 4 Blick Scholastic Wonder White Flat synthetic brush (or equivalent) to minimize self-absorption effects. This was confirmed by analyzing progressively thicker layers to identify the onset of observable self-absorption effects in the spectra. P K-edge XAS data were collected over three regions with 1.0 eV steps and 1.0 s dwell times between 2106 and 2136 eV (pre-edge), 0.08 eV steps and 1–2 s dwell times between 2136 and 2165 eV (edge), and 1.5 eV steps and 1 s dwell times between 2165 and 2380 eV (post-edge). Data were collected in triplicate scans over several experimental runs. P K-edge XAS data were calibrated to the white line of Na4P2O7 at 2152.40 eV (Fig. 1[link]). Calibration scans were collected before and after each triplicate data set.

[Figure 1]
Figure 1
Comparison of P K-edge XANES spectra of common energy calibration standards. First and second derivative traces depicted with red and blue dashed lines, respectively, are offset on the y-axis by +2 or +4 intensity units. P K-edge XANES data for hy­droxy­apatite, Ca5(PO4)3(OH), were obtained from the ID21 Phospho­rous XANES Spectra Database and are reported by Kim et al. (2015)[Kim, B., Gautier, M., Rivard, C., Sanglar, C., Michel, P. & Gourdon, R. (2015). Environ. Sci. Technol. 49, 4903-4910.].

Calibrations, background subtractions, normalizations and averaging were performed using the Athena program in the IFFEFIT XAS software package (Ravel & Newville, 2005[Ravel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537-541.]). All scans were ratioed against the incident radiation (I0), which was measured using an ionization chamber located prior to the sample chamber. Subsequent pre-edge and post-edge fits for background subtractions and normalizations were performed as previously described (Donahue et al., 2015[Donahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646-5659.]; Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]). The background was removed from each spectrum by fitting a first-order polynomial to the pre-edge region. The data were normalized by fitting the post-edge region with a first- or second-order polynomial. The step function intensity for data normalizations was set to 1.0 at 2165 eV. Curve fits were performed using IGOR Pro 6.3 using a modified version of EDGEFIT, as described previously (Donahue et al., 2015[Donahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646-5659.]; Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]).

2.2. DFT and TDDFT calculations

DFT calculations were performed as described previously using the Gaussian09 computational chemistry suite (Frisch et al., 2009[Frisch, M. J. T. G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, N. J., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). Gaussian 09, Revision D.01. Gaussian Inc., Pittsburgh, Pennsylvania, USA.]; Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]). To summarize, PPh3 and PPh4+ were evaluated at the dispersion-corrected hybrid density functional theory level known as B3LYP-d3 with the three-parameter exchange functional of Becke (Becke, 1993[Becke, A. D. (1993). J. Chem. Phys. 98, 5648-5652.]), the correlation functional of Lee, Yang and Parr (Lee et al., 1988[Lee, C., Yang, W. & Parr, R. G. (1988). Phys. Rev. B, 37, 785-789.]), and the addition of empirical dispersion with Grimme's D3 dispersion corrections (Grimme et al., 2010[Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.]). All atoms were modeled with the split-valence double-ζ plus polarization basis set 6-31G(d,p) (Hehre et al., 1972[Hehre, W. J., Ditchfield, R. & Pople, J. A. (1972). J. Chem. Phys. 56, 2257-2261.]; Hariharan & Pople, 1973[Hariharan, P. C. & Pople, J. A. (1973). Theor. Chim. Acta, 28, 213-222.]), and both structures were fully optimized in the gas phase. Frequency calculations were performed to ensure the absence of any imaginary vibrational modes. Wavefunction stability calculations were performed on all species to ensure that the resulting determinant was in fact a local minimum (Bauernschmitt & Ahlrichs, 1996[Bauernschmitt, R. & Ahlrichs, R. (1996). J. Chem. Phys. 104, 9047-9052.]). Time-dependent density functional theory (TDDFT) calculations were conducted to approximate excitation energies and amplitudes (oscillator strengths) for specific excited states, and to simulate the P K-edge XAS spectra. Energy shifts of +50.1 (PPh3) and +49.6 eV (PPh4+) were applied to the simulated transitions and TDDFT spectra to align them with the experimental data. As described previously, this difference in absolute energy arises due to the nonlocal nature of the calculated core hole, the inability to model the relaxation of the core hole, and inaccuracies of the functional. As shown here, once the error in absolute energy has been accounted for by applying a uniform energy shift, the calculations reproduce experimental ligand K-edge XAS data with very high accuracy for a wide-variety of complexes, including those containing transition metals, lanthanides and actinides (Kozimor et al., 2008[Kozimor, S. A., Yang, P., Batista, E. R., Boland, K. S., Burns, C. J., Christensen, C. N., Clark, D. L., Conradson, S. D., Hay, P. J., Lezama, J. S., Martin, R. L., Schwarz, D. E., Wilkerson, M. P. & Wolfsberg, L. E. (2008). Inorg. Chem. 47, 5365-5371.], 2009[Kozimor, S. A., Yang, P., Batista, E. R., Boland, K. S., Burns, C. J., Clark, D. L., Conradson, S. D., Martin, R. L., Wilkerson, M. P. & Wolfsberg, L. E. (2009). J. Am. Chem. Soc. 131, 12125-12136.]; Daly et al., 2012[Daly, S. R., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Kozimor, S. A. & Martin, R. L. (2012). J. Am. Chem. Soc. 134, 14408-14422.]; Minasian et al., 2012[Minasian, S. G., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Conradson, S. D., Kozimor, S. A., Martin, R. L., Schwarz, D. E., Shuh, D. K., Wagner, G. L., Wilkerson, M. P., Wolfsberg, L. E. & Yang, P. (2012). J. Am. Chem. Soc. 134, 5586-5597.], 2014[Minasian, S. G., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Kozimor, S. A., Martin, R. L., Shuh, D. K. & Tyliszczak, T. (2014). Chem. Sci. 5, 351-359.]; Spencer et al., 2013[Spencer, L. P., Yang, P., Minasian, S. G., Jilek, R. E., Batista, E. R., Boland, K. S., Boncella, J. M., Conradson, S. D., Clark, D. L., Hayton, T. W., Kozimor, S. A., Martin, R. L., MacInnes, M. M., Olson, A. C., Scott, B. L., Shuh, D. K. & Wilkerson, M. P. (2013). J. Am. Chem. Soc. 135, 2279-2290.]; Löble et al., 2015[Löble, M. W., Keith, J. M., Altman, A. B., Stieber, S. C. E., Batista, E. R., Boland, K. S., Conradson, S. D., Clark, D. L., Lezama Pacheco, J., Kozimor, S. A., Martin, R. L., Minasian, S. G., Olson, A. C., Scott, B. L., Shuh, D. K., Tyliszczak, T., Wilkerson, M. P. & Zehnder, R. A. (2015). J. Am. Chem. Soc. 137, 2506-2523.]; Donahue et al., 2014[Donahue, C. M., Lezama Pacheco, J. S., Keith, J. M. & Daly, S. R. (2014). Dalton Trans. 43, 9189-9201.], 2015[Donahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646-5659.]; Olson et al., 2014[Olson, A. C., Keith, J. M., Batista, E. R., Boland, K. S., Daly, S. R., Kozimor, S. A., MacInnes, M. M., Martin, R. L. & Scott, B. L. (2014). Dalton Trans. 43, 17283-17295.]; Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]).

3. Results and discussion

3.1. Review of common P K-edge XANES energy calibration standards

Numerous compounds have been used for P K-edge XANES energy referencing including Na4P2O7 (sodium pyrophosphate) (Blanchard et al., 2008[Blanchard, P. E. R., Grosvenor, A. P., Cavell, R. G. & Mar, A. (2008). Chem. Mater. 20, 7081-7088.]; Grosvenor et al., 2008[Grosvenor, A. P., Cavell, R. G. & Mar, A. (2008). J. Solid State Chem. 181, 2549-2558.]; Kruse & Leinweber, 2008[Kruse, J. & Leinweber, P. (2008). J. Plant Nutr. Soil Sci. 171, 613-620.]; Kruse et al., 2010[Kruse, J., Negassa, W., Appathurai, N., Zuin, L. & Leinweber, P. (2010). J. Environ. Qual. 39, 2179-2184.]), red phosphorus (Rafiuddin et al., 2014[Rafiuddin, M. R., Mueller, E. & Grosvenor, A. P. (2014). J. Phys. Chem. C, 118, 18000-18009.]; Prietzel et al., 2016[Prietzel, J., Harrington, G., Häusler, W., Heister, K., Werner, F. & Klysubun, W. (2016). J. Synchrotron Rad. 23, 532-544.], 2013[Prietzel, J., Dümig, A., Wu, Y., Zhou, J. & Klysubun, W. (2013). Geochim. Cosmochim. Acta, 108, 154-171.]; Rafiuddin & Grosvenor, 2016[Rafiuddin, M. R. & Grosvenor, A. P. (2016). Inorg. Chem. 55, 9685-9695.]), AlPO4·(H2O)n (aluminium phosphate; n = 0, 1.5 or 2) (Liu et al., 2015[Liu, J., Hu, Y., Yang, J., Abdi, D. & Cade-Menun, B. J. (2015). Environ. Sci. Technol. 49, 168-176.], 2013[Liu, J., Yang, J., Cade-Menun, B. J., Liang, X., Hu, Y., Liu, C. W., Zhao, Y., Li, L. & Shi, J. (2013). J. Environ. Qual. 42, 1763-1770.]; Kraal et al., 2015[Kraal, P., Bostick, B. C., Behrends, T., Reichart, G.-J. & Slomp, C. P. (2015). Mar. Chem. 168, 1-8.]; Eveborn et al., 2009[Eveborn, D., Gustafsson, J. P., Hesterberg, D. & Hillier, S. (2009). Environ. Sci. Technol. 43, 6515-6521.]; Baumann et al., 2017[Baumann, K., Glaser, K., Mutz, J.-E., Karsten, U., MacLennan, A., Hu, Y., Michalik, D., Kruse, J., Eckhardt, K.-U., Schall, P. & Leinweber, P. (2017). Soil Biol. Biochem. 109, 156-166.]; Huang & Tang, 2016[Huang, R. & Tang, Y. (2016). Water Res. 100, 439-447.]), Ca5(PO4)3(OH) (hy­droxy­apatite) (Takamoto & Hashimoto, 2014[Takamoto, A. & Hashimoto, Y. (2014). Chem. Lett. 43, 1696-1697.]; Hashimoto et al., 2014[Hashimoto, Y., Takamoto, A., Kikkawa, R., Murakami, K. & Yamaguchi, N. (2014). Environ. Sci. Technol. 48, 5486-5492.]), Zn3(PO4)2 (Kasrai et al., 1999[Kasrai, M., Vasiga, M., Suominen Fuller, M., Bancroft, G. M. & Fyfe, K. (1999). J. Synchrotron Rad. 6, 719-721.]; Sharma et al., 2016[Sharma, V., Doerr, N. & Aswath, P. B. (2016). RSC Adv. 6, 22341-22356.]) and OPPh3 (tri­phenyl­phosphine oxide) (Adhikari et al., 2008[Adhikari, D., Mossin, S., Basuli, F., Huffman, J. C., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2008). J. Am. Chem. Soc. 130, 3676-3682.]; Harkins et al., 2008[Harkins, S. B., Mankad, N. P., Miller, A. J. M., Szilagyi, R. K. & Peters, J. C. (2008). J. Am. Chem. Soc. 130, 3478-3485.]; Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]; Mossin et al., 2012[Mossin, S., Tran, B. L., Adhikari, D., Pink, M., Heinemann, F. W., Sutter, J., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2012). J. Am. Chem. Soc. 134, 13651-13661.]; Shearer et al., 2012[Shearer, J., Callan, P. E., Masitas, C. A. & Grapperhaus, C. A. (2012). Inorg. Chem. 51, 6032-6045.]; Soma et al., 2016[Soma, S., Van Stappen, C., Kiss, M., Szilagyi, R. K., Lehnert, N. & Fujisawa, K. (2016). J. Biol. Inorg. Chem. 21, 757-775.]). At the beginning of our P K-edge XAS studies, anhydrous Na4P2O7 was selected because it was readily available and had a reference value reported to 0.01 eV precision. The most intense peak (white line) located at 2152.40 eV in the P K-edge XANES spectrum is used for energy referencing (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]; Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). Unfortunately, we discovered that the white line was not well resolved from a lower-energy feature with similar intensity (Fig. 1[link]) and often required longer calibration dwell times to ensure sufficient signal-to-noise for accurate energy referencing via analysis of the first derivative trace. Our studies also prompted questions as to how the position of the features changes in response to absorption of water. Na4P2O7 is hygroscopic and can absorb up to as many as ten H2O molecules when exposed to ambient humidity (it is available commercially as an anhydrous powder or decahydrate). It is not clear if variations in Na4P2O7 hydration alter the energy and intensities of its P K-edge XANES features, but related hydrogen-bonding interactions are known to cause changes in XANES peak energies and intensities in other compounds (Dey et al., 2005[Dey, A., Okamura, T., Ueyama, N., Hedman, B., Hodgson, K. O. & Solomon, E. I. (2005). J. Am. Chem. Soc. 127, 12046-12053.]).

We next considered Ca5(PO4)3(OH) (hy­droxy­apatite) and red phosphorus. Red phosphorus was quickly ruled out because of its reactivity; it slowly oxidizes in air, which has been monitored by in situ P K-edge XANES measurements as a function of time and temperature (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]). In contrast, Ca5(PO4)3(OH) is a highly inert bioceramic found in bones and teeth. It has been studied extensively using P K-edge XANES and has an intense white line ideal for energy referencing. However, the analysis of 17 reported spectra by Oxmann showed subtle variations in the white line energy depending on the Ca5(PO4)3(OH) preparation method and commercial vendor (Oxmann, 2014[Oxmann, J. F. (2014). Biogeosciences, 11, 2169-2183.]). It is well known that Ca5(PO4)3(OH) morphology and Ca/P stoichiometry is sensitive to the preparation method used (Ferraz et al., 2004[Ferraz, M. P., Monteiro, F. J. & Manuel, C. M. (2004). J. Appl. Biomater. Biomech. 2, 74-80.]; Raynaud et al., 2002[Raynaud, S., Champion, E., Bernache-Assollant, D. & Thomas, P. (2002). Biomaterials, 23, 1065-1072.]), which is likely to account for the observed spectral differences (Nakahira et al., 2011[Nakahira, A., Nakata, K., Numako, C., Murata, H. & Matsunaga, K. (2011). Mater. Sci. Appl. 02, 1194-1198.]). It is unclear from Oxmann's plots of Ca5(PO4)3(OH) whether the batch-to-batch differences are severe enough to suggest precluding its use as an energy standard, but these differences, combined with the lack of post-synthetic methods to purify this high-melting ceramic, made it appear less desirable. Similar assessments were made for AlPO4·(H2O)n and other water-insoluble phosphates and high-melting solids.

We next turned our attention to molecular phospho­rus compounds. Tri­phenyl­phosphine oxide (OPPh3) has been used for ligand K-edge XAS measurements (Adhikari et al., 2008[Adhikari, D., Mossin, S., Basuli, F., Huffman, J. C., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2008). J. Am. Chem. Soc. 130, 3676-3682.]; Harkins et al., 2008[Harkins, S. B., Mankad, N. P., Miller, A. J. M., Szilagyi, R. K. & Peters, J. C. (2008). J. Am. Chem. Soc. 130, 3478-3485.]; Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]; Mossin et al., 2012[Mossin, S., Tran, B. L., Adhikari, D., Pink, M., Heinemann, F. W., Sutter, J., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2012). J. Am. Chem. Soc. 134, 13651-13661.]; Shearer et al., 2012[Shearer, J., Callan, P. E., Masitas, C. A. & Grapperhaus, C. A. (2012). Inorg. Chem. 51, 6032-6045.]; Soma et al., 2016[Soma, S., Van Stappen, C., Kiss, M., Szilagyi, R. K., Lehnert, N. & Fujisawa, K. (2016). J. Biol. Inorg. Chem. 21, 757-775.]). OPPh3 has a resolved pre-edge feature and is commercially available (Fig. 1[link]). Unlike Na4P2O7, OPPh3 is hydro­phobic with low solubility in water, suggesting that peak positions are less susceptible to hydration effects. While OPPh3 has desirable attributes for a P K-edge XANES energy standard, there are some discrepancies in its reported calibration energy. The first pre-edge maximum has been reported at 2147.3 eV (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]; Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]), 2147.5 eV (Adhikari et al., 2008[Adhikari, D., Mossin, S., Basuli, F., Huffman, J. C., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2008). J. Am. Chem. Soc. 130, 3676-3682.]; Harkins et al., 2008[Harkins, S. B., Mankad, N. P., Miller, A. J. M., Szilagyi, R. K. & Peters, J. C. (2008). J. Am. Chem. Soc. 130, 3478-3485.]; Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]; Mossin et al., 2012[Mossin, S., Tran, B. L., Adhikari, D., Pink, M., Heinemann, F. W., Sutter, J., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2012). J. Am. Chem. Soc. 134, 13651-13661.]; Shearer et al., 2012[Shearer, J., Callan, P. E., Masitas, C. A. & Grapperhaus, C. A. (2012). Inorg. Chem. 51, 6032-6045.]) and 2145.7 eV (Soma et al., 2016[Soma, S., Van Stappen, C., Kiss, M., Szilagyi, R. K., Lehnert, N. & Fujisawa, K. (2016). J. Biol. Inorg. Chem. 21, 757-775.]). However, the latter two energy values were probably annotated incorrectly because they reference the reported value of 2147.3 eV from Hormes and co-workers (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]; Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). To investigate these discrepancies, we collected the P K-edge XANES spectrum of OPPh3. The first peak in OPPh3 was observed at 2147.3 eV when referenced to Na4P2O7, matching the value reported previously by Hormes and co-workers.

The use of three different reference values for OPPh3 is not a significant concern as long as people are aware they exist and take them into account when referencing their data to other data sets. However, this is likely to be overlooked. Knowing that future P K-edge XANES users will continue to propagate multiple calibration values for OPPh3, it was important to highlight the differences and describe a potential alternative with a single-reference value for future P K-edge XAS studies.

3.2. Investigation of PPh3 and PPh4Br

Our search for an alternative P K-edge XANES standard focused on compounds that could meet the following criteria: (1) have a well resolved pre-edge feature for precise energy calibration, (2) be hydro­phobic and resistant to hydrogen-bonding interactions that could perturb the electronic structure of phospho­rus, (3) be air-stable, (4) be resistant to photodecomposition, and (5) be commercially available. The first candidate was PPh3, a commercially available phosphine reported to have a well defined pre-edge feature at 2145.4 eV (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). However, it was soon discovered that PPh3 undergoes photodecomposition at beamline 14-3 at the Stanford Synchrotron Radiation Lightsource (Fig. 2[link]). The pre-edge peak was observed at 2145.3 eV and decreased as a function of time, and a feature assigned to OPPh3 grew in at 2150.1 eV. This latter peak was not present in P K-edge XANES spectra reported previously for PPh3 (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]; Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]), indicating that our PPh3 had already oxidized to some extent before the first scan was collected. The peak at ca. 2147.3 eV is present in the spectra of both PPh3 and OPPh3, and therefore does not change significantly in intensity as PPh3 oxidation occurs. The presence of other OPPh3 transitions accounts for the slight discrepancy in our measured PPh3 pre-edge feature at 2145.3 eV compared with the value reported previously for oxide-free PPh3 at 2145.4 eV (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]).

[Figure 2]
Figure 2
P K-edge XANES spectra of PPh3 collected at beamline 14-3 at SSRL.

The lack of decomposition in earlier P K-edge XAS data sets collected for PPh3 can likely be ascribed to the spectra collection using unfocused beams (Mankad et al., 2009[Mankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878-3880.]; Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). In contrast, our data were collected using a focused beam at a beamline typically configured for imaging work. While the observed photodecomposition indicated PPh3 was not a suitable calibration standard, the experiments revealed that beamline 14-3 at SSRL would provide an excellent proving ground to test the photostability of other potential P K-edge XANES standards.

We next investigated PPh4Br. PPh4+ is a common non-coordinating cation used in coordination chemistry to offset the charge of complex ions (Daly et al., 2012[Daly, S. R., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Kozimor, S. A. & Martin, R. L. (2012). J. Am. Chem. Soc. 134, 14408-14422.]; Minasian et al., 2012[Minasian, S. G., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Conradson, S. D., Kozimor, S. A., Martin, R. L., Schwarz, D. E., Shuh, D. K., Wagner, G. L., Wilkerson, M. P., Wolfsberg, L. E. & Yang, P. (2012). J. Am. Chem. Soc. 134, 5586-5597.]; Donahue et al., 2014[Donahue, C. M., Lezama Pacheco, J. S., Keith, J. M. & Daly, S. R. (2014). Dalton Trans. 43, 9189-9201.]; Olson et al., 2014[Olson, A. C., Keith, J. M., Batista, E. R., Boland, K. S., Daly, S. R., Kozimor, S. A., MacInnes, M. M., Martin, R. L. & Scott, B. L. (2014). Dalton Trans. 43, 17283-17295.]), and is commercially available as its chloride and bromide salt. We reasoned that PPh4+ should be more stable than PPh3 with respect to photodecomposition; PPh4+ is a fully oxidized P(V) complex like OPPh3, which is corroborated by our P K-edge XANES data (see below). We postulated that the higher symmetry of PPh4+ versus OPPh3 (and corresponding differences in selection rules) would lead to a more resolved pre-edge feature in the P K-edge XANES spectrum, thereby making it more ideal for energy referencing. If one considers only the carbon and oxygen atoms attached to phospho­rus, the point group symmetry of PPh4+ is Td whereas for OPPh3 it is C3V. As discussed below, the actual point group symmetries are slightly lower when the orientation of the phenyl groups are taken into consideration.

The P K-edge XAS spectrum of PPh4Br is provided in Fig. 3[link]. Unlike PPh3, no decomposition was observed over repeated scans performed intermittently on the same sample over one week (we used the same sample for calibration scans over an entire week of data collection). Three prominent features were observed prior to the higher energy shape resonances starting at 2153 eV. The first derivative trace revealed two maxima (i.e. where the first derivative trace crosses 0) at 2146.96 and 2149.85 eV. The second derivative trace was used to assign the energy of the shoulder located next to the peak at 2149.85 eV and identify underlying features contributing to the observed peaks. Three minima were observed at 2146.82, 2148.90 and 2149.90 eV corresponding to the three most prominent XANES features. In addition, peak shouldering was observed in the second derivative trace at approximately 2147.2 and 2148.1 eV (Fig. 4[link]). No lower energy pre-edge features were observed in the range known for P(III) complexes such as PPh3 (2145.4 eV; Fig. 2[link]) or PEt3 (2145.8 eV) (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]), consistent with the assignment of PPh4+ as a P(V) complex. The peak positions fall within the range of those described for OPPh3 and other P(V) complexes with phenyl substituents (SPPh3 and SePPh3) (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]).

[Figure 3]
Figure 3
P K-edge XANES spectrum of PPh4Br. First and second derivative traces, depicted with red and blue dashed lines, respectively, are offset on the y-axis by +2 intensity units.
[Figure 4]
Figure 4
Enlarged P K-edge XANES spectrum (grey) and second derivative trace (blue) of PPh4Br. The second derivative trace is offset on the y-axis by +2 intensity units and scaled for comparison with the XANES trace.

Our P K-edge peak positions for PPh4Br are reported with 0.01 eV precision based on the following justification. First, all of our spectra are referenced to the white line for Na4P2O7 at 2152.40 ± 0.05 eV (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]; Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). This value was determined by Hormes and co-workers using a double InSb(111) crystal monochromator calibrated to the Ar K-edge of argon gas at 3203.54 eV (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]). Second, spectra from each run were obtained by averaging no less than three scans individually calibrated to Na4P2O7. Positions were then obtained from the first derivative trace of the merged scan, so peak positions were not limited by the 0.08 eV step size used over the edge in our experiments. Finally, averaged scans from multiple runs were used to establish a standard deviation in the peak position, which was ±0.02 eV. This measured repeatability is better than the ∼0.1 eV energy reproducibility used to establish the Cs2CuCl4 and Na2SO3·5H2O pre-edge peak positions at 2820.20 eV and 2472.02 eV, respectively, for use as Cl and S K-edge XAS energy calibration standards (Solomon et al., 2005[Solomon, E. I., Hedman, B., Hodgson, K. O., Dey, A. & Szilagyi, R. K. (2005). Coord. Chem. Rev. 249, 97-129.]).

It is important to note that peak positions can change slightly depending on the resolution of the beamline used. While a theoretical resolution can be calculated based on the monochromator crystals and energy of the element edge, the actual resolution of the beamline depends on the experimental setup and can be difficult to measure in a reliable way. To address this issue, we have uploaded our normalized P K-edge XAS data of PPh4Br in the supporting information. We also curve-fit our spectra to obtain the full width at half-maximum (FWHM) of our first pre-edge feature. As described by Solomon et al. (2005)[Solomon, E. I., Hedman, B., Hodgson, K. O., Dey, A. & Szilagyi, R. K. (2005). Coord. Chem. Rev. 249, 97-129.] for ligand K-edge XAS data, the spectra were modeled with pseudo-Voigt functions containing 1:1 mixtures of Gaussian and Lorentzian lineshapes. The step function was modeled with 1:1 ratio of arctangent and error function contributions. Curve fits of nine P K-edge XANES spectra of PPh4Br gave an average FWHM of 1.02 ± 0.03 eV for the first pre-edge peak.

3.3. DFT and TDDFT calculations

DFT and TDDFT calculations were performed on PPh4+ to assign features in the P K-edge XANES spectrum. Calculations were also performed on the aforementioned PPh3 to compare how changes in oxidation state and symmetry affected the electronic structure. Optimized gas-phase structures of PPh3 and PPh4+ are provided in Fig. 5[link] and a truncated molecular orbital (MO) diagram is provided in Fig. 6[link]. C3 and D2d point group symmetries were assigned to the calculated structures of PPh3 and PPh4+, respectively, which take into account the phenyl group orientations. Mulliken symbols were assigned by analyzing the symmetry of the Kohn–Sham orbitals, as shown in Fig. 6[link] for the highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO).

[Figure 5]
Figure 5
Calculated structures of PPh3 (left) and PPh4+ (right). P—C bond distances: 1.851–1.852 Å (PPh3) and 1.812 Å (PPh4+); C—P—C bond angles = 101.7° (PPh3) and 110.3° (PPh4+). The bond distances and angles are nearly identical to those obtained experimentally by single-crystal XRD (Alcock et al., 1985[Alcock, N. W., Pennington, M. & Willey, G. R. (1985). Acta Cryst. C41, 1549-1550.]; Dunne & Orpen, 1991[Dunne, B. J. & Orpen, A. G. (1991). Acta Cryst. C47, 345-347.]).
[Figure 6]
Figure 6
Molecular orbital diagram of PPh3 and PPh4+ from DFT calculations. Kohn–Sham orbitals of the highest occupied and lowest unoccupied molecular orbitals are shown on the left (PPh3) and right (PPh4+). Unoccupied MOs depicted in red are those involved in the calculated P K-edge XAS transitions (TDDFT) in Fig. 7[link].

The calculated MOs for PPh4+ all decrease in energy relative to PPh3 as a result of the change in P oxidation state and molecular charge (+1 and 0). The HOMO for PPh3 (1a) is effectively the P lone pair with some phenyl mixing (33.7% P p-character and 9.3% P s-character), whereas the HOMO 1a2 in PPh4+ is non-bonding aryl π*. Despite the differences in symmetry and charge, the unoccupied MOs for both compounds have similar groups of MOs that can be classified according to their bonding between P and the phenyl groups. At lower energy are a cluster of MOs assigned to P-C bonding with phenyl π* orbitals and MOs at higher energy are best described as P-C σ*.

TDDFT was used to simulate the P K-edge XAS spectra of PPh3 and PPh4+ using the ground-state DFT results provided in Fig. 6[link]. The simulated spectra yield relatively good agreement with the experimental spectra once an energy shift was applied to account for errors in determining absolute energies, which is well established in ligand K-edge TDDFT calculations (Fig. 7[link]) (Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]). For PPh3, the first pre-edge feature was assigned to P 1s → 1e (P-C π*) and the second feature at higher energy was assigned to P 1s → 3e (P-C σ*). These both are formally assigned as 1A→1E, but we use the primary orbital contributions to the electronic states here and below to relate the transitions back to the MOs in Fig. 6[link]. A higher-energy transition assigned to P 1s → 4a (P-C σ*) was not observed in our data, presumably a result of masking by the OPPh3 peak from photodecomposition. No transitions of significant intensity were calculated under the peak assigned to the OPPh3 impurity at 2150.1 eV, consistent with its assignment to photodecomposition.

[Figure 7]
Figure 7
Comparison of experimental (solid black line) and simulated (dashed black line) P K-edge XAS for PPh3 (top; first scan from Fig. 2[link]) and PPh4Br (bottom) from TDDFT calculations. The transition energies and relative oscillator strengths are depicted as red bars. Major unoccupied orbital components contributing to the electronic transitions have been assigned to correspond to the Mulliken symbols from Fig. 6[link].

The simulated spectrum for PPh4+ also provided relatively good agreement with its P K-edge XAS spectrum. The TDDFT spectrum indicates that the first feature at 2146.96 eV is attributed to a combination of two closely spaced transitions assigned to P 1s → b2 and 1s → e at 2146.55 and 2147.29 eV, respectively. The presence of two transitions in the TDDFT data accounts for the shoulder observed at ∼2147.3 eV in the second derivative trace of the experimental spectrum (Fig. 4[link]). Two P-C σ* transitions calculated at 2149.24 and 2150.49 eV were assigned to P 1s → 3e and 1s → 2b2. These transitions account for the second and third features in the P K-edge spectrum. Overall, the TDDFT data corroborated the P K-edge results and allowed the features to be assigned to P 1s → P-C π* and 1s → P-C σ*. These findings are consistent with assignments previously made by Hormes based on MS Xα calculations for OPPh3 and PPh3 (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]).

4. Conclusions

In summary, we presented the P K-edge XANES spectrum of PPh4Br and compared it with other phospho­rus compounds used as P K-edge energy calibration standards. In addition to its exceptional photostability and commercial availability, PPh4Br has a highly resolved pre-edge feature at 2146.96 eV ideal for precise P K-edge XANES energy referencing. This peak was calibrated to the white line of Na4P2O7 at 2152.40 eV (Küper et al., 1994[Küper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741-1753.]) and cross-referenced to the OPPh3 pre-edge peak reported by Hormes and co-workers at 2147.3 eV (Engemann et al., 1999[Engemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61-75.]). We have used PPh4Br as a calibration standard for several P K-edge XANES data collection runs at SSRL (Lee et al., 2016[Lee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774-9785.]), and we saw no decomposition after repeated calibration scans on the same PPh4Br sample during a week of P K-edge XANES data collection. DFT and TDDFT calculations were used to assign features in the P K-edge XANES spectrum of PPh4Br, and the simulated spectrum from TDDFT corroborated our experimental results.

Overall, our analysis suggests that PPh4Br is an excellent alternative to more commonly used energy calibration standards used for P K-edge XANES spectroscopy, especially for ligand K-edge XAS analysis of coordination complexes containing phospho­rus ligands. As such, we propose that the PPh4Br pre-edge position at 2146.96 eV be used for P K-edge energy calibration in future ligand K-edge XAS studies.

Supporting information


Acknowledgements

We would like to thank Erik Nelson, Matthew Latimer, and Cynthia Patty for their assistance collecting XANES data at the Stanford Synchrotron Radiation Lightsource. We also thank Erik Nelson and Matthew Latimer for assisting with technical questions during the preparation of this manuscript. Use of the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, is supported by the US Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515.

Funding information

The following funding is acknowledged: American Chemical Society Petroleum Research Fund (grant No. 55989-DNI3 to Scott Daly) and Colgate University for computational support and funding.

References

First citationAdhikari, D., Mossin, S., Basuli, F., Huffman, J. C., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2008). J. Am. Chem. Soc. 130, 3676–3682.  CSD CrossRef CAS Google Scholar
First citationAlcock, N. W., Pennington, M. & Willey, G. R. (1985). Acta Cryst. C41, 1549–1550.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBauernschmitt, R. & Ahlrichs, R. (1996). J. Chem. Phys. 104, 9047–9052.  CrossRef CAS Google Scholar
First citationBaumann, K., Glaser, K., Mutz, J.-E., Karsten, U., MacLennan, A., Hu, Y., Michalik, D., Kruse, J., Eckhardt, K.-U., Schall, P. & Leinweber, P. (2017). Soil Biol. Biochem. 109, 156–166.  CrossRef CAS Google Scholar
First citationBecke, A. D. (1993). J. Chem. Phys. 98, 5648–5652.  CrossRef CAS Web of Science Google Scholar
First citationBlanchard, P. E. R., Grosvenor, A. P., Cavell, R. G. & Mar, A. (2008). Chem. Mater. 20, 7081–7088.  Web of Science CrossRef CAS Google Scholar
First citationDaly, S. R., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Kozimor, S. A. & Martin, R. L. (2012). J. Am. Chem. Soc. 134, 14408–14422.  CrossRef CAS Google Scholar
First citationDey, A., Okamura, T., Ueyama, N., Hedman, B., Hodgson, K. O. & Solomon, E. I. (2005). J. Am. Chem. Soc. 127, 12046–12053.  CrossRef CAS Google Scholar
First citationDonahue, C. M., Lezama Pacheco, J. S., Keith, J. M. & Daly, S. R. (2014). Dalton Trans. 43, 9189–9201.  CrossRef CAS Google Scholar
First citationDonahue, C. M., McCollom, S. P., Forrest, C. M., Blake, A. V., Bellott, B. J., Keith, J. M. & Daly, S. R. (2015). Inorg. Chem. 54, 5646–5659.  CrossRef CAS Google Scholar
First citationDunne, B. J. & Orpen, A. G. (1991). Acta Cryst. C47, 345–347.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationEngemann, C., Franke, R., Hormes, J., Lauterbach, C., Hartmann, E., Clade, J. & Jansen, M. (1999). Chem. Phys. 243, 61–75.  CrossRef CAS Google Scholar
First citationEveborn, D., Gustafsson, J. P., Hesterberg, D. & Hillier, S. (2009). Environ. Sci. Technol. 43, 6515–6521.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFerraz, M. P., Monteiro, F. J. & Manuel, C. M. (2004). J. Appl. Biomater. Biomech. 2, 74–80.  CAS Google Scholar
First citationFrisch, M. J. T. G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, N. J., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). Gaussian 09, Revision D.01. Gaussian Inc., Pittsburgh, Pennsylvania, USA.  Google Scholar
First citationGlaser, T., Hedman, B., Hodgson, K. O. & Solomon, E. I. (2000). Acc. Chem. Res. 33, 859–868.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGrimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.  Web of Science CrossRef PubMed Google Scholar
First citationGrosvenor, A. P., Cavell, R. G. & Mar, A. (2008). J. Solid State Chem. 181, 2549–2558.  CrossRef CAS Google Scholar
First citationHariharan, P. C. & Pople, J. A. (1973). Theor. Chim. Acta, 28, 213–222.  CrossRef CAS Web of Science Google Scholar
First citationHarkins, S. B., Mankad, N. P., Miller, A. J. M., Szilagyi, R. K. & Peters, J. C. (2008). J. Am. Chem. Soc. 130, 3478–3485.  CSD CrossRef CAS Google Scholar
First citationHashimoto, Y., Takamoto, A., Kikkawa, R., Murakami, K. & Yamaguchi, N. (2014). Environ. Sci. Technol. 48, 5486–5492.  CrossRef CAS Google Scholar
First citationHehre, W. J., Ditchfield, R. & Pople, J. A. (1972). J. Chem. Phys. 56, 2257–2261.  CrossRef CAS Web of Science Google Scholar
First citationHu, Y., Maclennan, A. & Sham, T. K. (2015). J. Lumin. 166, 143–147.  CrossRef CAS Google Scholar
First citationHuang, R. & Tang, Y. (2016). Water Res. 100, 439–447.  CrossRef CAS Google Scholar
First citationIngall, E. D., Brandes, J. A., Diaz, J. M., de Jonge, M. D., Paterson, D., McNulty, I., Elliott, W. C. & Northrup, P. (2011). J. Synchrotron Rad. 18, 189–197.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKasrai, M., Vasiga, M., Suominen Fuller, M., Bancroft, G. M. & Fyfe, K. (1999). J. Synchrotron Rad. 6, 719–721.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKim, B., Gautier, M., Rivard, C., Sanglar, C., Michel, P. & Gourdon, R. (2015). Environ. Sci. Technol. 49, 4903–4910.  CrossRef CAS Google Scholar
First citationKozimor, S. A., Yang, P., Batista, E. R., Boland, K. S., Burns, C. J., Christensen, C. N., Clark, D. L., Conradson, S. D., Hay, P. J., Lezama, J. S., Martin, R. L., Schwarz, D. E., Wilkerson, M. P. & Wolfsberg, L. E. (2008). Inorg. Chem. 47, 5365–5371.  CrossRef CAS Google Scholar
First citationKozimor, S. A., Yang, P., Batista, E. R., Boland, K. S., Burns, C. J., Clark, D. L., Conradson, S. D., Martin, R. L., Wilkerson, M. P. & Wolfsberg, L. E. (2009). J. Am. Chem. Soc. 131, 12125–12136.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKraal, P., Bostick, B. C., Behrends, T., Reichart, G.-J. & Slomp, C. P. (2015). Mar. Chem. 168, 1–8.  CrossRef CAS Google Scholar
First citationKruse, J. & Leinweber, P. (2008). J. Plant Nutr. Soil Sci. 171, 613–620.  Web of Science CrossRef CAS Google Scholar
First citationKruse, J., Negassa, W., Appathurai, N., Zuin, L. & Leinweber, P. (2010). J. Environ. Qual. 39, 2179–2184.  CrossRef CAS Google Scholar
First citationKüper, G., Hormes, J. & Sommer, K. (1994). Macromol. Chem. Phys. 195, 1741–1753.  Google Scholar
First citationLee, C., Yang, W. & Parr, R. G. (1988). Phys. Rev. B, 37, 785–789.  CrossRef CAS Web of Science Google Scholar
First citationLee, K., Wei, H., Blake, A. V., Donahue, C. M., Keith, J. M. & Daly, S. R. (2016). Dalton Trans. 45, 9774–9785.  CSD CrossRef CAS Google Scholar
First citationLiu, J., Hu, Y., Yang, J., Abdi, D. & Cade-Menun, B. J. (2015). Environ. Sci. Technol. 49, 168–176.  CrossRef CAS Google Scholar
First citationLiu, J., Yang, J., Cade-Menun, B. J., Liang, X., Hu, Y., Liu, C. W., Zhao, Y., Li, L. & Shi, J. (2013). J. Environ. Qual. 42, 1763–1770.  CrossRef CAS Google Scholar
First citationLöble, M. W., Keith, J. M., Altman, A. B., Stieber, S. C. E., Batista, E. R., Boland, K. S., Conradson, S. D., Clark, D. L., Lezama Pacheco, J., Kozimor, S. A., Martin, R. L., Minasian, S. G., Olson, A. C., Scott, B. L., Shuh, D. K., Tyliszczak, T., Wilkerson, M. P. & Zehnder, R. A. (2015). J. Am. Chem. Soc. 137, 2506–2523.  Google Scholar
First citationMankad, N. P., Antholine, W. E., Szilagyi, R. K. & Peters, J. C. (2009). J. Am. Chem. Soc. 131, 3878–3880.  CSD CrossRef CAS Google Scholar
First citationMinasian, S. G., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Conradson, S. D., Kozimor, S. A., Martin, R. L., Schwarz, D. E., Shuh, D. K., Wagner, G. L., Wilkerson, M. P., Wolfsberg, L. E. & Yang, P. (2012). J. Am. Chem. Soc. 134, 5586–5597.  CrossRef CAS Google Scholar
First citationMinasian, S. G., Keith, J. M., Batista, E. R., Boland, K. S., Clark, D. L., Kozimor, S. A., Martin, R. L., Shuh, D. K. & Tyliszczak, T. (2014). Chem. Sci. 5, 351–359.  CrossRef CAS Google Scholar
First citationMorel, F. L., Pin, S., Huthwelker, T., Ranocchiari, M. & van Bokhoven, J. A. (2015). Phys. Chem. Chem. Phys. 17, 3326–3331.  CrossRef CAS Google Scholar
First citationMossin, S., Tran, B. L., Adhikari, D., Pink, M., Heinemann, F. W., Sutter, J., Szilagyi, R. K., Meyer, K. & Mindiola, D. J. (2012). J. Am. Chem. Soc. 134, 13651–13661.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationNakahira, A., Nakata, K., Numako, C., Murata, H. & Matsunaga, K. (2011). Mater. Sci. Appl. 02, 1194–1198.  CAS Google Scholar
First citationNorthrup, P., Leri, A. & Tappero, R. (2016). Protein Pept. Lett. 23, 300–308.  CrossRef CAS Google Scholar
First citationOlson, A. C., Keith, J. M., Batista, E. R., Boland, K. S., Daly, S. R., Kozimor, S. A., MacInnes, M. M., Martin, R. L. & Scott, B. L. (2014). Dalton Trans. 43, 17283–17295.  CrossRef CAS Google Scholar
First citationOxmann, J. F. (2014). Biogeosciences, 11, 2169–2183.  CrossRef CAS Google Scholar
First citationPereira, G., Lachenwitzer, A., Munoz-Paniagua, D., Kasrai, M., Norton, P. R., Abrecht, M. & Gilbert, P. U. P. A. (2006). Tribol. Lett. 23, 109–119.  CrossRef CAS Google Scholar
First citationPereira, G., Munoz-Paniagua, D., Lachenwitzer, A., Kasrai, M., Norton, P. R., Capehart, T. W., Perry, T. A. & Cheng, Y.-T. (2007). Wear, 262, 461–470.  Web of Science CrossRef CAS Google Scholar
First citationPrietzel, J., Dümig, A., Wu, Y., Zhou, J. & Klysubun, W. (2013). Geochim. Cosmochim. Acta, 108, 154–171.  Web of Science CrossRef CAS Google Scholar
First citationPrietzel, J., Harrington, G., Häusler, W., Heister, K., Werner, F. & Klysubun, W. (2016). J. Synchrotron Rad. 23, 532–544.  CrossRef CAS IUCr Journals Google Scholar
First citationQueen, M. S., Towey, B. D., Murray, K. A., Veldkamp, B. S., Byker, H. J. & Szilagyi, R. K. (2013). Coord. Chem. Rev. 257, 564–578.  Web of Science CrossRef CAS Google Scholar
First citationRafiuddin, M. R. & Grosvenor, A. P. (2016). Inorg. Chem. 55, 9685–9695.  CrossRef CAS Google Scholar
First citationRafiuddin, M. R., Mueller, E. & Grosvenor, A. P. (2014). J. Phys. Chem. C, 118, 18000–18009.  CrossRef CAS Google Scholar
First citationRavel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537–541.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationRaynaud, S., Champion, E., Bernache-Assollant, D. & Thomas, P. (2002). Biomaterials, 23, 1065–1072.  Web of Science CrossRef CAS Google Scholar
First citationSharma, V., Doerr, N. & Aswath, P. B. (2016). RSC Adv. 6, 22341–22356.  CrossRef CAS Google Scholar
First citationShearer, J., Callan, P. E., Masitas, C. A. & Grapperhaus, C. A. (2012). Inorg. Chem. 51, 6032–6045.  CrossRef CAS Google Scholar
First citationSolomon, E. I., Hedman, B., Hodgson, K. O., Dey, A. & Szilagyi, R. K. (2005). Coord. Chem. Rev. 249, 97–129.  Web of Science CrossRef CAS Google Scholar
First citationSoma, S., Van Stappen, C., Kiss, M., Szilagyi, R. K., Lehnert, N. & Fujisawa, K. (2016). J. Biol. Inorg. Chem. 21, 757–775.  CSD CrossRef CAS Google Scholar
First citationSpencer, L. P., Yang, P., Minasian, S. G., Jilek, R. E., Batista, E. R., Boland, K. S., Boncella, J. M., Conradson, S. D., Clark, D. L., Hayton, T. W., Kozimor, S. A., Martin, R. L., MacInnes, M. M., Olson, A. C., Scott, B. L., Shuh, D. K. & Wilkerson, M. P. (2013). J. Am. Chem. Soc. 135, 2279–2290.  CSD CrossRef CAS Google Scholar
First citationTakamoto, A. & Hashimoto, Y. (2014). Chem. Lett. 43, 1696–1697.  CrossRef CAS Google Scholar
First citationTan, Y., Sun, D., Yu, H., Yang, B., Gong, Y., Yan, S., Chen, Z., Cai, Q. & Wu, Z. (2014). CrystEngComm, 16, 9657–9668.  CrossRef CAS Google Scholar
First citationYoshida, M., Mineo, T., Mitsutomi, Y., Yamamoto, F., Kurosu, H., Takakusagi, S., Asakura, K. & Kondoh, H. (2016). Chem. Lett. 45, 277–279.  CrossRef CAS Google Scholar
First citationZhang, Z., Yamaguchi, E. S., Kasrai, M. & Bancroft, G. M. (2005). Tribol. Lett. 19, 211–220.  CrossRef CAS Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds