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

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Development of serial X-ray fluorescence holography for radiation-sensitive protein crystals

crossmark logo

aDepartment of Physical Science and Engineering, Nagoya Institute of Technology, Gokiso, Showa, Nagoya 466-8555, Japan, bSynchrotron Radiation Research Center, Nagoya University, Furo, Chikusa, Nagoya 466-8603, Japan, cDivision of Biophysics, Department of Physiology, Jichi Medical University, Yakushiji, Shimotsuke, Tochigi 329-0498, Japan, dDepartment of Computer and Network Engineering, Graduate School of Information Sciences, Hiroshima City University, Asa-Minami-ku, Hiroshima 731-3194, Japan, eJapan Synchrotron Radiation Research Institute (JASRI), Sayo, Hyôgo 679-5198, Japan, fGraduate School of Information Sciences, Hiroshima City University, Asa-Minami-ku, Hiroshima 731-3194, Japan, gGraduate School of Science and Technology, Nara Institute of Science and Technology, Ikoma 630-0192, Japan, hGraduate School of Frontier Sciences, University of Tokyo, Kashiwa, Chiba 277-8561, Japan, and iResearch Institute for Interdisciplinary Science and Graduate School of Natural Science and Technology, Okayama University, Tsushima Naka, Okayama 700-8530, Japan
*Correspondence e-mail: ang.artoni@toyota-ti.ac.jp

Edited by R. W. Strange, University of Essex, United Kingdom (Received 17 July 2022; accepted 12 December 2022; online 20 January 2023)

X-ray fluorescence holography (XFH) is a powerful atomic resolution technique capable of directly imaging the local atomic structure around atoms of a target element within a material. Although it is theoretically possible to use XFH to study the local structures of metal clusters in large protein crystals, the experiment has proven difficult to perform, especially on radiation-sensitive proteins. Here, the development of serial X-ray fluorescence holography to allow the direct recording of hologram patterns before the onset of radiation damage is reported. By combining a 2D hybrid detector and the serial data collection used in serial protein crystallography, the X-ray fluorescence hologram can be directly recorded in a fraction of the measurement time needed for conventional XFH measurements. This approach was demonstrated by obtaining the Mn Kα hologram pattern from the protein crystal Photosystem II without any X-ray-induced reduction of the Mn clusters. Furthermore, a method to interpret the fluorescence patterns as real-space projections of the atoms surrounding the Mn emitters has been developed, where the surrounding atoms produce large dark dips along the emitter–scatterer bond directions. This new technique paves the way for future experiments on protein crystals that aim to clarify the local atomic structures of their functional metal clusters, and for other related XFH experiments such as valence-selective XFH or time-resolved XFH.

1. Introduction

Atomic resolution holography techniques, such as photoelectron holography, X-ray fluorescence holography and neutron holography, are a family of powerful experimental techniques that allow the direct probing of the local structures around atoms of a target element (Daimon, 2018[Daimon, H. (2018). J. Phys. Soc. Jpn, 87, 061001.], 2020[Daimon, H. (2020). Jpn. J. Appl. Phys. 59, 010504.]). These techniques are capable of simultaneously recording the intensity and phase of the scattered beams, allowing the direct, model-free extraction of the 3D positions of the scatterer atoms around the emitter atoms.

X-ray fluorescence holography (XFH) (Tegze & Faigel, 1996[Tegze, M. & Faigel, G. (1996). Nature, 380, 49-51.]; Faigel & Tegze, 1999[Faigel, G. & Tegze, M. (1999). Rep. Prog. Phys. 62, 355-393.]; Hayashi et al., 2012[Hayashi, K., Happo, N., Hosokawa, S., Hu, W. & Matsushita, T. (2012). J. Phys. Condens. Matter, 24, 093201.]; Hayashi & Korecki, 2018[Hayashi, K. & Korecki, P. (2018). J. Phys. Soc. Jpn, 87, 061003.]), in particular, is a robust and powerful technique that has been used to reveal the local structures around active sites in various functional materials. In XFH, atoms of a target element are excited by an incident X-ray, which then emit fluorescent X-rays. These fluorescent X-rays act as a reference wave, which is then scattered by the surrounding atoms (object wave). The interference between the reference and object waves stores information about the 3D arrangement of the scatterers. This interference pattern can be treated as a hologram pattern, where the 3D arrangement of atoms can be directly reconstructed using Barton's method (Barton, 1991[Barton, J. J. (1991). Phys. Rev. Lett. 67, 3106-3109.]), which is a Fourier transform based algorithm, or advanced reconstruction algorithms such as the L1-regularized linear regression method (Matsushita, 2018[Matsushita, T. (2018). Phys. Status Solidi B, 255, 1800091.]).

While almost all reported XFH experiments have been on inorganic functional materials, XFH is also expected to be capable of resolving the local structures around metal atoms in organic materials. This was first proposed back in 1996 (Fadley & Len, 1996[Fadley, C. S. & Len, P. M. (1996). Nature, 380, 27-28.]); however, experimental difficulties in dealing with organic samples have significantly stalled progress in this field. These difficulties included low metal concentrations, large complicated unit cells, samples consisting mostly of light elements and the susceptibility of these samples to radiation damage. However, recent advances in synchrotron radiation facilities, X-ray detectors and cryogenic cooling have provided possible solutions to these challenges.

Work on adapting XFH for organic samples started recently, with the first bioXFH setup for protein crystal samples developed and tested on human hemoglobin (Hb) crystals in 2016 (Sato-Tomita et al., 2016[Sato-Tomita, A., Shibayama, N., Happo, N., Kimura, K., Okabe, T., Matsushita, T., Park, S. Y., Sasaki, Y. C. & Hayashi, K. (2016). Rev. Sci. Instrum. 87, 063707.]). The experimental difficulties were addressed by using large P41212 Hb crystals, a toroidal graphite energy analyzer crystal to collect and focus the fluorescent X-rays, an N2 cold-gas flow system and the introduction of a new χ-circle stage and sample holder. While the experiment successfully prevented radiation damage to the protein crystal, the atomic reconstruction has not yet been fully interpreted. The resulting reconstructed atomic image is a complicated superposition of the 16 different Fe local environments, requiring more sophisticated methods of analysis. A more recent attempt on sperm whale myoglobin, a simpler crystal with the space group P21, successfully obtained the reconstructed image (Sato-Tomita et al., 2022[Sato-Tomita, A., Ang, A. K. R., Kimura, K., Marumi, R., Happo, N., Matsushita, T., Park, S., Shibayama, N., Sasaki, Y. C. & Hayashi, K. (2022). Biochem. Biophys. Res. Commun. 635, 277-282.]). In this work, the atomic image around the Fe heme was reconstructed from the measured holograms, and some of the atomic images reflected the actual atomic positions.

Recently, clear atomic reconstruction was obtained and interpreted from XFH experiments on the organic superconductor κ-(BEDT-TTF)2Cu[N(CN)2]Br or κ-Br (Ang et al., 2021[Ang, A. K. R., Marumi, R., Sato-Tomita, A., Kimura, K., Happo, N., Akagi, K., Sasaki, T. & Hayashi, K. (2021). Phys. Rev. B, 103, 214106.]). κ-Br is an organic charge-transfer salt that is widely studied in strongly correlated electron physics, where an Anderson-type metal–insulator transition is induced by the introduction of random defects by X-ray irradiation (Sano et al., 2010[Sano, K., Sasaki, T., Yoneyama, N. & Kobayashi, N. (2010). Phys. Rev. Lett. 104, 217003.]). Atomic reconstructions from XFH, molecular dynamics simulations and hologram calculations have shown experimental evidence for the previously proposed `bond-shifted' model (Kang et al., 2017[Kang, L., Akagi, K., Hayashi, K. & Sasaki, T. (2017). Phys. Rev. B, 95, 214106.]). Diffraction images and resistivity measurements before and after XFH experiments showed no additional radiation damage. Contrary to the complicated local structures around the Fe heme in hemoglobin or myoglobin, there are only four inequivalent Cu sites in κ-Br and they all lie in the anion layer, significantly simplifying the interpretation of the atomic reconstructions.

To date, all bioXFH experiments reported have been performed using the bioXFH apparatus, in which the angular dependence of the fluorescent X-rays is recorded as the sample is rotated point-by-point along the polar and azimuthal angles, taking several hours to record a full hologram pattern (Sato-Tomita et al., 2016[Sato-Tomita, A., Shibayama, N., Happo, N., Kimura, K., Okabe, T., Matsushita, T., Park, S. Y., Sasaki, Y. C. & Hayashi, K. (2016). Rev. Sci. Instrum. 87, 063707.], 2022[Sato-Tomita, A., Ang, A. K. R., Kimura, K., Marumi, R., Happo, N., Matsushita, T., Park, S., Shibayama, N., Sasaki, Y. C. & Hayashi, K. (2022). Biochem. Biophys. Res. Commun. 635, 277-282.]; Ang et al., 2021[Ang, A. K. R., Marumi, R., Sato-Tomita, A., Kimura, K., Happo, N., Akagi, K., Sasaki, T. & Hayashi, K. (2021). Phys. Rev. B, 103, 214106.]). For more robust crystals like hemoglobin or myoglobin, cooling the sample to 100 K and controlling the photon flux enable collection of the hologram pattern before the onset of radiation-induced global damage. No changes in the diffraction patterns were observed after recording the holograms. However, at these dosage levels, specific damage around water molecules and active redox sites can still occur (Garman, 2010[Garman, E. F. (2010). Acta Cryst. D66, 339-351.]), and this bioXFH apparatus is no longer appropriate for these kinds of sensitive protein crystals.

An example of sensitive protein crystals is the protein membrane complex Photosystem II [PSII (Dau & Haumann, 2008[Dau, H. & Haumann, M. (2008). Coord. Chem. Rev. 252, 273-295.]; Shen, 2015[Shen, J. R. (2015). Annu. Rev. Plant Biol. 66, 23-48.])]. The oxygen-evolving complex (OEC) of PSII is the catalytic center of the photosynthetic oxidation of water by green plants. In this process, the OEC, which consists of an Mn4CaO5 cluster, cycles through a series of redox states as described by the Kok cycle, where these states are designated as Si (where i = 0–4). The structure of PSII, and its OEC, have been extensively studied by synchrotron X-ray diffraction (XRD) with gradually increasing resolution (Zouni et al., 2001[Zouni, A., Witt, H. T., Kern, J., Fromme, P., Krauss, N., Saenger, W. & Orth, P. (2001). Nature, 409, 739-743.]; Kamiya & Shen, 2003[Kamiya, N. & Shen, J.-R. (2003). Proc. Natl Acad. Sci. 100, 98-103.]; Ferreira et al., 2004[Ferreira, K. N., Iverson, T. M., Maghlaoui, K., Barber, J. & Iwata, S. (2004). Science, 303, 1831-1838.]; Guskov et al., 2009[Guskov, A., Kern, J., Gabdulkhakov, A., Broser, M., Zouni, A. & Saenger, W. (2009). Nat. Struct. Mol. Biol. 16, 334-342.]; Umena et al., 2011[Umena, Y., Kawakami, K., Shen, J. R. & Kamiya, N. (2011). Nature, 473, 55-60.]). The Mn4CaO5 cluster consists of a cubane-like structure, with one Ca and three Mn atoms (Mn1D, Mn2C, Mn3B) occupying four corners, and O occupying the other four, while the last Mn atom (Mn4A) is connected to the cubane by two di-μ-oxo-bridges [Fig. 1[link]]. Although the PSII structure obtained at 1.9 Å resolution provided remarkable details (Umena et al., 2011[Umena, Y., Kawakami, K., Shen, J. R. & Kamiya, N. (2011). Nature, 473, 55-60.]), the average Mn–ligand and Mn–Mn distances were slightly longer than those extracted from extended X-ray absorption fine structure [EXAFS (Yano, Pushkar et al., 2005[Yano, J., Pushkar, Y., Glatzel, P., Lewis, A., Sauer, K., Messinger, J., Bergmann, U. & Yachandra, V. (2005). J. Am. Chem. Soc. 127, 14974-14975.]; Dau et al., 2008[Dau, H., Grundmeier, A., Loja, P. & Haumann, M. (2008). Philos. Trans. R. Soc. B, 363, 1237-1244.]; Glöckner et al., 2013[Glöckner, C., Kern, J., Broser, M., Zouni, A., Yachandra, V. & Yano, J. (2013). J. Biol. Chem. 288, 22607-22620.])]. Recently, the `low-dose' structure of PSII was revealed using extremely low-dose synchrotron XRD (Tanaka et al., 2017[Tanaka, A., Fukushima, Y. & Kamiya, N. (2017). J. Am. Chem. Soc. 139, 1718-1721.]), and the `damage-free' structure was obtained from a combination of large crystals and the femtosecond X-ray pulses of an X-ray free-electron laser (XFEL) (Suga et al., 2015[Suga, M., Akita, F., Hirata, K., Ueno, G., Murakami, H., Nakajima, Y., Shimizu, T., Yamashita, K., Yamamoto, M., Ago, H. & Shen, J. R. (2015). Nature, 517, 99-103.]), which provided shorter Mn–Mn distances, consistent with those obtained from EXAFS studies. The valence of each Mn ion in the Mn4CaO5 cluster of PSII S1 [2 × Mn(III), 2 × Mn(IV)] have been asserted based on the results of experiments such as X-ray absorption spectroscopy [XAS (Dau et al., 2008[Dau, H., Grundmeier, A., Loja, P. & Haumann, M. (2008). Philos. Trans. R. Soc. B, 363, 1237-1244.]; Glöckner et al., 2013[Glöckner, C., Kern, J., Broser, M., Zouni, A., Yachandra, V. & Yano, J. (2013). J. Biol. Chem. 288, 22607-22620.]; Roelofs et al., 1996[Roelofs, T. A., Liang, W., Latimer, M. J., Cinco, R. M., Rompel, A., Andrews, J. C., Sauer, K., Yachandra, V. K. & Klein, M. P. (1996). Proc. Natl Acad. Sci. USA, 93, 3335-3340.]; Yachandra et al., 1996[Yachandra, V. K., Sauer, K. & Klein, M. P. (1996). Chem. Rev. 96, 2927-2950.]; Robblee et al., 2001[Robblee, J. H., Cinco, R. M. & Yachandra, V. K. (2001). Biochim. Biophys. Acta, 1503, 7-23.]], Fourier transform infrared spectroscopy (Chu et al., 2001[Chu, H. A., Hillier, W., Law, N. A. & Babcock, G. T. (2001). Biochim. Biophys. Acta, 1503, 69-82.], 2004[Chu, H. A., Hillier, W. & Debus, R. J. (2004). Biochemistry, 43, 3152-3166.]; Debus et al., 2005[Debus, R. J., Strickler, M. A., Walker, L. M. & Hillier, W. (2005). Biochemistry, 44, 1367-1374.]), various electron paramagnetic resonance (EPR) spectroscopy techniques (Kulik et al., 2007[Kulik, L. V., Epel, B., Lubitz, W. & Messinger, J. (2007). J. Am. Chem. Soc. 129, 13421-13435.]; Cox et al., 2011[Cox, N., Rapatskiy, L., Su, J. H., Pantazis, D. A., Sugiura, M., Kulik, L., Dorlet, P., Rutherford, A. W., Neese, F., Boussac, A., Lubitz, W. & Messinger, J. (2011). J. Am. Chem. Soc. 133, 3635-3648.]; Stich et al., 2011[Stich, T. A., Yeagle, G. J., Service, R. J., Debus, R. J. & Britt, R. D. (2011). Biochemistry, 50, 7390-7404.]; Asada et al., 2013[Asada, M., Nagashima, H., Koua, F. H. M., Shen, J. R., Kawamori, A. & Mino, H. (2013). Biochim. Biophys. Acta, 1827, 438-445.]), through the analysis of the Jahn–Teller distortion effects from XRD studies (Suga et al., 2015[Suga, M., Akita, F., Hirata, K., Ueno, G., Murakami, H., Nakajima, Y., Shimizu, T., Yamashita, K., Yamamoto, M., Ago, H. & Shen, J. R. (2015). Nature, 517, 99-103.]) and from theoretical calculations. These results suggest that the typical doses in protein crystallography experiments reduce, or partially reduce, the Mn ions and change the local structure around Mn, making the determination of the structural arrangement and the valence states of Mn difficult.

[Figure 1]
Figure 1
Crystal structure of PSII determined from protein X-ray crystallography (PDB entry 3wu2), with the relative positions of the Mn4CaO5 clusters highlighted. Inset: the Mn4CaO5 cluster.

In this work, we develop a novel approach to determine the local structure of the metal clusters in protein crystals. By adapting XFH for low doses, the local structure of these metal clusters can be directly obtained. To minimize X-ray exposure of the crystals during the XFH measurements, the holograms can be directly imaged using a 2D hybrid pixel detector allowing much faster data acquisition. We have previously demonstrated the direct imaging of the Fe Kα holograms of the mixed-valence compound magnetite [Fe3O4 (Ang et al., 2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.])]. Building from this work, the principles of serial crystallography were incorporated into the experiment to further reduce the X-ray exposure of the protein crystal samples. This new approach is demonstrated on large PSII S1 crystals, where the crystal structure and Mn valence states are already well established in the literature. By scanning the irradiation point across the surface of several PSII crystals, the holograms were directly imaged using a 2D hybrid pixel detector prior to the onset of X-ray-induced reduction of the metal clusters. By interpreting the holograms in terms of dips in the fluorescent X-ray intensity in the forward-scattering directions, real-space projections of the arrangement of the atoms in the Mn clusters were obtained. Furthermore, we attempted valence-sensitive XFH by tuning the incident X-ray energy based on the small shifts in the Mn K-edge of the different Mn ions. The real space projections in the Mn(III) and Mn(IV) holograms showed modest indications of the different local structure around these ions. These results demonstrate a new and straightforward approach for XFH experiments on highly sensitive protein crystals and the future possibility of simultaneous valence-selective XFH experiments on the metal clusters in protein crystals.

2. Material and methods

2.1. X-ray fluorescence holography measurements

Serial X-ray fluorescence holography (sXFH) experiments were performed on BL39XU of SPring-8, Japan, using X-rays focused by Kirkpatrick–Baez (KB) mirrors (Suzuki et al., 2013[Suzuki, M., Kawamura, N., Mizumaki, M., Terada, Y., Uruga, T., Fujiwara, A., Yamazaki, H., Yumoto, H., Koyama, T., Senba, Y., Takeuchi, T., Ohashi, H., Nariyama, N., Takeshita, K., Kimura, H., Matsushita, T., Furukawa, Y., Ohata, T., Kondo, Y., Ariake, J., Richter, J., Fons, P., Sekizawa, O., Ishiguro, N., Tada, M., Goto, S., Yamamoto, M., Takata, M. & Ishikawa, T. (2013). J. Phys. Conf. Ser. 430, 012017.]). Using the KB mirrors and a variable Al film attenuator, the beam spot size and the X-ray photon flux were set to 7 µm × 10 µm and ∼2 × 109 photons s−1, respectively. To perform sXFH experiments, the bioXFH apparatus previously described by Sato-Tomita et al. (2016[Sato-Tomita, A., Shibayama, N., Happo, N., Kimura, K., Okabe, T., Matsushita, T., Park, S. Y., Sasaki, Y. C. & Hayashi, K. (2016). Rev. Sci. Instrum. 87, 063707.]) was modified to allow the scanning of the irradiation point, and the use of a 2D X-ray detector. A schematic of the sXFH apparatus is shown in Fig. 2[link].

[Figure 2]
Figure 2
sXFH experiment to visualize the local structure around metal atoms in protein crystals at 100 K. (a) Schematic of the sXFH measurements. (b) Photograph of the experimental apparatus taken at BL39-XU, SPring-8. Reproduced with permission from The Japan Society of Applied Physics (Copyright 2020) (Ang et al., 2020[Ang, A. K. R., Sato-Tomita, A., Shibayama, N., Umena, Y., Happo, N., Marumi, R., Kimura, K., Matsushita, T., Akagi, K., Sasaki, T., Sasaki, Y. C. & Hayashi, K. (2020). Jpn. J. Appl. Phys. 59, 010505.]).

The sXFH apparatus consists of a precision motorized four-axis (YZ-swivel-tilt) sample stage that is mounted on the χ-circle stage of the bioXFH apparatus (Sato-Tomita et al., 2016[Sato-Tomita, A., Shibayama, N., Happo, N., Kimura, K., Okabe, T., Matsushita, T., Park, S. Y., Sasaki, Y. C. & Hayashi, K. (2016). Rev. Sci. Instrum. 87, 063707.]). This entire assembly is mounted on a 2θ goniometer, where the sample stage assembly and the detector assembly can be rotated independently.

The samples are cooled to 100 K using a liquid-nitro­gen gas-flow system (Cryostream 800, Oxford Cryosystems, Inc.) with the cryostream nozzle set according to the protocols for cryogenic X-ray crystallography.

The scanning of the irradiation point necessary for these experiments is incompatible with inverse-mode XFH, where the sample is rotated point-by-point along the θ axis (0–75°) and ϕ axis (0–360°) as the fluorescent X-ray is recorded. In normal-mode XFH, the hologram can be directly imaged with a hybrid pixel detector (Ang et al., 2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]; Bortel et al., 2019[Bortel, G., Faigel, G., Tegze, M. & Angelov, B. (2019). J. Synchrotron Rad. 26, 170-174.]), allowing faster data acquisition.

To record the holograms, a Medipix3-based quad chip version of the Merlin 2D detector system (Quantum Detectors) was used (Plackett et al., 2013[Plackett, R., Horswell, I., Gimenez, E. N., Marchal, J., Omar, D. & Tartoni, N. (2013). J. Instrum. 8, C01038.]; Ballabriga et al., 2018[Ballabriga, R., Campbell, M. & Llopart, X. (2018). Nucl. Instrum. Methods Phys. Res. A, 878, 10-23.]). This is the same detector used in a previous valence-sensitive normal-mode XFH experiment (Ang et al., 2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]), and in a related experiment on Kossel lines (Bortel et al., 2016[Bortel, G., Faigel, G., Tegze, M. & Chumakov, A. (2016). J. Synchrotron Rad. 23, 214-218.]; Faigel et al., 2016[Faigel, G., Bortel, G. & Tegze, M. (2016). Sci. Rep. 6, 22904.]). The detector features high spatial resolution and dynamic range (256 × 256 pixels × 24 bit at 110 µm pixel size), which makes it suitable for holography experiments. The much larger EIGER X 1M 2D detector has also been used in a normal-mode XFH apparatus optimized for speed, where statistically relevant Ni holograms from an NiO crystal were recorded in a single image taken in 1 s (Bortel et al., 2019[Bortel, G., Faigel, G., Tegze, M. & Angelov, B. (2019). J. Synchrotron Rad. 26, 170-174.]). However, the Merlin quad chip detector allows two simultaneous energy thresholds to measure photons in a narrow-energy-window mode, which allows the user to record element-selective holograms from more complicated samples.

In this setup, the Merlin detector was set 30 mm from the sample. A Cr filter (4 µm film on a 8 µm Kapton sheet) was placed in front of the detector array to block the incident X-rays. Additional shielding was also set in front of the detector to prevent unwanted scattering reaching the detector.

2.2. Data collection

Each PSII crystal is scanned along the y and z directions as indicated in Fig. 2[link](a), with step sizes of 30 µm and 40 µm, respectively, to distribute the dosage and minimize radiation damage (Yano, Kern et al., 2005[Yano, J., Kern, J., Irrgang, K.-D., Latimer, M. J., Bergmann, U., Glatzel, P., Pushkar, Y., Biesiadka, J., Loll, B., Sauer, K., Messinger, J., Zouni, A. & Yachandra, V. K. (2005). Proc. Natl Acad. Sci. 102, 12047-12052.]). Images of the fluorescence intensity patterns were recorded at each point at an integration time of 10 s, resulting in an average dose of 0.15 MGy per point, as calculated using the program RADDOSE (Zeldin et al., 2013[Zeldin, O. B., Gerstel, M. & Garman, E. F. (2013). J. Appl. Cryst. 46, 1225-1230.]). Table 1[link] summarizes the experimental conditions of the sXFH measurements on PSII.

Table 1
Summary of the experimental conditions of the sXFH measurements on PSII

  Unit-cell parameters   Total photons collected (×106)
Sample (Å) (°) No. of scan positions At χH, 6.565 keV At χL, 6.551 keV
001 a = 123.807, b = 228.591, c = 286.008 α = β = γ = 90 231 4.30 3.58
002 a = 125.444, b = 228.707, c = 286.563 α = β = γ = 90 144 3.17 2.59
003 a = 124.75, b = 229.485, c = 285.433 α = β = γ = 90 72 2.19 1.70
004 a = 124.234, b = 228.938, c = 285.951 α = β = γ = 90 80 2.22 1.65
005 a = 134.579, b = 227.073, c = 286.66 α = β = γ = 90 179 2.12 1.77
006 a = 121.072, b = 227.512, c = 286.063 α = β = γ = 90 138 1.76 1.43
007 a = 123.024, b = 228.847, c = 286.000 α = β = γ = 90 77 2.01 1.57
008 a = 124.049, b = 229.029, c = 290.245 α = β = γ = 90 143 4.64 3.74
009 a = 121.633, b = 228.184, c = 286.12 α = β = γ = 90 164 5.07 3.76

The fluorescence intensity patterns were recorded using the 24 bit differential mode of the Merlin detector, with two simultaneous energy thresholds set at 5.15 keV and 6.15 keV. The fluorescence spectra at the detector position were confirmed with an SDD detector (XR-100SDD, Amptek Co. Ltd). A typical fluorescence spectrum from a PSII crystal, taken at an incident X-ray energy of 6.565 keV, is shown in Fig. 3[link](a), where the Mn Kα and Ca Kα peaks and a peak from the incident X-ray can be clearly observed. The Ca Kα peak contains signals from both the Ca atoms in the OEC of PSII and the Ca atoms in the cryoprotectant. An additional fluorescence peak at ∼4.5 keV was also detected and was attributed to the Litholoop. Aside from the energy-windowing from the simultaneous energy thresholds, a Cr thin-film filter was also placed in front of the detector to block the scattered incident X-rays.

[Figure 3]
Figure 3
(a) Typical fluorescence spectra from a PSII crystal showing the Mn Kα and Ca Kα peaks, and from an empty Litholoop mesh taken at an incident X-ray energy of 6.565 keV. A typical Mn Kα hologram image from a PSII crystal at different stages of data processing: (b) the total image from a single crystal, (c) normalized by dividing with background image, (d) after converting to a spherical projection, and (e) after flattening and rotating the pattern to the correct orientation.

2.3. Hologram data processing

Before data processing, the raw images (energy-windowed 5.15–6.15 keV) obtained from each PSII crystal are first integrated, and any hot or dead pixels are removed using a 3 × 3 median filter. Figs. 3[link](b)–3(e) show a typical hologram pattern obtained from a PSII crystal at different stages of data processing.

To extract the hologram pattern from the fluorescence images, several data processing steps are needed. These are described in more detail by Ang et al. (2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]); Bortel et al. (2019[Bortel, G., Faigel, G., Tegze, M. & Angelov, B. (2019). J. Synchrotron Rad. 26, 170-174.]); and Matsushita, Muro, Matsui et al. (2020[Matsushita, T., Muro, T., Matsui, F., Happo, N. & Hayashi, K. (2020). Jpn. J. Appl. Phys. 59, 020502.]). First, all the images taken while scanning the irradiation point from each sample are integrated [Fig. 3[link](b)]. Then, each image (In) is normalized using a normalizing pattern (P), obtained by integrating several images taken as the sample is slowly rotated in-plane. By rotating the sample, the holographic signal is averaged out in P, and only the low-frequency fluorescence background attributed to the sample–detector geometry remain. By normalizing each image pixel-by-pixel, Hn = In/P [Fig. 3[link](c)]; this low-frequency background and any inhomogeneity in the sensitivity of the detector pixels can be removed. An additional histogram filter was applied to reduce noise.

The images Hn, which are gnomonic projections, 2D projections of the spherical hologram pattern on the detector surface, are converted to a spherical projection [Fig. 3[link](d)]. To further remove low-frequency fluorescence background signals, the images are flattened by applying a Gaussian convolution on the images (Gn), and dividing the original image with the background χn = Hn/Gn. The images are then rotated based on their orientations, as determined from the indexing of the diffraction images obtained from each sample. At this point, the fluorescence intensity oscillates around 1. The final hologram fragments from the different samples are composed into one hologram χ. Finally, a constant 1 is subtracted, and a low pass filter (σ = 10°) is applied.

To expand the recorded hologram in k-space, symmetry operations were applied based on the symmetry of the PSII S1 crystal.

2.4. Hologram pattern simulations

The intensity of the X-ray fluorescence holograms, χ(k), is expressed as

[\chi ({\bf k}) \,\simeq\, 2\,{\rm{Re}}\sum _{j}\left\{ {{{r}_{\rm e\,}\,{f}_{\!j}\,({\theta }_{{{\bf r}}_{\,{j}}}^{\,{\bf k}}\,)}\over{{r}_{j}}}\exp \left [i\,\left(-{\bf k}\cdot {\bf r}_{\,j}-kr_j\right) \right ]\right\}, \eqno (1)]

where re is the free electron radius, fj is the atomic structure factor of the jth atom, and θrj is the angle between k and rj. For organic materials, however, the dynamic fluctuations caused by thermal vibrations are stronger than in inorganic materials, making it necessary to consider the effects of thermal vibrations. Thermal vibrations can be considered as atoms oscillating around their ideal positions with a Gaussian distribution with a standard deviation, σ, given as 〈u2〉/6 = σ2/2. This introduces an additional term in the holographic oscillation (Matsushita, Muro, Yokoya et al., 2020[Matsushita, T., Muro, T., Yokoya, T., Terashima, K., Kato, Y., Matsui, H., Maejima, N., Hashimoto, Y. & Matsui, F. (2020). Phys. Status Solidi B, 257, 2000117.]),

[\eqalignno { \chi ({\bf k}) \,\simeq\, & \,2\, {\rm{Re}}\sum _{j}\left\{{{{r}_{\rm{e}}\,{f}_{\!j}\,({\theta }_{{{\bf r}}_{{\,j}}}^{\,{\bf k}}\,)}\over{{r}_{j}}}\exp\left[i\,\left(-{\bf k}\cdot {{\bf r}}_{\,{j}}-k{r}_{j}\right)\right]\right\}\cr & \times \exp{\left[-{\sigma }^{2}{|{\bf k}|}^{2}\left(1-\cos{\theta }_{r\,j}\right)\right]}. & (2) }]

To simplify the calculations and reduce the computational times, the holograms were calculated using clusters constructed from only the Mn, O and Ca atoms within the OEC of PSII, based on atomic positions in the PDB entry 3wu2. For the total Mn Kα holograms, 32 clusters, each centered on an Mn emitter atom, were created and used for the calculations. The angular positions (θ, ϕ) of the forward scattering (FS) dips for all Mn and Ca scatterer atoms from each Mn emitter atom were also calculated using the atomic positions in 3wu2 and are indicated by the purple and green circles overlaid on the hologram patterns.

2.5. Sample purification and crystallization

For the XFH experiments, the holograms were measured from many large isomorphous PSII crystals. The XFH signal from the sample was weak due to the low-dose irradiation that was necessary to avoid the X-ray reduction of the Mn atoms in the Mn4CaO5 cluster as much as possible. The samples were extracted from the thermophilic cyano­bacteria, Thermo­synechococcus vulcanus, and the PSII crystals were prepared based on the methods reported by Kawakami & Shen (2018[Kawakami, K. & Shen, J. R. (2018). Methods Enzymol. 613, 1-16.]).

To maintain the isomorphism between PSII crystals, the crystals were replaced with different concentrations of the cryo-protectant solutions step-by-step using a gentle dialysis method. The post-crystallization process using dialysis membrane (MWCO6000-8000 purchased from Spectra/Por dialysis) with a molecular weight cutoff of 6–8 kDa pore size was divided into six steps, in which each step was treated every hour, from the first crystal solution containing 10% PEG3000 to the final cryo-protectant solution containing 25% PEG3000 (polyethlyne glycol) and 20% di­methyl sulfoxide (DMSO). When a PSII crystal is frozen by flash-cooling, excess cryo-protectant solution covering the surface of a PSII crystal must be excluded, as this causes the attenuation of X-ray fluorescence. Therefore, the PSII crystal was placed on a mesh loop of diameter 400 µm or 1000 µm with 40 µm × 40 µm spacing, and the excess solution was absorbed from the back side by filter paper before flash-cooling the PSII crystal. To obtain different sections of the hologram pattern, PSII crystals oriented in different directions were prepared.

2.6. Evaluation of radiation damage

To evaluate the radiation damage introduced into the samples during the XFH experiments, the XANES spectra taken before and after the XFH scans are compared. The `before' spectrum was obtained at a point on the surface of the sample outside the XFH scan region, and another spectrum was obtained from the center of the sample after the XFH scans with 1 eV steps and 2 s integration time. The Mn K-edge inflection point is determined from the zero point in the second derivative.

2.7. Determination of crystal orientation

The orientation of each PSII crystal relative to the Merlin detector was obtained from the indexing of the diffraction images taken by the Pilatus 100 K. Diffraction images were taken at a wavelength of 1.89 Å and indexed using XDS (Kabsch, 2010[Kabsch, W. (2010). Acta Cryst. D66, 125-132.]).

3. Results

3.1. Serial X-ray fluorescence holography

Fig. 4[link](a) shows typical Mn K-edge absorption spectra of a PSII S1 crystal taken before and after the XFH experiment, which shows no observable shift in the Mn K-edge. The inflection points (zero point of the second derivative) at 6.551 keV agree with the previously reported inflection points for the S1 state (Roelofs et al., 1996[Roelofs, T. A., Liang, W., Latimer, M. J., Cinco, R. M., Rompel, A., Andrews, J. C., Sauer, K., Yachandra, V. K. & Klein, M. P. (1996). Proc. Natl Acad. Sci. USA, 93, 3335-3340.]; Robblee et al., 2001[Robblee, J. H., Cinco, R. M. & Yachandra, V. K. (2001). Biochim. Biophys. Acta, 1503, 7-23.]). These results demonstrate that the measures taken to minimize radiation damage were sufficient to prevent radiation-induced reduction of the Mn ions in the OEC.

[Figure 4]
Figure 4
Mn K-edge absorption spectra and Mn Kα holograms of PSII. (a) Mn absorption spectra of PSII taken before and after the sXFH experiment and various reference Mn oxide powders. (b) Mn Kα hologram obtained at an incident X-ray energy of 6.565 keV (χH).

Fig. 4[link](b) shows the Mn Kα hologram integrated from nine PSII crystals. The hologram pattern appears to be dominated by the FS dips of the holographic oscillations (black in the color scale). The application of the low pass filter, the larger thermal vibration in protein crystals and the mostly light elements surrounding the Mn4CaO5 cluster (water molecules and amino-acid residues) suggest that the hologram consists mostly of holographic signal from scatterers near the Mn emitters.

To investigate how the larger thermal vibrations in proteins affect the holograms, the holograms were calculated with and without root-mean-square displacements of the scatterers due to static positional fluctuations within the crystal and dynamic fluctuations from thermal vibrations. The root-mean-square displacements were introduced into the hologram calculation by assuming isotropic vibrations of the scatterer atoms relative to the fixed emitter (〈u2Rel), which is represented by a Gaussian distribution with a standard deviation σ.

First, Fig. 5[link](a) shows the holographic oscillations calculated from a simple Mn–Mn dimer system using different relative root-mean-square displacement values, 〈u2Rel, whereas Fig. 5[link](b) shows the corresponding Debye–Waller factor or damping term. The holograms are composed of minima in the FS direction, surrounded by concentric higher-order interference rings, as shown in the inset of Fig. 5[link](b). As the 〈u2Rel values increase, the backscattering is suppressed and the FS dips become the prominent features of the hologram patterns.

[Figure 5]
Figure 5
Evolution of XFH patterns with increasing root-mean-square displacements. (a) Holographic oscillations as a function of scattering angle simulated from a simple Mn–Mn dimer at different relative root-mean-square displacement values, χH, and (b) their corresponding damping factors. Inset: spherical projection of the hologram calculated from the Mn–Mn dimer. (c)–(f) Evolution of the Mn Kα holograms calculated from the OEC of PSII using different χH values: (c) Mn Kα holograms simulated without thermal vibrations, and those with χH = 0.35 Å2 (d), χH = 0.60 Å2 (e) and χH = 2.00 Å2 (f).

Next, the Mn Kα holograms of PSII were calculated. Since PSII has a large unit cell (122.2 Å × 228.5 Å × 286.4 Å) and consists mostly of light elements, only the atoms within the Mn4CaO5 cluster are used for the calculations. The atomic positions were extracted from the PDB entry 3wu2 (Umena et al., 2011[Umena, Y., Kawakami, K., Shen, J. R. & Kamiya, N. (2011). Nature, 473, 55-60.]). The total calculated hologram is the superposition of the hologram patterns calculated for each of the 32 Mn atoms in a PSII unit cell. Each hologram was calculated using the Mn, Ca and O atoms located within the same Mn4CaO5 cluster as the emitter Mn (Fig. 1[link]).

Fig. 5[link](c) shows the Mn Kα hologram of PSII calculated without any thermal vibrations. With increasing 〈u2Rel values, the similarities with the experiment become more apparent. In Figs. 5[link](d)–5(f), a series of holograms were calculated with different 〈u2Rel values. As the 〈u2Rel value increases, the positive signals (bright yellow) in the hologram patterns are suppressed, and in Fig. 5[link](f), with an extreme 〈u2Rel = 2.00 Å2, the pattern is reduced to a large dark feature that bears strong resemblance to feature A in Fig. 4[link](b).

To determine the best 〈u2Rel value, an R-factor analysis is performed (Kuznetsov et al., 2014[Kuznetsov, M. V., Ogorodnikov, I. I. & Vorokh, A. S. (2014). Russ. Chem. Rev. 83, 13-37.]). The R-factor describes the agreement between the calculated and experimental hologram patterns, where a smaller R-factor means a better agreement. Fig. 6[link](a) shows the dependence of the R-factor on 〈u2Rel. The minimization of the R-factor shows that the hologram calculated with 〈u2Rel = 0.35 Å2 shows the best agreement with the experiment. This calculated hologram pattern is shown in Fig. 6[link](b) and the experimental hologram is shown again in Fig. 6[link](c) for comparison and discussion.

[Figure 6]
Figure 6
R-factor analysis between the calculated and experimental hologram patterns. (a) R-factor as a function of the relative mean-square displacements, 〈u2Rel, of the scatterer atoms used in the Mn Kα hologram calculations. (b) Calculated hologram with 〈u2Rel = 0.35 Å2 and (c) the experimental hologram. The dashed white lines are visual guides for features discussed in the text. Circular marks indicate the FS directions of Mn (purple) and Ca (green) scatterers calculated from each Mn emitter from PDB entry 3wu2.

There are several key features in the calculated hologram that are also clearly observed in the experimental hologram [highlighted by the dashed white lines in Figs. 6[link](b) and 6[link](c)]: region A, the large, dark and dog-bone-shaped region; region B, the bright feature observed at an azimuthal angle of 45°; and region C, a combination of dark and bright features.

For further interpretation of these features, the FS directions of all Mn and Ca scatterers were calculated from each Mn emitter and are superimposed on the lower images of Figs. 6[link](b) and 6[link](c). The angular positions of the circles on the images indicate the angular directions of the FS dips, and the relative sizes of the circles indicate the relative distances of the scatterer from the Mn emitters. These FS dips can be thought of as real-space projections of the positions of the scatterers around the emitter.

The origin of region A becomes apparent from the superimposed FS directions, where a dense collection of FS directions appear over the dark regions of the experimental hologram pattern in Fig. 6[link](c). Among these is the FS direction marked as (1) in Fig. 6[link](c), where a dark dip in the hologram pattern is observed in both the experiment and the calculation. This FS direction corresponds to an Mn4A–Mn3B emitter–scatterer pair where the Mn–Mn distance is 2.89 Å. The short emitter–scatter distance will result in a lower-frequency holographic signal and a larger FS dip (amplitude and width).

On the other hand, some FS directions appear as bright signals in both the experiment and the calculation, an example of which is the FS dip marked as (2). This occurs when the FS dip interferes with the first interference ring of an adjacent FS dip.

4. Discussion

The good agreement of the experimental results with the calculated hologram and calculated FS directions in Figs. 6[link](b) and 6[link](c) is a strong indication that the projection pattern obtained using sXFH contains structural information about the Mn4CaO5 cluster of PSII. Furthermore, the fact that the data were collected before any shifts in the Mn absorption spectra shows the tremendous potential of this method in studying the local structures of metal clusters in proteins before any radiation-induced damage.

In the current study, we interpreted the hologram patterns as real-space projections of the local structure around the emitter by analyzing the FS dips. This is similar to the tomographic interpretation of the directional fine structure in the absorption of white X-rays (Korecki & Materlik, 2001[Korecki, P. & Materlik, G. (2001). Phys. Rev. Lett. 86, 2333-2336.]; Korecki et al., 2006[Korecki, P., Tolkiehn, M., Novikov, D. V., Materlik, G. & Szymonski, M. (2006). Phys. Rev. B, 74, 184116.], 2009[Korecki, P., Novikov, D. V. & Tolkiehn, M. (2009). Phys. Rev. B, 80, 014119.]), also later termed white XFH (Dul & Korecki, 2012[Dul, D. T. & Korecki, P. (2012). New J. Phys. 14, 113044.]; Dąbrowski et al., 2013[Dąbrowski, K. M., Dul, D. T., Roszczynialski, T. P. & Korecki, P. (2013). Phys. Rev. B, 87, 064111.]). By employing polychromatic X-rays, a decrease in the coherence length results in the suppression of the higher-order interference fringes in the higher scattering angles, while the FS dip, which is largely energy independent, remains relatively unchanged. In this work, the combination of the large positional fluctuations in the PSII crystal and the application of the low pass filter on the hologram pattern also results in the suppression of the higher-order interference fringes while leaving the FS dip unchanged.

Furthermore, the FS dip approach used in this study can be thought of as the X-ray fluorescence analog of X-ray photoelectron diffraction (XPD) (Kuznetsov et al., 2014[Kuznetsov, M. V., Ogorodnikov, I. I. & Vorokh, A. S. (2014). Russ. Chem. Rev. 83, 13-37.]). In XPD, the local structure information is extracted by analyzing the forward-focusing peaks (FFP) in the angular distribution of the photoelectron intensity. Using photoelectrons with high kinetic energies allows the FFP to dominate the XPD patterns and a real-space projection of the scatterers around the emitter can be obtained. XPD has been extensively used in studying local structures in 2D layers, thin films and interfaces. Aside from the FS dips, other features such as Kossel lines (Bortel et al., 2016[Bortel, G., Faigel, G., Tegze, M. & Chumakov, A. (2016). J. Synchrotron Rad. 23, 214-218.]; Faigel et al., 2016[Faigel, G., Bortel, G. & Tegze, M. (2016). Sci. Rep. 6, 22904.]) or X-ray standing wave lines that may also appear in the hologram can also be used to extract structural information about the sample.

To improve the accuracy of hologram calculations used in this study, the effects of the positional fluctuations of the scatterers were directly introduced into the calculations by a Debye–Waller factor or damping term [Fig. 5[link](b) and equation (2)[link]]. Until now, the effects of atomic fluctuations have been introduced into the hologram calculations by introducing a distribution of the atomic positions in the atomic model used in the calculations. This can include random Gaussian distributions in the atomic positions (Hayashi et al., 2014[Hayashi, K., Happo, N. & Hosokawa, S. (2014). J. Electron Spectrosc. Relat. Phenom. 195, 337-346.]; Hosokawa et al., 2013[Hosokawa, S., Happo, N., Ozaki, T., Ikemoto, H., Shishido, T. & Hayashi, K. (2013). Phys. Rev. B, 87, 094104.]; Kimura et al., 2020[Kimura, K., Yamamoto, K., Hayashi, K., Tsutsui, S., Happo, N., Yamazoe, S., Miyazaki, H., Nakagami, S., Stellhorn, J. R., Hosokawa, S., Matsushita, T., Tajiri, H., Ang, A. K. R. & Nishino, Y. (2020). Phys. Rev. B, 101, 024302.]) or, alternatively, using atomic positions extracted from molecular dynamics simulations (Ang et al., 2021[Ang, A. K. R., Marumi, R., Sato-Tomita, A., Kimura, K., Happo, N., Akagi, K., Sasaki, T. & Hayashi, K. (2021). Phys. Rev. B, 103, 214106.]). Although these are good approximations that can result in the adequate reproduction of the atomic reconstruction, using a DWF or damping term directly in the hologram calculation is a simpler and more direct way to calculate the hologram pattern.

In the hologram calculations in Fig. 6[link], a root-mean-square displacement value of χH = 0.35 Å2 was used. This value obtained from the R-factor analysis in Fig. 6[link](a) is relative and the average 〈u2〉 of the atoms in the Mn4CaO5 cluster can be obtained using (Kimura et al., 2020[Kimura, K., Yamamoto, K., Hayashi, K., Tsutsui, S., Happo, N., Yamazoe, S., Miyazaki, H., Nakagami, S., Stellhorn, J. R., Hosokawa, S., Matsushita, T., Tajiri, H., Ang, A. K. R. & Nishino, Y. (2020). Phys. Rev. B, 101, 024302.])

[\langle{u^2}\rangle_{\rm Rel} = \langle{u^2}\rangle_{\rm Mn} + \langle{u^2}\rangle_{\rm Scatterer}, \eqno (3)]

resulting in an average 〈u2〉 = 0.29 Å2, or an average B-factor of 23.25 Å2 for all the atoms in Mn4CaO5. This is comparable with the average R-factors of the Mn atoms of the synchrotron radiation [B-factor ∼26 Å2 (Umena et al., 2011[Umena, Y., Kawakami, K., Shen, J. R. & Kamiya, N. (2011). Nature, 473, 55-60.])] or XFEL [B-factor ∼24 Å2 (Suga et al., 2015[Suga, M., Akita, F., Hirata, K., Ueno, G., Murakami, H., Nakajima, Y., Shimizu, T., Yamashita, K., Yamamoto, M., Ago, H. & Shen, J. R. (2015). Nature, 517, 99-103.])] structures, suggesting that, aside from dynamic fluctuations caused by thermal vibrations at 100 K, the atoms in the Mn4CaO5 cluster also have large positional fluctuations within the crystal. Though we were only able to extract average isotropic fluctuations in this work, we expect that in future work, the radial and angular positional fluctuations of Mn atoms can be extracted from more accurate hologram patterns, as this has already been demonstrated in XFH experiments on inorganic samples and organic crystals (Hosokawa et al., 2013[Hosokawa, S., Happo, N., Ozaki, T., Ikemoto, H., Shishido, T. & Hayashi, K. (2013). Phys. Rev. B, 87, 094104.]; Hayashi et al., 2014[Hayashi, K., Happo, N. & Hosokawa, S. (2014). J. Electron Spectrosc. Relat. Phenom. 195, 337-346.]; Ang et al., 2021[Ang, A. K. R., Marumi, R., Sato-Tomita, A., Kimura, K., Happo, N., Akagi, K., Sasaki, T. & Hayashi, K. (2021). Phys. Rev. B, 103, 214106.]; Kizaki et al., 2022[Kizaki, H., Hayashi, K., Lu, C., Happo, N., Hosokawa, S., Hidaka, S., Hayashi, S., Suzuki, M. & Uchitomi, N. (2022). Phys. Rev. B, 106, 064434.]). Once this is realized, valuable information to help understand the mechanisms of the oxidation process in PSII will be obtainable.

Aside from the Mn Kα hologram obtained in Fig. 4[link](b), the experimental configuration will also allow valence-selective XFH. By carefully tuning the incident X-ray energy based on the small shifts in the Mn K-edge of the different Mn ions, valence-sensitive hologram patterns can be recorded. We performed preliminary valence-selective XFH experiments on the same PSII samples by recording an additional hologram pattern at 6.551 keV [below the Mn(IV) K-edge, but above the Mn(III) edge]. Fig. 7[link] shows the high (χH) and low (χL) energy Mn Kα holograms taken at 6.565 keV and 6.551 keV, respectively. Both hologram patterns were taken from the same set of samples, and the only difference is the incident X-ray energy. The difference in the incident X-ray energies will result in a small change in the Mn(III):Mn(IV) contribution ratio, which means that, aside from small differences, the χH and χL holograms should be almost the same.

[Figure 7]
Figure 7
Mn Kα hologram obtained at (a) a high incident X-ray energy of 6.565 keV (χH) and (b) a low incident X-ray energy of 6.551 keV (χL).

At 6.551 keV, the absorption and the subsequent fluorescence from Mn(IV) is suppressed, resulting in a hologram pattern that is mostly from the Mn(III) ions. The total hologram in Fig. 7[link](a), taken at an incident X-ray energy of 6.565 keV [above the Mn K-edge of both Mn(III) and Mn(IV)], χH, is a superposition of hologram contributions from all Mn ions, χMn(III) and χMn(IV). Using the known Mn valence distribution of PSII S1 [2 Mn(III) and 2 Mn(IV) (Suga et al., 2015[Suga, M., Akita, F., Hirata, K., Ueno, G., Murakami, H., Nakajima, Y., Shimizu, T., Yamashita, K., Yamamoto, M., Ago, H. & Shen, J. R. (2015). Nature, 517, 99-103.])], the total hologram can be expressed as

[{\chi_{\rm H}} = {{{2{\chi_{\rm Mn(III)}} + 2{\chi_{\rm Mn(IV)}}} }\over{4}}. \eqno (4)]

At the lower incident X-ray energy of 6.551 keV, both χMn(III) and χMn(IV) signals are suppressed due to the lower X-ray absorption, and the hologram χL can be expressed as

[\chi_{\rm L} = {{{2\alpha {\chi_{\rm Mn(III)}} + 2\beta {\chi_{\rm Mn(IV)}}} }\over{2\left({\alpha + \beta } \right)}}, \eqno (5)]

where the α and β parameters represent the decrease in the absorption spectra of χMn(III) and χMn(IV) relative to the absorption at 6.565 keV. From equations (3)[link] and (4)[link], χMn(III) can be numerically extracted using

[{{{2\alpha - 2\beta }} \over \beta }{\chi_{\rm Mn(III)}} = {{{2\alpha + 2\beta }} \over \beta }{\chi_{\rm L}} - 4{\chi_{\rm H}}, \eqno (6)]

and subsequently χMnIV can be obtained using χH, χMn(III) and equation (4)[link].

From the normalized intensity of the absorption spectra in Fig. 4[link](a), the α and β parameters are estimated to be 0.75 and 0.55, respectively. Using these parameters, the χMn(III) and χMn(IV) hologram patterns were extracted and are shown in Figs. 8[link](a) and 8[link](b).

[Figure 8]
Figure 8
Valence-sensitive Mn Kα hologram patterns χMn(III) and χMn(IV). (aχMn(III) and (b) χMn(IV) hologram patterns extracted numerically from the experimental holograms using (α, β) = (0.75, 0.55) as determined from Mn absorption spectra. Calculated (c) χMn(III) and (d) χMn(IV) holograms with 〈u2Rel = 0.35 Å2. The FS directions of neighboring Mn and Ca atoms were calculated from either χMn(III) (Mn1D and Mn4A sites) or χMn(IV) (Mn2C and Mn3B sites) emitters and are indicated by the purple (Mn) or green (Ca) circles.

To calculate the hologram patterns for Mn(III) or Mn(IV), only 16 clusters were used, with Mn emitters at the known sites of either the Mn(III) or Mn(IV) ions. Using the same χH parameter used for the total hologram pattern in Fig. 6[link](b), the hologram patterns were calculated for Mn(III) and Mn(IV) emitters in PSII and are shown in Figs. 8[link](c) and 8[link](d). The FS directions are once again superimposed on the holograms.

From the calculations of the hologram patterns and the FS directions, a distinct difference in the χMn(III) and χMn(IV) holograms can be observed: the large dark feature (region A) in χMn(IV) that results from the dense arrangement of the FS directions. For the χMn(III) hologram, there is no clear distinct feature in the hologram, the FS directions are evenly distributed.

This difference between χMn(III) and χMn(IV) can also be observed in the experimentally extracted holograms. The dark region in χMn(IV), marked as region D in Fig. 8[link](b), is clearly observed in both the experimentally extracted and the calculated χMn(IV) hologram patterns. On the other hand, the experimentally extracted χMn(III), like the calculations, shows no distinct features. These results show that, while the agreement between the experiment and the calculations are not clear as shown in the the total hologram in Fig. 6[link], the distinct features in the hologram patterns can allow differentiation of the χMn(III) and χMn(IV) hologram patterns. In the case of the S2 state, where only one Mn(III) remains in the OEC, the distinction between χMn(III) and χMn(IV) will be clearer, making XFH analysis more straightforward. Thus, analysis of the χMn(III) and χMn(IV) hologram patterns can be an alternative approach to experimentally determining which Mn site is the remaining Mn(III).

In our experiment, a careful compromise had to be made between the total X-ray dose on the sample, the total photons collected by the 2D detector and the limited experiment times at synchrotron radiation facilities. Even under such limiting experimental conditions, we could already reach a certain agreement between the experimental and calculated holograms, as shown and discussed in the preceding sections. However, further optimization of the experiment would be necessary to obtain more accurate valence-selective holograms, because they must extract the small difference in the holograms taken at two different energies. By increasing the total photons collected during the experiment, either through longer exposure times, larger single-crystal samples or by measuring samples, the quality of the experimental hologram might be significantly improved. Alternatively, novel approaches to the extraction of the χMnIII and χMnIV holograms from the high- and low-energy holograms recorded in the experiment can also be developed. The subtraction in equation (6)[link] can result in the propagation of uncertainty, and this can be seen in the higher contrast scale used in Fig. 8[link].

Several valence-selective XFH experiments have already been reported for inorganic crystals: a direct-imaging experiment using the same 2D hybrid pixel detector on magnetite (Fe3O4) (Ang et al., 2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]) and conventional XFH experiments using 0D detectors (avalanche photodiode or SDD) on yttrium oxide thin films (YO/Y2O3) and YbInCu4 (Stellhorn et al., 2017[Stellhorn, J. R., Hosokawa, S., Happo, N., Tajiri, H., Matsushita, T., Kaminaga, K., Fukumura, T., Hasegawa, T. & Hayashi, K. (2017). J. Appl. Cryst. 50, 1583-1589.]; Hosokawa et al., 2019[Hosokawa, S., Happo, N., Matsushita, T., Stellhorn, J. R., Kimura, K. & Hayashi, K. (2019). Jpn. J. Appl. Phys. 58, 120601.], 2020[Hosokawa, S., Happo, N., Hayashi, K., Kimura, K., Matsushita, T., Stellhorn, J. R., Mizumaki, M., Suzuki, M., Sato, H. & Hiraoka, K. (2020). J. Phys. Soc. Jpn, 89, 034603.]). In the experiments on Fe3O4, the difference in the Kossel line features between the high- and low-energy hologram patterns clearly show the valence selectivity (Ang et al., 2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]). Atomic reconstructions of these valence-selective hologram patterns have resulted in mixed results. Ang et al. (2018[Ang, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.]) numerically extracted the Fe(II) hologram using an equation similar to equation (6)[link], and, although the reconstruction can be distinguished from the reconstruction obtained from the total [Fe(II) and Fe(III)], the quality of the results shows that it will be difficult to use the same approach on samples with larger and more complicated structures. Hosokawa et al. (2020[Hosokawa, S., Happo, N., Hayashi, K., Kimura, K., Matsushita, T., Stellhorn, J. R., Mizumaki, M., Suzuki, M., Sato, H. & Hiraoka, K. (2020). J. Phys. Soc. Jpn, 89, 034603.]) used a different approach to extract the Yb(II) hologram, where the Yb(II) hologram was recorded at an energy where only the Yb 2p3/2 electrons in Yb(II) were excited and emitted fluorescent X-rays. The reconstructions obtained showed the expected f.c.c. structure around Yb(III), whereas the reconstructed atomic image around Yb(II) showed large positional fluctuations. In principle, this would be a reasonable approach to obtain valence-sensitive holograms. However, for samples with low concentration of metal emitters, or for radiation-sensitive samples such as metalloproteins, it will be experimentally difficult to obtain valence-selective holograms using this approach.

5. Conclusions

In this work, we develop a novel approach for obtaining the local atomic structure around metallic clusters in protein crystals. By employing a 2D hybrid pixel detector to directly image the hologram patterns, a statistically significant hologram pattern was obtained in serial data acquisition mode, similar to those used in serial protein crystallography. The Mn Kα hologram was recorded from PSII crystals prior to the onset of radiation-damage-induced reduction of the Mn4CaO5 clusters. The good agreement between the calculated and experimental hologram patterns and the analysis of the FS directions show that the recorded hologram pattern can be treated as a real-space projection of the atoms around the Mn emitters in the Mn4CaO5 clusters.

Furthermore, our new approach also allows valence-sensitive XFH experiments on protein crystals. Our preliminary results show that, by tuning the incident X-ray energy, the holograms from Mn(III) and Mn(IV) were extracted, and some distinct features observed in both the experimental and the calculated holograms allow some differentiation between them. Though additional data processing and more advanced reconstruction algorithms will be necessary to selectively reconstruct the local atomic structure around either Mn(III) or Mn(IV) from the holograms, the analysis of the FS patterns in the hologram, in combination with R-factor analysis, may allow the determination of which Mn sites each Mn ion occupies. This paves the way for future valence-sensitive XFH studies of other metal clusters in proteins and, considering that the valence states of these metal clusters play key roles in the functions of these proteins, there can be significant developments unlocked by further refinement of this approach.

Footnotes

Current address: Toyota Technological Institute, Nagoya, Japan.

Funding information

This work was supported by the Japan Society for the Promotion of Science (JSPS) Grants-in-Aid for Scientific Research (grant nos. JP16KT0058; JP17H06372) and Grant-in-Aid for Scientific Research on Innovative Areas `3D Active-Site Science' (grant nos. 17H05219; 26105005; 26105006) and Grant-in-Aid for Transformative Research Areas (A) `Hyper-Ordered Structures Science' (grant nos. 20H05878; 20H05881; 20H05883; 20H05884). The XFH experiments were performed in the SPring-8 Synchrotron Radiation Facility in Japan with the approval of JASRI (proposal nos. 2015A1071; 2016A1391; 2016A2514; 2016B1328; 2017A2529; 2017A1137; 2017B1355; 2018A1097; 2018A2519; 2018B1153; 2019A1230; 2019B1149; 2019B1151).

References

First citationAng, A. K. R., Marumi, R., Sato-Tomita, A., Kimura, K., Happo, N., Akagi, K., Sasaki, T. & Hayashi, K. (2021). Phys. Rev. B, 103, 214106.  Web of Science CrossRef Google Scholar
First citationAng, A. K. R., Matsushita, T., Hashimoto, Y., Happo, N., Yamamoto, Y., Mizuguchi, M., Sato-Tomita, A., Shibayama, N., Sasaki, Y. C., Kimura, K., Taguchi, M., Daimon, H. & Hayashi, K. (2018). Phys. Status Solidi B, 255, 1800100.  Web of Science CrossRef Google Scholar
First citationAng, A. K. R., Sato-Tomita, A., Shibayama, N., Umena, Y., Happo, N., Marumi, R., Kimura, K., Matsushita, T., Akagi, K., Sasaki, T., Sasaki, Y. C. & Hayashi, K. (2020). Jpn. J. Appl. Phys. 59, 010505.  Web of Science CrossRef Google Scholar
First citationAsada, M., Nagashima, H., Koua, F. H. M., Shen, J. R., Kawamori, A. & Mino, H. (2013). Biochim. Biophys. Acta, 1827, 438–445.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBallabriga, R., Campbell, M. & Llopart, X. (2018). Nucl. Instrum. Methods Phys. Res. A, 878, 10–23.  Web of Science CrossRef CAS Google Scholar
First citationBarton, J. J. (1991). Phys. Rev. Lett. 67, 3106–3109.  CrossRef PubMed CAS Web of Science Google Scholar
First citationBortel, G., Faigel, G., Tegze, M. & Angelov, B. (2019). J. Synchrotron Rad. 26, 170–174.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBortel, G., Faigel, G., Tegze, M. & Chumakov, A. (2016). J. Synchrotron Rad. 23, 214–218.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationChu, H. A., Hillier, W. & Debus, R. J. (2004). Biochemistry, 43, 3152–3166.  Web of Science CrossRef PubMed CAS Google Scholar
First citationChu, H. A., Hillier, W., Law, N. A. & Babcock, G. T. (2001). Biochim. Biophys. Acta, 1503, 69–82.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCox, N., Rapatskiy, L., Su, J. H., Pantazis, D. A., Sugiura, M., Kulik, L., Dorlet, P., Rutherford, A. W., Neese, F., Boussac, A., Lubitz, W. & Messinger, J. (2011). J. Am. Chem. Soc. 133, 3635–3648.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDąbrowski, K. M., Dul, D. T., Roszczynialski, T. P. & Korecki, P. (2013). Phys. Rev. B, 87, 064111.  Google Scholar
First citationDaimon, H. (2018). J. Phys. Soc. Jpn, 87, 061001.  Web of Science CrossRef Google Scholar
First citationDaimon, H. (2020). Jpn. J. Appl. Phys. 59, 010504.  Web of Science CrossRef Google Scholar
First citationDau, H., Grundmeier, A., Loja, P. & Haumann, M. (2008). Philos. Trans. R. Soc. B, 363, 1237–1244.  Web of Science CrossRef CAS Google Scholar
First citationDau, H. & Haumann, M. (2008). Coord. Chem. Rev. 252, 273–295.  Web of Science CrossRef CAS Google Scholar
First citationDebus, R. J., Strickler, M. A., Walker, L. M. & Hillier, W. (2005). Biochemistry, 44, 1367–1374.  Web of Science CrossRef PubMed CAS Google Scholar
First citationDul, D. T. & Korecki, P. (2012). New J. Phys. 14, 113044.  Web of Science CrossRef Google Scholar
First citationFadley, C. S. & Len, P. M. (1996). Nature, 380, 27–28.  CrossRef CAS Web of Science Google Scholar
First citationFaigel, G., Bortel, G. & Tegze, M. (2016). Sci. Rep. 6, 22904.  Web of Science CrossRef PubMed Google Scholar
First citationFaigel, G. & Tegze, M. (1999). Rep. Prog. Phys. 62, 355–393.  Web of Science CrossRef CAS Google Scholar
First citationFerreira, K. N., Iverson, T. M., Maghlaoui, K., Barber, J. & Iwata, S. (2004). Science, 303, 1831–1838.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGarman, E. F. (2010). Acta Cryst. D66, 339–351.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGlöckner, C., Kern, J., Broser, M., Zouni, A., Yachandra, V. & Yano, J. (2013). J. Biol. Chem. 288, 22607–22620.  Web of Science PubMed Google Scholar
First citationGuskov, A., Kern, J., Gabdulkhakov, A., Broser, M., Zouni, A. & Saenger, W. (2009). Nat. Struct. Mol. Biol. 16, 334–342.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHayashi, K., Happo, N. & Hosokawa, S. (2014). J. Electron Spectrosc. Relat. Phenom. 195, 337–346.  Web of Science CrossRef CAS Google Scholar
First citationHayashi, K., Happo, N., Hosokawa, S., Hu, W. & Matsushita, T. (2012). J. Phys. Condens. Matter, 24, 093201.  Web of Science CrossRef PubMed Google Scholar
First citationHayashi, K. & Korecki, P. (2018). J. Phys. Soc. Jpn, 87, 061003.  Web of Science CrossRef Google Scholar
First citationHosokawa, S., Happo, N., Hayashi, K., Kimura, K., Matsushita, T., Stellhorn, J. R., Mizumaki, M., Suzuki, M., Sato, H. & Hiraoka, K. (2020). J. Phys. Soc. Jpn, 89, 034603.  Web of Science CrossRef Google Scholar
First citationHosokawa, S., Happo, N., Matsushita, T., Stellhorn, J. R., Kimura, K. & Hayashi, K. (2019). Jpn. J. Appl. Phys. 58, 120601.  Web of Science CrossRef Google Scholar
First citationHosokawa, S., Happo, N., Ozaki, T., Ikemoto, H., Shishido, T. & Hayashi, K. (2013). Phys. Rev. B, 87, 094104.  Web of Science CrossRef Google Scholar
First citationKabsch, W. (2010). Acta Cryst. D66, 125–132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKamiya, N. & Shen, J.-R. (2003). Proc. Natl Acad. Sci. 100, 98–103.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKang, L., Akagi, K., Hayashi, K. & Sasaki, T. (2017). Phys. Rev. B, 95, 214106.  Web of Science CrossRef Google Scholar
First citationKawakami, K. & Shen, J. R. (2018). Methods Enzymol. 613, 1–16.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKimura, K., Yamamoto, K., Hayashi, K., Tsutsui, S., Happo, N., Yamazoe, S., Miyazaki, H., Nakagami, S., Stellhorn, J. R., Hosokawa, S., Matsushita, T., Tajiri, H., Ang, A. K. R. & Nishino, Y. (2020). Phys. Rev. B, 101, 024302.  Web of Science CrossRef Google Scholar
First citationKizaki, H., Hayashi, K., Lu, C., Happo, N., Hosokawa, S., Hidaka, S., Hayashi, S., Suzuki, M. & Uchitomi, N. (2022). Phys. Rev. B, 106, 064434.  Web of Science CrossRef Google Scholar
First citationKorecki, P. & Materlik, G. (2001). Phys. Rev. Lett. 86, 2333–2336.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKorecki, P., Novikov, D. V. & Tolkiehn, M. (2009). Phys. Rev. B, 80, 014119.  Web of Science CrossRef Google Scholar
First citationKorecki, P., Tolkiehn, M., Novikov, D. V., Materlik, G. & Szymonski, M. (2006). Phys. Rev. B, 74, 184116.  Web of Science CrossRef Google Scholar
First citationKulik, L. V., Epel, B., Lubitz, W. & Messinger, J. (2007). J. Am. Chem. Soc. 129, 13421–13435.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKuznetsov, M. V., Ogorodnikov, I. I. & Vorokh, A. S. (2014). Russ. Chem. Rev. 83, 13–37.  Web of Science CrossRef Google Scholar
First citationMatsushita, T. (2018). Phys. Status Solidi B, 255, 1800091.  Web of Science CrossRef Google Scholar
First citationMatsushita, T., Muro, T., Matsui, F., Happo, N. & Hayashi, K. (2020). Jpn. J. Appl. Phys. 59, 020502.  Web of Science CrossRef Google Scholar
First citationMatsushita, T., Muro, T., Yokoya, T., Terashima, K., Kato, Y., Matsui, H., Maejima, N., Hashimoto, Y. & Matsui, F. (2020). Phys. Status Solidi B, 257, 2000117.  Web of Science CrossRef Google Scholar
First citationPlackett, R., Horswell, I., Gimenez, E. N., Marchal, J., Omar, D. & Tartoni, N. (2013). J. Instrum. 8, C01038.  Web of Science CrossRef Google Scholar
First citationRobblee, J. H., Cinco, R. M. & Yachandra, V. K. (2001). Biochim. Biophys. Acta, 1503, 7–23.  Web of Science CrossRef PubMed CAS Google Scholar
First citationRoelofs, T. A., Liang, W., Latimer, M. J., Cinco, R. M., Rompel, A., Andrews, J. C., Sauer, K., Yachandra, V. K. & Klein, M. P. (1996). Proc. Natl Acad. Sci. USA, 93, 3335–3340.  CrossRef PubMed CAS Web of Science Google Scholar
First citationSano, K., Sasaki, T., Yoneyama, N. & Kobayashi, N. (2010). Phys. Rev. Lett. 104, 217003.  Web of Science CrossRef PubMed Google Scholar
First citationSato-Tomita, A., Ang, A. K. R., Kimura, K., Marumi, R., Happo, N., Matsushita, T., Park, S., Shibayama, N., Sasaki, Y. C. & Hayashi, K. (2022). Biochem. Biophys. Res. Commun. 635, 277–282.  Web of Science CAS PubMed Google Scholar
First citationSato-Tomita, A., Shibayama, N., Happo, N., Kimura, K., Okabe, T., Matsushita, T., Park, S. Y., Sasaki, Y. C. & Hayashi, K. (2016). Rev. Sci. Instrum. 87, 063707.  Web of Science PubMed Google Scholar
First citationShen, J. R. (2015). Annu. Rev. Plant Biol. 66, 23–48.  Web of Science CrossRef CAS PubMed Google Scholar
First citationStellhorn, J. R., Hosokawa, S., Happo, N., Tajiri, H., Matsushita, T., Kaminaga, K., Fukumura, T., Hasegawa, T. & Hayashi, K. (2017). J. Appl. Cryst. 50, 1583–1589.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationStich, T. A., Yeagle, G. J., Service, R. J., Debus, R. J. & Britt, R. D. (2011). Biochemistry, 50, 7390–7404.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSuga, M., Akita, F., Hirata, K., Ueno, G., Murakami, H., Nakajima, Y., Shimizu, T., Yamashita, K., Yamamoto, M., Ago, H. & Shen, J. R. (2015). Nature, 517, 99–103.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSuzuki, M., Kawamura, N., Mizumaki, M., Terada, Y., Uruga, T., Fujiwara, A., Yamazaki, H., Yumoto, H., Koyama, T., Senba, Y., Takeuchi, T., Ohashi, H., Nariyama, N., Takeshita, K., Kimura, H., Matsushita, T., Furukawa, Y., Ohata, T., Kondo, Y., Ariake, J., Richter, J., Fons, P., Sekizawa, O., Ishiguro, N., Tada, M., Goto, S., Yamamoto, M., Takata, M. & Ishikawa, T. (2013). J. Phys. Conf. Ser. 430, 012017.  CrossRef Google Scholar
First citationTanaka, A., Fukushima, Y. & Kamiya, N. (2017). J. Am. Chem. Soc. 139, 1718–1721.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTegze, M. & Faigel, G. (1996). Nature, 380, 49–51.  CrossRef CAS Web of Science Google Scholar
First citationUmena, Y., Kawakami, K., Shen, J. R. & Kamiya, N. (2011). Nature, 473, 55–60.  Web of Science CrossRef CAS PubMed Google Scholar
First citationYachandra, V. K., Sauer, K. & Klein, M. P. (1996). Chem. Rev. 96, 2927–2950.  CrossRef PubMed CAS Web of Science Google Scholar
First citationYano, J., Kern, J., Irrgang, K.-D., Latimer, M. J., Bergmann, U., Glatzel, P., Pushkar, Y., Biesiadka, J., Loll, B., Sauer, K., Messinger, J., Zouni, A. & Yachandra, V. K. (2005). Proc. Natl Acad. Sci. 102, 12047–12052.  Web of Science CrossRef PubMed CAS Google Scholar
First citationYano, J., Pushkar, Y., Glatzel, P., Lewis, A., Sauer, K., Messinger, J., Bergmann, U. & Yachandra, V. (2005). J. Am. Chem. Soc. 127, 14974–14975.  Web of Science CrossRef PubMed CAS Google Scholar
First citationZeldin, O. B., Gerstel, M. & Garman, E. F. (2013). J. Appl. Cryst. 46, 1225–1230.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZouni, A., Witt, H. T., Kern, J., Fromme, P., Krauss, N., Saenger, W. & Orth, P. (2001). Nature, 409, 739–743.  Web of Science CrossRef PubMed CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

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