Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2021 Nov 5;12(45):10969–10974. doi: 10.1021/acs.jpclett.1c03411

Suppression of the Phase Coexistence of the fcc–fct Transition in Hafnium-Hydride Thin Films

Lars J Bannenberg †,*, Herman Schreuders , Hyunjeong Kim , Kouji Sakaki , Shigenobu Hayashi §, Kazutaka Ikeda , Toshiya Otomo , Kohta Asano , Bernard Dam
PMCID: PMC8607497  PMID: 34738818

Abstract

graphic file with name jz1c03411_0005.jpg

Metal hydrides may play a paramount role in a future hydrogen economy. While most applications are based on nanostructured and confined materials, studies considering the structural response of these materials to hydrogen concentrate on bulk material. Here, using in situ in- and out-of-plane X-ray diffraction and reflectometry, we study the fcc ↔ fct transition in Hf thin films, an optical hydrogen-sensing material. We show that the confinement of Hf affects this transition: compared to bulk Hf, the transition is pushed to a higher hydrogen-to-metal ratio, the tetragonality of the fct phase is reduced, and phase coexistence is suppressed. These nanoconfinement effects ensure the hysteresis-free response of hafnium to hydrogen, enabling its remarkable performance as a hydrogen-sensing material. In a wider perspective, the results highlight the profound influences of the nanostructuring and nanoconfinement of metal hydrides on their structural response to hydrogen with a significant impact on their applicability in future devices.


Metal hydrides are a class of materials that may play an important role in a hydrogen-powered economy. Traditionally considered to be hydrogen storage materials,13 other applications such as in hydrogen-purifying membranes,4,5 switchable mirrors,6,7 fuel cells,8 and especially hydrogen sensors3,914 have attracted attention recently. In general, the x (hydrogen-to-metal ratio) – T (temperature) phase diagrams of metal hydrides are relatively complex, involving a combination of body- or face-centered cubic, tetragonal, orthorhombic, or hexagonal metal hydride phases that may or may not coexist.100 A detailed understanding of (the transitions between) these phases is of vital importance because it determines, together with the entropy and enthalpy of formation, the performance of metal hydrides in real-life applications.

Although many applications of metal hydrides rely on nanostructered materials such as thin films and nanoparticles, most studies determining the phase diagram and nature of phase transitions focused on the bulk. Nanostructuring may significantly alter the properties of metal hydrides, changing their stability and the occurrence of phases and phase transitions (see, e.g., refs (1523)). Such effects may arise from an increased surface-to-volume ratio or, especially for thin films, geometric constraints on the possibility to expand volumetrically and clamping to the supporting substrate. In particular, for hydrogen-sensing applications, metal hydrides are typically structured as thin films or nanoparticles,3,1114,24 and understanding the occurrence and nature of phase transitions is of vital importance to designing hysteresis-free hydrogen sensing materials that are stable over time with repeated exposure to hydrogen.

Hafnium (Hf) is a promising optical hydrogen-sensing material that features an extraordinary hysteresis-free sensing range spanning over 6 orders of magnitude in hydrogen pressure.3,25,26 In the bulk, similar to other group IV elements titanium (Ti)2729 and zirconium (Zr),30,31 a large two-phase region separates the hexagonal-close-packed (hcp) solid solution at low x in HfHx from the face-centered-cubic (fcc) phase (x ≳ 1.5), where hydrogen occupies the interstitial tetrahedral sites (4f).27,32,33 Upon further hydrogenation, a transformation to the face-centered tetragonal (fct) phase is observed, in which the fcc lattice is compressed along the c axis. The attractiveness of hafnium as a hydrogen-sensing material stems from the fact that hafnium thin films show a highly reproducible change in optical transmission in response to a hydrogen exposure ranging over 6 orders of magnitude in hydrogen pressure, which can be used to accurately determine the hydrogen pressure.3,13,25,26 The sensing range corresponds to a hysteresis-free hydrogenation of Hf from about HfH1.5 to HfH1.98.25,26 The appearance of a hysteresis-free sensing range is most remarkable, as for bulk Hf a first-order fcc ↔ fct phase transition involving hysteresis is reported between HfH1.78 and HfH1.86.27,33 As such, it is unclear whether the phase transition is suppressed in thin films or whether nanoconfinement pushes the transition toward second order, thereby eliminating the hysteresis. Studying this and other phase transitions in thin films is typically complicated by the strong epi-texturing of the films: while powder X-ray diffraction (XRD) provides information on multiple lattice reflections, typical Bragg–Brentano XRD measurements of thin films provide limited information on the out-of-plane direction only: in the case of hafnium-hydride only the (111) and (222) reflections can be observed, preventing the observation of the fcc ↔ fct phase transition.

Here, we combine in situ out-of-plane, in-plane XRD and X-ray reflectometry (XRR) to elucidate the nature of the fcc ↔ fct phase transition in Hf thin films. We unambiguously identify the fcc ↔ fct phase transition in Hf thin films, but at a much larger value of x than in the bulk and with a smaller degree of tetragonality. Different from bulk HfHx, no sign of phase coexistence is observed, suggesting that the confinement of the film suppresses the phase coexistence. In a more general perspective, these results highlight the profound influence of nanoconfinement on the presence and nature of phase transitions in thin films and other nanostructured metal hydrides.

Before discussing the results on thin films, we first consider the nature of powder Hf to confirm the first-order nature of the fcc ↔ fct phase transition in the bulk. Figure 1 displays X-ray and neutron diffraction results as well as 2H magic angle spinning nuclear magnetic resonance (MAS NMR) spectroscopy results of HfDx in three different deuteration states—HfD1.66, HfD1.78, and HfD1.95—obtained after pressure–composition isotherm measurements at 300 °C (Figure S3). As previous results indicate that the hafnium–hydrogen and hafnium–deuterium phase diagrams are very similar,33 we use deuterium instead of hydrogen to suppress the incoherent background in the neutron diffraction experiments and to enable a higher resolution in NMR.

Figure 1.

Figure 1

Diffraction and NMR results on powder HfDx at three different deuteration states: HfD1.66, HfD1.78, and HfD1.95. (a) X-ray diffraction patterns (Cu Kα, λ = 0.1542 nm). The continuous lines show the fit to the data, while the additional lines underneath the diffraction patterns indicate the corresponding residuals of the fits. (b) Neutron diffraction patterns obtained at NOVA located at the Japan Proton Accelerator Research Complex (J-PARC). The insets shows the split of the {200} reflection at HfD1.66 to (002) and (110) at HfD1.95, while the coexistence of these three reflections is observed at HfD1.78. (c) 2H magic angle spinning nuclear magnetic resonance (MAS NMR) spectra.

The X-ray and neutron diffraction patterns (Figure 1(a,b)) of HfD1.66 and HfD1.95 are well described by single-phase models of fcc (Fm3̅m) and fct (I4/mmm), respectively, while the pattern of HfD1.78 can be fitted using only a two-phase model where the fcc and fct phases coexist. While for fitting the XRD pattern the contribution of deuterium is assumed to be negligible, neutron diffraction confirms that deuterium atoms occupy the tetrahedral sites. The lattice parameters that are refined by neutron diffraction are a = 0.487648(6) nm for HfD1.66 and a = 0.487805(6) nm and c = 0.434134(10) nm (c/a = 0.89) for HfD1.95. The refinement of the pattern for HfD1.78 using the two-phase model provides a = 0.467416(33) nm for the fcc phase and a = 0.481491(12) nm and c = 0.443678(26) nm (c/a = 0.92) for the fct phase, respectively. The obtained weight fraction of these two phases is fcc/fct = 39:61.

The 2H MAS NMR spectra confirm the single-phase behavior for HfD1.66 and HfD1.95 and the phase coexistence for HfD1.78. The spectra of HfD1.66 and HfD1.95 show a single component (Figure 1(c)), which have negative Knight shifts of −9.9 and −33.9 ppm, respectively. The Knight shift is attributed to the density of states at the Fermi energy: HfD1.66 is relatively more insulating than HfD1.95, consistent with the reduction of the optical transmission upon increased hydrogenation (Figure 4(d,e) and refs (25) and (26)). The spectrum of HfD1.78 consists of two contributions due to the phase coexistence. As such, it was fitted using two Gaussian functions with an area ratio of fcc/fct = 4:6. Assuming that the deuterium contents of the fcc and fct phases are D/Hf = 1.75 and 2, respectively, and using the fact that the intensity of NMR signals is proportional to the molar ratio of deuterium, the weight fraction is estimated to be fcc/fct = 43:57, which is close to the one obtained by neutron diffraction. Most importantly, the phase coexistence observed for HfD1.78 by both X-ray and neutron diffraction as well as NMR unambiguously confirms the first-order nature of the fcc–fct phase transition.

Figure 4.

Figure 4

Reflectometry and optical transmission measurements on the 60 nm Hf films capped with a 10 nm Pd layer and with a 3 nm Ti adhesion layer. (a) In situ X-ray reflectometry (XRR) measurements at T = 120 °C and the partial hydrogen pressures PH2 indicated. During the experiment, the hydrogen pressure was stepwise increased. The dots indicate the measurement points, while the continuous lines indicate the fits to the data. (b) Partial hydrogen pressure dependence of the Hf layer thickness at T = 120 °C derived from fitting the XRR data that was collected by stepwise increasing and decreasing the partial hydrogen pressure. (c) Hydrogen content at T = 120 °C of 40 nm Hf films capped with a 10 nm Pd layer as deduced from neutron reflectometry measurements. Data were obtained from ref (26). (d) Changes in the white light optical transmission where the film was exposed to various increasing and decreasing pressure steps of PH2 = 0.22–9.5 Pa at T = 120 °C. The optical transmission Inline graphic is measured relative to the transmission of the film after unloading in air (Inline graphic). The dashed lines indicate levels of the same transmission and pressure. (e) Partial hydrogen pressure dependence of changes in the white light optical transmission at T = 90, 120, and 150 °C.

Turning to the results of the thin films, Figure 2 displays the out-of-plane and in-plane diffraction measurements obtained for different partial hydrogen pressures at 120 °C. The thin films are produced by magnetron sputtering on fused quartz substrates and are composed of a 4 nm Ti adhesion, a 60 nm Hf, and a 10 nm Pd capping layer to prevent oxidation and promote hydrogen dissociation. As the films are highly textured, the out-of-plane results reveal only the (111) peak (Figures S4 Figure S5). As such, we performed in-plane diffraction by tilting the sample in the direction perpendicular to the beam by ϕ = 35, 55, and 70° in order to monitor the hydrogen pressure dependence of the (200), (220), and in-plane (111) reflections.

Figure 2.

Figure 2

In situ out-of-plane and in-plane XRD measurements (Cu Kα, λ = 0.1542 nm) on the 60 nm Hf films capped with a 10 nm Pd layer and with a 3 nm Ti adhesion layer at T = 120 °C and the partial hydrogen pressures PH2 indicated. During the experiment, the hydrogen pressure was increased stepwise. (a) Out-of-plane diffraction and in-plane diffraction with the sample tilted in the direction perpendicular to the X-ray beam by (b) χ = 35°, (c) χ = 55°, and (d) χ = 70°. The continuous lines show fits of pseudo-Voigt function(s) to the data from which the d spacing (Figure 3) has been derived.

Most importantly, the diffraction patterns show the hallmarks of the fcc ↔ fct transformation: comparing the diffraction patterns in the unloaded state and the highest hydrogen pressure measured, PH2 = 613 Pa, indicates that the (111) reflection (Figure 2(a,d)) remains unsplit, both in the in- and out-of-plane directions, while the (200) (Figure 2(b)) and (220) (Figure 2(c)) reflections, both oriented in-plane, each split into two peaks. On the basis of the peak positions, we observe a substantial expansion of the a axes from 0.472 nm in the unloaded state to 0.479 nm in the fully hydrated state. At the same time, the c axis contracts to 0.457 nm, resulting in c/a ≈ 0.95. As such, the volume of the unit cell in the fully hydrogenated state of V = 0.0524 nm3 is approximately the same as for the bulk (V = 0.0516 nm3), while the compression of the c axis and expansion of the a axes are much less pronounced in the thin films than in the bulk (c/a ≈ 0.90).

The fcc ↔ fct transition at PH2 = 10+1 Pa is accompanied by a considerable reduction in volume. The out-of-plane diffraction results indicate a reduction of the out-of-plane d111 spacing by about 1% (Figure 3(a)) when fcc ↔ fct occurs. Differently, the in-plane d111 spacing remains unaltered and thus does not change upon increased hydrogenation. Consistent with this, the XRR measurements of Figure 4(a,b) indicate that the Hf layer thickness decreases by about 1%, thus indicating that the reduction of the d111 spacing is translated completely in the out-of-plane direction.

Figure 3.

Figure 3

Partial hydrogen pressure dependence of the in situ out-of-plane and in-plane d spacing of 60 nm Hf films capped with a 10 nm Pd layer and with a 3 nm Ti adhesion layer at T = 120 °C. During the experiment, the hydrogen pressure was stepwise increased and decreased. The d spacing was obtained by fitting pseudo-Voigt functions to the data of Figure 2 and normalized to the d spacing of the unloaded state as measured in air. d spacing was obtained by (a) out-of-plane diffraction and in-plane diffraction with the sample tilted in the direction perpendicular to the X-ray beam by (b) χ = 35°, (c) χ = 55°, and (d) χ = 70°.

Different from bulk measurements, where phase coexistence of fcc- and fct-HfHx is observed in the region between x ≈ 1.78 and x ≈ 1.86,33 the XRD patterns of thin film Hf do not show any indication of the simultaneous occurrence of these two phases. The absence of phase coexistence in thin films could imply that the confinement of the film suppresses the phase coexistence associated with a first order phase transition. It has the important implication that the hysteresis associated with first-order phase transition is (practically) eliminated. Indeed, both earlier25,26 and present (Figure 4(d,e)) optical transmission measurements of Hf indicate a completely hysteresis-free response to changing hydrogen pressures: The optical transmission at a certain PH2 is exactly the same after increasing and decreasing the pressure, making the response of hafnium as an optical sensing material completely free of hysteresis.

In addition, we observe that the phase transition in thin films occurs at a much larger x than in the bulk. Both our data (Figure 1) and data in the literature (see, e.g., refs (27) and (33)) indicate that the onset of the transition occurs at x ≈ 1.78 in the bulk. For thin films, neutron reflectometry measurements (Figure 4(c)) indicate that the partial pressure of the transition, PH2 ≈ 10 Pa, corresponds to x ≈ 1.92. Different from bulk materials, two-dimensional films that are clamped to the substrate have to obey constraints on lateral expansion because expansion can be realized only in the out-of-plane direction, which may result in a very high in-plane stress and affect the stability of phases.16,21,22,34,35 Indeed, the diffraction results indicate that the unit cell expands only in the out-of-plane direction and continuously deforms from its unloaded state: at PH2 = 9.5 Pa (HfH1.90), the out-of-plane d111 spacing is approximately 2% longer than the in-plane spacing (Figure 2(a,d)). Moreover, the nucleation of domains, inducing locally large stresses, may also be hindered considerably by the clamping of the film to the substrate.21,23,34,3639 The collective result of the clamping is that the onset of the transition is pushed to higher values of x, the coexistence of phases is suppressed, and the tetragonality is much smaller than for the bulk.

Most remarkable is that an abrupt increase in both the out-of-plane d111 spacing (Figure 3(a)) and layer thickness (Figure 4(b)) of about 1% is observed at PH2 = 10+2 Pa, while no changes are seen in the in-plane XRD results. Neutron reflectometry measurements (Figure 4(c)) reproduced from ref (26) indicate that this abrupt change in layer thickness and d111 spacing occurs when the layer is nearly fully hydrogenated, i.e., at x ≈ 1.98. While we have no direct explanation for this remarkable observation, a direction for future research would be to determine if such a volume reduction also occurs in other group IV elements Zr and Ti.

In the literature, the root cause of changes in the optical transmission of metal hydride thin films upon exposure to hydrogen has been debated. Often, Beer-Lambert's law is taken as a staring point: T = I/I(d) = e- αd, with T the optical transmission of the film, d the thickness and α the linear attanuation coefficient. In this framework, changes in the optical transmission upon hydrogenation can be due to the volumetric expansion of the film (that needs to be completely translated in an expansion of the film thickness d) or to differences in hydrogen concentration in the metallic thin film causing changes to the dielectric constants and thus the adsorption coeficient α. Different from a previous study of vanadium films for which it was reported that the optical transmission changes are driven by the volumetric expansion of the thin film,40 our results are consistent with optical transmission changes dictated by changes in the hydrogen concentration affecting the dielectric properties. Indeed, the fcc ↔ fct transition including the pronounced reduction in volume is not reflected in the partial pressure dependence of the optical transmission that changes monotonically with increasing partial hydrogen pressure (Figure 4(d,e)). Thus, our results suggests that the changes in optical transmission are fully dictated by the change in hydrogen concentration, causing changes to the dielectric properties, and not by the variation of the thickness of the thin film. This is consistent with neutron reflectometry and optical transmission measurements of Hf25 (and Ta,26 Pd1–yAuy23) that indicate a linear relationship between the optical transmission and the hydrogen concentration of the metallic layer. The fact that the optical transmission is unaffected by the fcc ↔ fct transition has the convenient implication that Hf can be used as a hydrogen sensor in which the optical response monotonically changes with the hydrogen pressure/concentration in the environment over a large range.

In conclusion, using in situ in- and out-of-plane X-ray diffraction and reflectometry, we unambiguously identify that the fcc ↔ fct transition also exists in Hf thin films. The confinement of Hf and the clamping to the substrate significantly affect the transition: compared to bulk Hf, the transition is pushed to higher values of x, the tetragonality of the fct phase is reduced, and the coexistence of phases is suppressed. These results highlight the profound influences of the nanoconfinement of metal hydrides on their structural response to hydrogen. In the case of Hf, the nanoconfinement effects facilitate the extraordinary performance as an optical hydrogen-sensing material including a hysteresis-free sensing range over 6 orders of magnitude in partial hydrogen pressure.

Acknowledgments

We thank Malte Verleg for designing the sample holders for the in-plane in situ XRD measurements, Martin Exeter for fabricating these sample holders from Al, Joost Middelkoop for 3D printing the sample holders, and Reinier den Oudsten for making a spacer for the Anton Paar XRK900 reactor chamber such that the tilted in situ measurements could be performed at the correct height. Quan van der Knokke is acknowledged for fruitful discussions. The neutron scattering experiments were approved by the Neutron Scattering Program Advisory Committee of the Institute of Materials Structure Science (IMSS), High Energy Accelerator Research Organization (KEK) (proposal no. 2014S06).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.1c03411.

  • Experimental details; spectrum of the LED lights used for the optical transmission measurements; photographs of the in situ XRD/XRR setup. Pressure–composition isotherm of bulk powder HfDx; out-of-plane and in-plane XRD measurements of the Hf thin films; and rocking curves of the Hf thin films (PDF)

The authors declare no competing financial interest.

Supplementary Material

References

  1. Rusman N. A. A.; Dahari M. A Review on the Current Progress of Metal Hydrides Material for Solid-State Hydrogen Storage Applications. Int. J. Hydrogen Energy 2016, 41, 12108–12126. 10.1016/j.ijhydene.2016.05.244. [DOI] [Google Scholar]
  2. Schneemann A.; White J. L.; Kang S.; Jeong S.; Wan L. F.; Cho E. S.; Heo T. W.; Prendergast D.; Urban J. J.; Wood B. C.; et al. Nanostructured Metal Hydrides for Hydrogen Storage. Chem. Rev. 2018, 118, 10775–10839. 10.1021/acs.chemrev.8b00313. [DOI] [PubMed] [Google Scholar]
  3. Bannenberg L. J.; Heere M.; Benzidi H.; Montero J.; Dematteis E.; Suwarno S.; Jaron T.; Winny M.; OrÅĆowski P.; Wegner W.; et al. Metal (boro-) Hydrides for High Energy Density Storage and Relevant Emerging Technologies. Int. J. Hydrogen Energy 2020, 45, 33687–33730. 10.1016/j.ijhydene.2020.08.119. [DOI] [Google Scholar]
  4. Nishimura C.; Komaki M.; Hwang S.; Amano M. V. –Ni Alloy Membranes for Hydrogen Purification. J. Alloys Compd. 2002, 330, 902–906. 10.1016/S0925-8388(01)01648-6. [DOI] [Google Scholar]
  5. Dolan M. D.; Viano D. M.; Langley M. J.; Lamb K. E. Tubular Vanadium Membranes for Hydrogen Purification. J. Membr. Sci. 2018, 549, 306–311. 10.1016/j.memsci.2017.12.031. [DOI] [Google Scholar]
  6. Huiberts J. N.; Griessen R.; Rector J. H.; Wijngaarden R. J.; Dekker J. P.; De Groot D. G.; Koeman N. J. Yttrium and Lanthanum Hydride Films with Switchable Optical Properties. Nature 1996, 380, 231–234. 10.1038/380231a0. [DOI] [Google Scholar]
  7. Maiorov V. A. Metal Hydride Switchable Mirrors. Opt. Spectrosc. 2020, 128, 148–165. 10.1134/S0030400X20010154. [DOI] [Google Scholar]
  8. Lototskyy M. V.; Tolj I.; Pickering L.; Sita C.; Barbir F.; Yartys V. The Use of Metal Hydrides in Fuel Cell Applications. Prog. Nat. Sci. 2017, 27, 3–20. 10.1016/j.pnsc.2017.01.008. [DOI] [Google Scholar]
  9. Hübert T.; Boon-Brett L.; Black G.; Banach U. Hydrogen Sensors–a Review. Sens. Actuators, B 2011, 157, 329–352. 10.1016/j.snb.2011.04.070. [DOI] [Google Scholar]
  10. Wadell C.; Syrenova S.; Langhammer C. Plasmonic hydrogen sensing with nanostructured metal hydrides. ACS Nano 2014, 8, 11925–11940. 10.1021/nn505804f. [DOI] [PubMed] [Google Scholar]
  11. Bannenberg L. J.; Boelsma C.; Asano K.; Schreuders H.; Dam B. Metal Hydride Based Optical Hydrogen Sensors. J. Phys. Soc. Jpn. 2020, 89, 051003. 10.7566/JPSJ.89.051003. [DOI] [Google Scholar]
  12. Darmadi I.; Nugroho F. A. A.; Langhammer C. High-Performance Nanostructured Palladium-Based Hydrogen Sensors–Current Limitations and Strategies for Their Mitigation. ACS Sensors 2020, 5, 3306–3327. 10.1021/acssensors.0c02019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Koo W.-T.; Cho H.-J.; Kim D.-H.; Kim Y. H.; Shin H.; Penner R. M.; Kim I.-D. Chemiresistive Hydrogen Sensors: Fundamentals, Recent Advances, and Challenges. ACS Nano 2020, 14, 14284–14322. 10.1021/acsnano.0c05307. [DOI] [PubMed] [Google Scholar]
  14. Chen K.; Yuan D.; Zhao Y. Review of Optical Hydrogen Sensors Based on Metal Hydrides: Recent Developments and Challenges. Opt. Laser Technol. 2021, 137, 106808. 10.1016/j.optlastec.2020.106808. [DOI] [Google Scholar]
  15. Manchester F. D.Phase Diagrams of Binary Hydrogen Alloys; ASM International, 2000. [Google Scholar]
  16. Feenstra R.; de Bruin-Hordijk G. J.; Bakker H. L. M.; Griessen R.; de Groot D. G. Critical Point Lowering in Thin PdHx Films. J. Phys. F: Met. Phys. 1983, 13, L13. 10.1088/0305-4608/13/2/001. [DOI] [Google Scholar]
  17. Pivak Y.; Gremaud R.; Gross K.; Gonzalez-Silveira M.; Walton A.; Book D.; Schreuders H.; Dam B.; Griessen R. Effect of the Substrate on the Thermodynamic Properties of PdHx Films Studied by Hydrogenography. Scr. Mater. 2009, 60, 348–351. 10.1016/j.scriptamat.2008.11.012. [DOI] [Google Scholar]
  18. Bloch J.; Pejova B.; Jacob J.; Hjörvarsson B. Hydrogen-Vanadium System in Thin Films: Effect of Film Thickness. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 82, 245428. 10.1103/PhysRevB.82.245428. [DOI] [Google Scholar]
  19. Baldi A.; Narayan T. C.; Koh A. L.; Dionne J. A. In situ Detection of Hydrogen-Induced Phase Transitions in Individual Palladium Nanocrystals. Nat. Mater. 2014, 13, 1143–1148. 10.1038/nmat4086. [DOI] [PubMed] [Google Scholar]
  20. Syrenova S.; Wadell C.; Nugroho F. A. A.; Gschneidtner T. A.; Fernandez Y. A. D.; Nalin G.; Świtlik D.; Westerlund F.; Antosiewicz T. J.; Zhdanov V. P.; et al. Hydride Formation Thermodynamics and Hysteresis in Individual Pd Nanocrystals with Different Size and Shape. Nat. Mater. 2015, 14, 1236–1244. 10.1038/nmat4409. [DOI] [PubMed] [Google Scholar]
  21. Griessen R.; Strohfeldt N.; Giessen H. Thermodynamics of the Hybrid Interaction of Hydrogen with Palladium Nanoparticles. Nat. Mater. 2016, 15, 311–317. 10.1038/nmat4480. [DOI] [PubMed] [Google Scholar]
  22. Burlaka V.; Wagner S.; Hamm M.; Pundt A. Suppression of Phase Transformation in Nb–H Thin Films below Switchover Thickness. Nano Lett. 2016, 16, 6207–6212. 10.1021/acs.nanolett.6b02467. [DOI] [PubMed] [Google Scholar]
  23. Wagner S.; Pundt A. Quasi-Thermodynamic Model on Hydride Formation in Palladium–Hydrogen Thin Films: Impact of Elastic and Microstructural Constraints. Int. J. Hydrogen Energy 2016, 41, 2727–2738. 10.1016/j.ijhydene.2015.11.063. [DOI] [Google Scholar]
  24. Bannenberg L. J.; Nugroho F. A. A.; Schreuders H.; Norder B.; Trinh T. T.; Steinke N.-J.; Van Well A. A.; Langhammer C.; Dam B. Direct Comparison of PdAu Alloy Thin Films and Nanoparticles upon Hydrogen Exposure. ACS Appl. Mater. Interfaces 2019, 11, 15489–15497. 10.1021/acsami.8b22455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Bannenberg L. J.; Boshuizen B.; Ardy Nugroho F. A.; Schreuders H. Hydrogenation kinetics of metal hydride catalytic layers. ACS Appl. Mater. Interfaces 2021, 10.1021/acsami.1c13240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Boelsma C.; Bannenberg L. J.; van Setten M. J.; Steinke N.-J.; Van Well A. A.; Dam B. Hafnium - an Optical Hydrogen Sensor Spanning Six Orders in Pressure. Nat. Commun. 2017, 8, 15718. 10.1038/ncomms15718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Bannenberg L. J.; Boelsma C.; Schreuders H.; Francke S.; Steinke N.-J.; Van Well A. A.; Dam B. Optical Hydrogen Sensing Beyond Palladium: Hafnium and Tantalum as Effective Sensing Materials. Sens. Actuators, B 2019, 283, 538–548. 10.1016/j.snb.2018.12.029. [DOI] [Google Scholar]
  28. Sidhu S. S.; Heaton L.; Zauberis D. D. Neutron Diffraction Studies of Hafnium–Hydrogen and Titanium–Hydrogen Systems. Acta Crystallogr. 1956, 9, 607–614. 10.1107/S0365110X56001649. [DOI] [Google Scholar]
  29. San-Martin A.; Manchester F. D. The H-Ti (hydrogen-titanium) system. Bull. Alloy Phase Diagrams 1987, 8, 30–42. 10.1007/BF02868888. [DOI] [Google Scholar]
  30. Lewkowicz I. Titanium-Hydrogen. Solid State Phenomena 1996, 49-50, 239–280. 10.4028/www.scientific.net/ssp.49-50.239. [DOI] [Google Scholar]
  31. Zuzek E.; Abriata J. P.; San-Martin A.; Manchester F. D. The H-Zr (Hydrogen-Zirconium) System. Bull. Alloy Phase Diagrams 1990, 11, 385–395. 10.1007/BF02843318. [DOI] [Google Scholar]
  32. Aladjem A. Zirconium-Hydrogen. Solid State Phenomena 1996, 49-50, 281–330. 10.4028/www.scientific.net/SSP.49-50.281. [DOI] [Google Scholar]
  33. Sidhu S. S.; McGuire J. C. An X-ray Diffraction Study of the Hafnium-Hydrogen System. J. Appl. Phys. 1952, 23, 1257–1261. 10.1063/1.1702043. [DOI] [Google Scholar]
  34. Mintz M. H. Hafnium-Hydrogen. Solid State Phenomena 1996, 49-50, 331–356. 10.4028/www.scientific.net/SSP.49-50.331. [DOI] [Google Scholar]
  35. Pundt A.; Kirchheim R. Hydrogen in Metals: Microstructural Aspects. Annu. Rev. Mater. Res. 2006, 36, 555–608. 10.1146/annurev.matsci.36.090804.094451. [DOI] [Google Scholar]
  36. Wagner S.; Pundt A. Mechanical Stress Impact on Thin Pd1–xFex Film Thermodynamic Properties. Appl. Phys. Lett. 2008, 92, 051914. 10.1063/1.2841636. [DOI] [Google Scholar]
  37. Mooij L.; Dam B. Hysteresis and the Role of Nucleation and Growth in the Hydrogenation of Mg Nanolayers. Phys. Chem. Chem. Phys. 2013, 15, 2782–2792. 10.1039/c3cp44441d. [DOI] [PubMed] [Google Scholar]
  38. Bannenberg L. J.; Schreuders H.; van Eijck L.; Heringa J. R.; Steinke N.-J.; Dalgliesh R.; Dam B.; Mulder F. M.; van Well A. A. Impact of Nanostructuring on the Phase Behavior of Insertion Materials: The Hydrogenation Kinetics of a Magnesium Thin Film. J. Phys. Chem. C 2016, 120, 10185–10191. 10.1021/acs.jpcc.6b02302. [DOI] [Google Scholar]
  39. Baldi A.; Mooij L.; Palmisano V.; Schreuders H.; Krishnan G.; Kooi B. J.; Dam B.; Griessen R. Elastic versus Alloying Effects in Mg-Based Hydride Films. Phys. Rev. Lett. 2018, 121, 255503. 10.1103/PhysRevLett.121.255503. [DOI] [PubMed] [Google Scholar]
  40. Wagner S.; Klose P.; Burlaka V.; Nörthemann K.; Hamm M.; Pundt A. Structural Phase Transitions in Niobium Hydrogen Thin Films: Mechanical Stress, Phase Equilibria and Critical Temperatures. ChemPhysChem 2019, 20, 1890–1904. 10.1002/cphc.201900247. [DOI] [PubMed] [Google Scholar]
  41. Droulias S.; Granas O.; Hartmann O.; Komander K.; Hjorvarsson B.; Wolff M.; Palsson G.. The Root Cause of Hydrogen Induced Changes in Optical Transmission of Vanadium. arXiv:1812.04917, 2018. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from The Journal of Physical Chemistry Letters are provided here courtesy of American Chemical Society

RESOURCES