Iron–hydrogen alloy

This article is about the iron–hydrogen alloy. For the chemical compounds, see Iron(I) hydride and Iron(II) hydride.
An 'iron-hydrogen resistor', or barretter -- a glass bulb containing an iron filament in a hydrogen atmosphere that forms a temperature-dependent iron-hydrogen alloy during operation.

An iron–hydrogen alloy, also known as an iron hydride,[1] is an alloy of iron and hydrogen formed when hydrogen atoms dissolve in the crystal structure of metallic iron. This article deals with two broad classes of the iron-hydrogen system: solid iron at ordinary pressure that can incorporate a small amount of hydrogen into its crystal structure; and iron-hydrogen phases that exist at extreme temperatures and pressures, such as might be found in the Earth's core. These substances are the subject of study in industrial metallurgy and planetary geology.

Low pressure phases

Iron is commonly found in the Earth's crust in the form of an ore, usually an iron oxide, such as magnetite, hematite, etc. Iron is smelted from iron ore by a number of chemical processes. One such process, known as hydrogen roasting, is more commonly applied to metals such as tungsten and molybdenum, but can be used to produce iron-hydrogen alloys.

In the narrow range of concentrations that make up iron hydride at atmospheric pressure, mixtures of hydrogen and iron can form a small number of different structures. At room temperature, the most stable form of iron is the body-centred cubic (BCC) structure, α-ferrite. It is a fairly soft metal that can dissolve only a very small concentration of hydrogen, no more than 0.0004 wt% at 912 °C (1,674 °F), and only 0.0002% at 25 °C (77 °F). If iron hydride is at iron hydride-making temperatures it transforms into a face-centred cubic (FCC) structure, called austenite or γ-iron. It is also soft and metallic and can dissolve somewhat more hydrogen, as much as 0.0009% hydrogen at 1,394 °C (2,541 °F). If iron hydride is at temperatures higher than 1,394 °C (2,541 °F) it transforms into a different BCC structure, called δ-iron. It can dissolve even more hydrogen, as much as 0.001% hydrogen at 1,538 °C (2,800 °F), which reflects the upper hydrogen content of iron hydride.[2]

Hydrogen and other elements within iron act as softening agents that promote the movement of dislocations that naturally exist in the iron atom crystal lattices.[3] Varying the amount of alloying elements -- their formation in the iron hydride either as solute elements, or as precipitated phases -- enables the movement of those dislocations that make iron so ductile and weak, and thus controls qualities such as the hardness, ductility, and tensile strength of the resulting iron hydride. Iron hydride's ductility compared to pure iron is only possible at the expense of strength, of which elemental iron has a relative deficiency. However, iron hydride is metastable and subject to the rapid formation of a precipitated phase of gaseous dihydrogen, a process known as hydrogen embrittlement.[4] Thus, the material properties of iron hydride are time-sensitive. One alloying element intentionally added to modify the characteristics of iron hydride is titanium.[5]

High pressure phases

The common form of iron is the “α” form, with body centred cubic (bcc) crystalline structure;[1] in the absence of reactive chemicals, at ambient temperature and 13 GPa of pressure it converts to the “ε” form, with hexagonal close packing (hcp) structure.[6] In an atmosphere of hydrogen at ambient temperature, α-Fe retains its structure up to 3.5 GPa (35,000 atmospheres), with only small amounts of hydrogen diffusing into it forming a solid interstitial solution.[7]

Starting at about 3.5 GPa of pressure, hydrogen H
2
rapidly diffuses into metallic iron (with diffusion length of about 500 mm per 10 s at 5 GPa[8]) to form a crystalline solid with formula close to FeH. This reaction, in which the iron expands significantly, was first inferred from the unexpected deformation of steel gaskets in diamond anvil cell experiments. In 1991 J. V. Badding and others identified the compound by X-ray diffraction as having the approximate composition FeH0.94 and double hexagonal close packed (dhcp) structure.[1]

Since then the phase diagram of this high-pressure iron-hydrogen compound has been intensively investigated up to 70 GPa. Three stable crystalline forms have been observed, denoted “ε’” (the original dhcp form),[1] “ε” (hexagonal close packed, hpc), and “γ” (face centered cubic, fcc).[7][9][10] A fourth metastable “α” form (body-centered cubic, bcc) has also been identified.[7] In all these phases the packing of iron atoms is less dense than in pure iron. The hcp and fcc forms have the same iron lattice but have different number of hydrogen neighbors, and have different local magnetic moments.[11] The hydrogen and iron atoms are electrically neutral in this compound.[8]

At low temperatures the stable forms are bcc below 5 GPa and ε’ (dhcp) above 5 GPa at least up to 80 GPa; at higher temperatures γ (fcc) exists at least up to 20 GPa.[12] The triple point ε'-γ-melt is predicted to be at 60 GPa and 2000 K.[12] Theoretical calculations however predict that, at 300 K, the stable structures should be dhcp below 37 GPa, hcp between 37–83 GPa, and fcc above 83 GPa.[12]

Other hydrogenated forms FeHx with x = 0.25 (Fe
4
H
), x = 0.50 (Fe
2
H
), and x = 0.75 (Fe
4
H
3
) have been the subject of theoretical studies.[11]

These compounds dissociate spontaneously at ordinary pressures, but at very low temperatures they will survive long enough in a metastable state to be studied.[7] At ordinary temperatures, rapid depressurization of FeH from 7.5 GPa (at 1.5 GPa/s) results in metallic iron containing many small hydrogen bubbles; with slow depressurization the hydrogen diffuses out of the metal.[8]

Hydrogen is a solid at room temperature above 5 GPa.[12]

ε’ (dhcp) form

The double hexagonal close packed (dhcp) structure with ABAC alignment of FeH. Each sphere is an iron atom. Hydrogen are located in the interstices.

The best-known form of FeH (characterized by V. E. Antonov and others, 1989) has a double hexagonal close packed (dhcp) structure. It consists of layers of hexagonal packed iron atoms, offset in a pattern ABAC; which means that even-numbered layers are vertically aligned, while the odd-numbered ones alternate between the two possible relative alignments. The c axis of the unit cell is 0.87 nm. Hydrogen atoms occupy octahedral cavities between the layers. The hydrogen layers come in vertically aligned pairs, bracketing the B and C layers and shifted like them.[7] For each hydrogen added the unit cell expands by 1.8 Å3 (0.0018 nm3). This phase was denoted ε’, after the similar structure that iron assumes above 14 GPa.[6]

This form of FeH is rapidly created at room temperature and 3.8 GPa from hydrogen and α-iron.[12] The transformation entails an expansion by 17–20% in volume.[11][13] The reaction is complex and may involve a metastable hcp intermediate form; at 9 GPa and 350 °C there are still noticeable amounts of unreacted α-Fe in the solid.[7] The same form is obtained from by reacting hydrogen with the higher-pressure hcp form of iron (ε-Fe) at 1073 K and 20 GPa for 20 min;[13] and also from α-iron and H
2
O
at 84 GPa and 1300 K.[12]

This form of FeH is stable at room temperature at least up to 80 GPa,[12] but turns into the γ form between 1073 and 1173 K and 20 GPa.[13] There may be several dhcp forms with slightly different stoichiometries,[9] but the material is stoichiometric above 10 GPa in the presence of excess hydrogen.[13]

This material has metallic appearance and is an electrical conductor.[6] Its resistivity is higher than that of iron, and decreases down to a minimum at 8 GPa. Above 13 GPa the resistivity increases with pressure. The material is ferromagnetic at the lowest pressure range, but the ferromagnetism begins to decrease at 20 GPa and disappears at 32 GPa t.[6][9]

The bulk elasticity modulus of this compound is 121 ± 19 GPa, substantially lower than iron’s 160 GPa. This difference means that at 3.5 GPa FeH has 51% less volume than the mixture of hydrogen and iron that forms it.[1]

The speed of compressional sound waves in FeH rises as pressure rises, at 10 GPa it is at 6.3 km/s, at 40 GPa 8.3 km/s and 70 GPa 9 km/s.[12]

The dhcp form of FeH can be preserved in a metastable form at ambient pressures by first lowering the temperature below 100 K.[7]

ε (hcp) form

A hexagonal close packed (hcp) form of FeH also exists at lower pressure hydrogen, also described by M. Yamakata and others in 1992. This is called the ε phase (no prime).[7] The hcp phase is not ferromagnetic,[9] probably paramagnetic.[7] This appears to be the most stable form in a wide pressure range.[11] It seems to have a composition between FeH
0.42
.[7]

The hcp form of FeH can be preserved in a metastable form at ambient pressures by first lowering the temperature below 100 K.[7]

High pressure stability of different iron hydrides was systematically studied using density-functional calculations and evolutionary crystal structure prediction by Bazhanova et al.,[10] who found that at pressures of the Earth's inner core FeH, FeH3 and an unexpected compound FeH4 are thermodynamically stable, whereas FeH2 is not.

Melting point

These high pressure iron-hydrogen alloys melt at a significantly lower temperature than pure iron:[8][13]

Pressure (Gpa) 7.5 10 11.5 15 18 20
Melting point (C) 1150 1473 1448 1538 1548 1585

The slope of the melting point curve with pressure (dT/dP) is 13 K/GPa.[13]

Occurrence in the Earth’s core

Very little is known about the composition of Earth’s inner core. The only parameters that are known with confidence are the speed of the pressure and shear sound waves (the existence of the latter implying that it is a solid). The pressure at the boundary between the inner core and the liquid outer core is estimated at 330 GPa,[12] still somewhat beyond the range of laboratory experiments. The density of the outer and inner cores can only be estimated by indirect means. The inner core was at first thought to be 10% less dense than pure iron at the predicted conditions,[1][8] but this presumed “density deficit” has later been revised downwards: 2 to 5% by some estimates[12] or 1 to 2% by others.[9]

The density deficit is thought to be due to mixture of lighter elements such as silicon or carbon.[1] Hydrogen has been thought unlikely because of its volatility, but recent studies have uncovered plausible mechanisms for its incorporation and permanence in the core. It is estimated that hcp FeH would be stable under those conditions.[12] Iron–hydrogen alloys could have been formed in a reaction of iron with water in magma during the formation of the earth. Above 5 GPa, iron will split water yielding the hydride and ferrous ions:[9]

3Fe + H
2
O
→ 2FeH + FeO

Indeed, Okuchi obtained magnetite and iron hydride by reacting magnesium silicate, magnesium oxide, silica and water with metallic iron in a diamond cell at 2000 C.[8][14] Okuchi argues that most of the hydrogen accreted to Earth should have dissolved into the primeval magma ocean; and if the pressure at the bottom of the magma was 7.5 GPa or more, then almost all of that hydrogen would have reacted with iron to form the hydride, which then would have sunk to the core where it would be stabilized by the increased pressure.[8] Moreover, it appears that at those pressures iron binds hydrogen in preference to carbon.[9]

Based on density and sound velocity measurements at room temperature and up to 70 GPa, extrapolated to core conditions, Shibazaki and others claim that the presence of 0.23 ± 0.06% hydrogen in weight (that is, a mean atomic composition of FeH0.13 ± 0.03) would explain a 2–5% density deficit.[12] and match the observed speed of pressure and shear sound waves in the solid inner core.[12] A different study predicts 0.08–0.16% (weight) hydrogen in the inner core,[9] while others proposed from 50% to 95% FeH (by mole count) If the core has this much hydrogen it would amount to ten times as much as in the oceans.[14]

The liquid outer core also appears to have density 5–10% lower than iron.[9][13] Shibazaki and others estimate that it should have a somewhat higher proportion of hydrogen than the inner core, but there is not enough data about molten FeHx for accurate estimates.[12] Narygina and others estimate 0.5–1.0% (weight) of hydrogen in the melt.[9] Similar, but without extrapolations in pressure, theoretical estimates give a narrower range of concentrations 0.4-0.5% (weight),[10] however, this results to too low mean atomic mass of the inner core (43.8-46.5) and hydrogen seems to be less likely than other elements (S, Si, C, O) to be the main light alloying element in the core.

See also

References

  1. 1 2 3 4 5 6 7 J.V. Badding, R.J. Hemley, and H.K. Mao (1991), "High-pressure chemistry of hydrogen in metals: in situ study of iron hydride." Science, American Association for the Advancement of Science, volume 253, issue 5018, pages 421-424 doi:10.1126/science.253.5018.421
  2. San-Martin, A.; Manchester, F. D. (1 April 1990). "The Fe-H (Iron-Hydrogen) System". Bulletin of Alloy Phase Diagrams. Springer-Verlag. 11 (2): 173–184. doi:10.1007/BF02841704. ISSN 0197-0216.
  3. Sakaki, K.; Kawase, T.; Hirato, M.; Mizuno, M.; Araki, H.; Shirai, Y.; Nagumo, M. (December 2006). "The Effect of Hydrogen on Vacancy Generation in Iron by Plastic Deformation". Scripta Materialia. Elsevier Ltd. 55 (11): 1031–1034. doi:10.1016/j.scriptamat.2006.08.030.
  4. Itakura, M.; Kaburaki, H.; Yamaguchi, M.; Okita, T. (October 2013). "The Effect of Hydrogen Atoms on the Screw Dislocation Mobility in BCC Iron: A First-Principles Study". Acta Materialia. Elsevier Ltd. 61 (18): 6857–6867. doi:10.1016/j.actamat.2013.07.064.
  5. Sarkar, Arindam; Banerjee, Rangan (July 2005). "Net energy analysis of hydrogen storage options". International Journal of Hydrogen Energy. Elsevier Ltd. 30 (8): 867–877. doi:10.1016/j.ijhydene.2004.10.021.
  6. 1 2 3 4 Takahiro Matsuoka, Naohisa Hirao,Yasuo Ohishi, Katsuya Shimizu, Akihiko Machida and Katsutoshi Aoki (), "Structural and electrical transport properties of FeHx under high pressures and low temperatures". High Pressure Research, volume 31, issue 1, pages 64–67 doi:10.1080/08957959.2010.522447
  7. 1 2 3 4 5 6 7 8 9 10 11 V. E. Antonov, K. Cornell, V.K. Fedotov, A. I. Kolesnikov E.G. Ponyatovsky, V.I. Shiryaev, H. Wipf (1998) "Neutron diffraction investigation of the dhcp and hcp iron hydrides and deuterides". Journal of Alloys and Compounds, volume 264, pages 214–222 doi:10.1016/S0925-8388(97)00298-3
  8. 1 2 3 4 5 6 7 Takuo Okuchi (1997), "Hydrogen partitioning into molten iron at high pressure: implications for Earth's core." Science (American Association for the Advancement of Science), volume 278, pages 1781-1784. doi:10.1126/science.278.5344.1781
  9. 1 2 3 4 5 6 7 8 9 10 Olga Narygina, Leonid S. Dubrovinsky, Catherine A. McCammon, Alexander Kurnosov, Innokenty Yu. Kantor, Vitali B. Prakapenka, and Natalia A. Dubrovinskaia (2011), "FeH at high pressures and implications for the composition of the Earth's core". Earth and Planetary Science Letters, volume 307, issue 3–4, pages 409–414 doi:10.1016/j.epsl.2011.05.015
  10. 1 2 3 Zulfiya G. Bazhanova, Artem R. Oganov, Omar Gianola (2012) "Fe-C-H system at pressures of the Earth's inner core". Physics-Uspekhi, volume 55, pages 489-497
  11. 1 2 3 4 A. S. Mikhaylushkin, N. V. Skorodumova, R. Ahuja, B. Johansson (2006), "Structural and magnetic properties of FeHx (x=0.25; 0.50;0.75)". In: Hydrogen in Matter: A Collection from the Papers Presented at the Second International Symposium on Hydrogen in Matter (ISOHIM), AIP Conference Proceedings, volume 837, pages 161–167 doi:10.1063/1.2213072
  12. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Shibazaki, Yuki; Eiji Ohtani; Hiroshi Fukui; Takeshi Sakai; Seiji Kamada; Daisuke Ishikawa; Satoshi Tsutsui; Alfred Q.R. Baron; Naoya Nishitani; Naohisa Hirao; Kenichi Takemura (1 January 2012). "Sound velocity measurements in dhcp-FeH up to 70 GPa with inelastic X-ray scattering: Implications for the composition of the Earth's core". Earth and Planetary Science Letters. 313–314: 79–85. doi:10.1016/j.epsl.2011.11.002.
  13. 1 2 3 4 5 6 7 Sakamaki, K; Takahashi, E.; Nakajima, Y.; Nishihara, Y.; Funakoshi, K.; Suzuki, T.; Fukai, Y. (May 2009). "Melting phase relation of FeHx up to 20GPa: Implication for the temperature of the Earth's core". Physics of the Earth and Planetary Interiors. 174: 192–201. doi:10.1016/j.pepi.2008.05.017.
  14. 1 2 Surendra K. Saxena, Hanns-Peter Liermann, and Guoyin Shen (2004), "Formation of iron hydride and high-magnetite at high pressure and temperature". Physics of the Earth and Planetary Interiors, volume 146, pages 313-317. doi:10.1016/j.pepi.2003.07.030
This article is issued from Wikipedia - version of the 7/10/2015. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.