Samarium compounds

Last updated

Samarium compounds are compounds formed by the lanthanide metal samarium (Sm). In these compounds, samarium generally exhibits the +3 oxidation state, such as SmCl3, Sm(NO3)3 and Sm(C2O4)3. Compounds with samarium in the +2 oxidation state are also known, for example SmI2.

Contents

Properties of samarium compounds

Formulacolorsymmetry space group No Pearson symbol a (pm)b (pm)c (pm)Zdensity,
g/cm3
Smsilverytrigonal [1] R3m166hR9362.9362.92621.397.52
Smsilveryhexagonal [1] P63/mmc194hP4362362116847.54
Smsilverytetragonal [2] I4/mmm139tI2240.2240.2423.1220.46
SmOgoldencubic [3] Fm3m225cF8494.3494.3494.349.15
Sm2O3trigonal [4] P3m1164hP5377.8377.859417.89
Sm2O3monoclinic [4] C2/m12mS301418362.4885.567.76
Sm2O3cubic [5] Ia3206cI80109310931093167.1
SmH2cubic [6] Fm3m225cF12537.73537.73537.7346.51
SmH3hexagonal [7] P3c1165hP24377.1377.1667.26
Sm2B5graymonoclinic [8] P21/c14mP28717.9718720.546.49
SmB2hexagonal [9] P6/mmm191hP3331331401.917.49
SmB4tetragonal [10] P4/mbm127tP20717.9717.9406.746.14
SmB6cubic [11] Pm3m221cP7413.4413.4413.415.06
SmB66cubic [12] Fm3c226cF19362348.72348.72348.7242.66
Sm2C3cubic [13] I43d220cI40839.89839.89839.8987.55
SmC2tetragonal [13] I4/mmm139tI6377377633.126.44
SmF2purple [14] cubic [15] Fm3m225cF12587.1587.1587.146.18
SmF3white [14] orthorhombic [15] Pnma62oP16667.22705.85440.4346.64
SmCl2brown [14] orthorhombic [16] Pnma62oP12756.28450.77901.0944.79
SmCl3yellow [14] hexagonal [15] P63/m176hP8737.33737.33416.8424.35
SmBr2brown [14] orthorhombic [17] Pnma62oP12797.7475.4950.645.72
SmBr3yellow [14] orthorhombic [18] Cmcm63oS16404126590825.58
SmI2green [14] monoclinicP21/c14mP12
SmI3orange [14] trigonal [19] R363hR24749749208065.24
SmNcubic [20] Fm3m225cF835735735748.48
SmPcubic [21] Fm3m225cF857657657646.3
SmAscubic [22] Fm3m225cF8591.5591.5591.547.23

Chalcogenides

Oxides

Samarium sesquioxide Samarium(III) oxide.jpg
Samarium sesquioxide

The most stable oxide of samarium is the sesquioxide Sm2O3. Like many samarium compounds, it exists in several crystalline phases. The trigonal form is obtained by slow cooling from the melt. The melting point of Sm2O3 is high (2345 °C), so it is usually melted not by direct heating, but with induction heating, through a radio-frequency coil. Sm2O3 crystals of monoclinic symmetry can be grown by the flame fusion method (Verneuil process) from Sm2O3 powder, that yields cylindrical boules up to several centimeters long and about one centimeter in diameter. The boules are transparent when pure and defect-free and are orange otherwise. Heating the metastable trigonal Sm2O3 to 1900 °C converts it to the more stable monoclinic phase. [4] Cubic Sm2O3 has also been described. [5]

Samarium is one of the few lanthanides that form a monoxide, SmO. This lustrous golden-yellow compound was obtained by reducing Sm2O3 with samarium metal at high temperature (1000 °C) and pressure above 50 kbar; lowering the pressure resulted in incomplete reaction. SmO has cubic rock-salt lattice structure. [3] [23]

Other chalcogenides

Samarium forms a trivalent sulfide, selenide and telluride. Divalent chalcogenides SmS, SmSe and SmTe with cubic rock-salt crystal structure are also known. They are remarkable by converting from semiconducting to metallic state at room temperature upon application of pressure. Whereas the transition is continuous and occurs at about 20–30 kbar in SmSe and SmTe, it is abrupt in SmS and requires only 6.5 kbar. This effect results in spectacular color change in SmS from black to golden yellow when its crystals of films are scratched or polished. The transition does not change lattice symmetry, but there is a sharp decrease (~15%) in the crystal volume. [24] It shows hysteresis, that is when the pressure is released, SmS returns to the semiconducting state at much lower pressure of about 0.4 kbar. [25] [26]

Halides

Samarium trichloride hexahydrate Samarium trichloride hexahydrate.jpg
Samarium trichloride hexahydrate

Samarium metal reacts with all the halogens, forming trihalides: [27]

2 Sm (s) + 3 X2 (g) → 2 SmX3 (s) (X = F, Cl, Br or I)

Their further reduction with samarium, lithium or sodium metals at elevated temperatures (about 700–900 °C) yields dihalides. [16] The diiodide can also be prepared by heating SmI3, or by reacting the metal with 1,2-diiodoethane in anhydrous tetrahydrofuran at room temperature: [28]

Sm (s) + ICH2-CH2I → SmI2 + CH2=CH2.

In addition to dihalides, the reduction also produces many non-stoichiometric samarium halides with a well-defined crystal structure, such as Sm3F7, Sm14F33, Sm27F64, [15] Sm11Br24, Sm5Br11 and Sm6Br13. [29]

As reflected in the table above, samarium halides change their crystal structures when one type of halide atom is substituted for another, which is an uncommon behavior for most elements (e.g. actinides). Many halides have two major crystal phases for one composition, one being significantly more stable and another being metastable. The latter is formed upon compression or heating, followed by quenching to ambient conditions. For example, compressing the usual monoclinic samarium diiodide and releasing the pressure results in a PbCl2-type orthorhombic structure (density 5.90 g/cm3), [30] and similar treatment results in a new phase of samarium triiodide (density 5.97 g/cm3). [31]

Borides

Sintering powders of samarium oxide and boron, in vacuum, yields a powder containing several samarium boride phases, and their volume ratio can be controlled through the mixing proportion. [32] The powder can be converted into larger crystals of a certain samarium boride using arc melting or zone melting techniques, relying on the different melting/crystallization temperature of SmB6 (2580 °C), SmB4 (about 2300 °C) and SmB66 (2150 °C). All these materials are hard, brittle, dark-gray solids with the hardness increasing with the boron content. [11] Samarium diboride is too volatile to be produced with these methods and requires high pressure (about 65 kbar) and low temperatures between 1140 and 1240 °C to stabilize its growth. Increasing the temperature results in the preferential formations of SmB6. [9]

Samarium hexaboride

Samarium hexaboride is a typical intermediate-valence compound where samarium is present both as Sm2+ and Sm3+ ions at the ratio 3:7. [32] It belongs to a class of Kondo insulators, that is at high temperatures (above 50 K), its properties are typical of a Kondo metal, with metallic electrical conductivity characterized by strong electron scattering, whereas at low temperatures, it behaves as a non-magnetic insulator with a narrow band gap of about 4–14 meV. [33] The cooling-induced metal-insulator transition in SmB6 is accompanied by a sharp increase in the thermal conductivity, peaking at about 15 K. The reason for this increase is that electrons themselves do not contribute to the thermal conductivity at low temperatures, which is dominated by phonons, but the decrease in electron concentration reduced the rate of electron-phonon scattering. [34]

New research seems to show that it may be a topological insulator. [35] [36] [37]

Other inorganic compounds

Samarium sulfate, Sm2(SO4)3 Samarium-sulfate.jpg
Samarium sulfate, Sm2(SO4)3

Samarium carbides are prepared by melting a graphite-metal mixture in an inert atmosphere. After the synthesis, they are unstable in air and are studied also under inert atmosphere. [13] Samarium monophosphide SmP is a semiconductor with the bandgap of 1.10 eV, the same as in silicon, and high electrical conductivity of n-type. It can be prepared by annealing at 1100 °C an evacuated quartz ampoule containing mixed powders of phosphorus and samarium. Phosphorus is highly volatile at high temperatures and may explode, thus the heating rate has to be kept well below 1 °C/min. [21] Similar procedure is adopted for the monarsenide SmAs, but the synthesis temperature is higher at 1800 °C. [22]

Numerous crystalline binary compounds are known for samarium and one of the group-14, 15 or 16 element X, where X is Si, Ge, Sn, Pb, Sb or Te, and metallic alloys of samarium form another large group. They are all prepared by annealing mixed powders of the corresponding elements. Many of the resulting compounds are non-stoichiometric and have nominal compositions SmaXb, where the b/a ratio varies between 0.5 and 3. [38] [39] [40]

Organosamarium compounds

Crystal of the complex Sm(Cp )3 Sm(Cp-tet)3 crystals.png
Crystal of the complex Sm(Cp )3

Samarium forms a cyclopentadienide Sm(C5H5)3 and its chloroderivatives Sm(C5H5)2Cl and Sm(C5H5)Cl2. They are prepared by reacting samarium trichloride with NaC5H5 in tetrahydrofuran. Contrary to cyclopentadienides of most other lanthanides, in Sm(C5H5)3 some C5H5 rings bridge each other by forming ring vertexes η1 or edges η2 toward another neighboring samarium, thus creating polymeric chains. [41] The chloroderivative Sm(C5H5)2Cl has a dimer structure, which is more accurately expressed as (η(5)−C5H5)2Sm(−Cl)2(η(5)−C5H5)2. There, the chlorine bridges can be replaced, for instance, by iodine, hydrogen or nitrogen atoms or by CN groups. [42]

The (C5H5) ion in samarium cyclopentadienides can be replaced by the indenide (C9H7) or cyclooctatetraenide (C8H8)2− ring, resulting in Sm(C9H7)3 or KSm(η(8)−C8H8)2. The latter compound has a structure similar to uranocene. There is also a cyclopentadienide of divalent samarium, Sm(C5H5)2− a solid that sublimates at about 85 °C. Contrary to ferrocene, the C5H5 rings in Sm(C5H5)2 are not parallel but are tilted by 40°. [42] [43]

A metathesis reaction in tetrahydrofuran or ether gives alkyls and aryls of samarium: [42]

SmCl3 + 3LiR → SmR3 + 3LiCl
Sm(OR)3 + 3LiCH(SiMe3)2 → Sm{CH(SiMe3)2}3 + 3LiOR

Here R is a hydrocarbon group and Me = methyl.

Pictures of samarium compounds

See also

Related Research Articles

<span class="mw-page-title-main">Berkelium</span> Chemical element, symbol Bk and atomic number 97

Berkelium is a synthetic chemical element; it has symbol Bk and atomic number 97. It is a member of the actinide and transuranium element series. It is named after the city of Berkeley, California, the location of the Lawrence Berkeley National Laboratory where it was discovered in December 1949. Berkelium was the fifth transuranium element discovered after neptunium, plutonium, curium and americium.

<span class="mw-page-title-main">Holmium</span> Chemical element, symbol Ho and atomic number 67

Holmium is a chemical element; it has symbol Ho and atomic number 67. It is a rare-earth element and the eleventh member of the lanthanide series. It is a relatively soft, silvery, fairly corrosion-resistant and malleable metal. Like many other lanthanides, holmium is too reactive to be found in native form, as pure holmium slowly forms a yellowish oxide coating when exposed to air. When isolated, holmium is relatively stable in dry air at room temperature. However, it reacts with water and corrodes readily, and also burns in air when heated.

The lanthanide or lanthanoid series of chemical elements comprises at least the 14 metallic chemical elements with atomic numbers 57–70, from lanthanum through ytterbium. In the periodic table, they fill the 4f orbitals. Lutetium is also sometimes considered a lanthanide, despite being a d-block element and a transition metal.

<span class="mw-page-title-main">Samarium</span> Chemical element, symbol Sm and atomic number 62

Samarium is a chemical element; it has symbol Sm and atomic number 62. It is a moderately hard silvery metal that slowly oxidizes in air. Being a typical member of the lanthanide series, samarium usually has the oxidation state +3. Compounds of samarium(II) are also known, most notably the monoxide SmO, monochalcogenides SmS, SmSe and SmTe, as well as samarium(II) iodide.

<span class="mw-page-title-main">Ytterbium</span> Chemical element, symbol Yb and atomic number 70

Ytterbium is a chemical element; it has symbol Yb and atomic number 70. It is a metal, the fourteenth and penultimate element in the lanthanide series, which is the basis of the relative stability of its +2 oxidation state. Like the other lanthanides, its most common oxidation state is +3, as in its oxide, halides, and other compounds. In aqueous solution, like compounds of other late lanthanides, soluble ytterbium compounds form complexes with nine water molecules. Because of its closed-shell electron configuration, its density, melting point and boiling point are much lower than those of most other lanthanides.

<span class="mw-page-title-main">Cubic crystal system</span> Crystallographic system where the unit cell is in the shape of a cube

In crystallography, the cubiccrystal system is a crystal system where the unit cell is in the shape of a cube. This is one of the most common and simplest shapes found in crystals and minerals.

<span class="mw-page-title-main">Samarium(III) chloride</span> Chemical compound

Samarium(III) chloride, also known as samarium trichloride, is an inorganic compound of samarium and chloride. It is a pale yellow salt that rapidly absorbs water to form a hexahydrate, SmCl3.6H2O. The compound has few practical applications but is used in laboratories for research on new compounds of samarium.

Samarium(III) sulfide (Sm2S3) is a chemical compound of the rare earth element samarium, and sulfur. In this compound samarium is in the +3 oxidation state, and sulfur is an anion in the −2 state.

<span class="mw-page-title-main">Calcium disilicide</span> Chemical compound

Calcium disilicide (CaSi2) is an inorganic compound, a silicide of calcium. It is a whitish or dark grey to black solid matter with melting point 1033 °C. It is insoluble in water, but may decompose when subjected to moisture, evolving hydrogen and producing calcium hydroxide. It decomposes in hot water, and is flammable and may ignite spontaneously in air.

<span class="mw-page-title-main">Yttrium borides</span> Chemical compound

Yttrium boride refers to a crystalline material composed of different proportions of yttrium and boron, such as YB2, YB4, YB6, YB12, YB25, YB50 and YB66. They are all gray-colored, hard solids having high melting temperatures. The most common form is the yttrium hexaboride YB6. It exhibits superconductivity at relatively high temperature of 8.4 K and, similar to LaB6, is an electron cathode. Another remarkable yttrium boride is YB66. It has a large lattice constant (2.344 nm), high thermal and mechanical stability, and therefore is used as a diffraction grating for low-energy synchrotron radiation (1–2 keV).

Samarium monochalcogenides are chemical compounds with the composition SmX, where Sm stands for the lanthanide element samarium and X denotes any one of three chalcogen elements, sulfur, selenium or tellurium, resulting in the compounds SmS, SmSe or SmTe. In these compounds, samarium formally exhibits oxidation state +2, whereas it usually assumes the +3 state, resulting in chalcogenides with the chemical formula Sm2X3.

<span class="mw-page-title-main">Berkelium compounds</span> Chemical compounds

Berkelium forms a number of chemical compounds, where it normally exists in an oxidation state of +3 or +4, and behaves similarly to its lanthanide analogue, terbium. Like all actinides, berkelium easily dissolves in various aqueous inorganic acids, liberating gaseous hydrogen and converting into the trivalent oxidation state. This trivalent state is the most stable, especially in aqueous solutions, but tetravalent berkelium compounds are also known. The existence of divalent berkelium salts is uncertain and has only been reported in mixed lanthanum chloride-strontium chloride melts. Aqueous solutions of Bk3+ ions are green in most acids. The color of the Bk4+ ions is yellow in hydrochloric acid and orange-yellow in sulfuric acid. Berkelium does not react rapidly with oxygen at room temperature, possibly due to the formation of a protective oxide surface layer; however, it reacts with molten metals, hydrogen, halogens, chalcogens and pnictogens to form various binary compounds. Berkelium can also form several organometallic compounds.

<span class="mw-page-title-main">Samarium hexaboride</span> Chemical compound

Samarium hexaboride (SmB6) is an intermediate-valence compound where samarium is present both as Sm2+ and Sm3+ ions at the ratio 3:7. It is a Kondo insulator having a metallic surface state.

<span class="mw-page-title-main">Werner Urland</span> German chemist (born 1944)

Werner Urland is a German chemist whose name is imprinted in the pioneering implementation of the Angular Overlap Model for the interpretation of optical and magnetic properties of rare-earth coordination compounds. This approach receives a renewed value in the context of the vogue around the lanthanide-based new materials, such as achieving magnets at molecular scale, or designing new phosphor materials.

<span class="mw-page-title-main">Lanthanocene</span>

A lanthanocene is a type of metallocene compound that contains an element from the lanthanide series. The most common lanthanocene complexes contain two cyclopentadienyl anions and an X type ligand, usually hydride or alkyl ligand.

An yttrium compound is a chemical compound containing yttrium. Among these compounds, yttrium generally has a +3 valence. The solubility properties of yttrium compounds are similar to those of the lanthanides. For example oxalates and carbonates are hardly soluble in water, but soluble in excess oxalate or carbonate solutions as complexes are formed. Sulfates and double sulfates are generally soluble. They resemble the "yttrium group" of heavy lanthanide elements.

<span class="mw-page-title-main">Europium compounds</span> Chemical compounds

Europium compounds are compounds formed by the lanthanide metal europium (Eu). In these compounds, europium generally exhibits the +3 oxidation state, such as EuCl3, Eu(NO3)3 and Eu(CH3COO)3. Compounds with europium in the +2 oxidation state are also known. The +2 ion of europium is the most stable divalent ion of lanthanide metals in aqueous solution. Many europium compounds fluoresce under ultraviolet light due to the excitation of electrons to higher energy levels. Lipophilic europium complexes often feature acetylacetonate-like ligands, e.g., Eufod.

Ytterbium compounds are chemical compounds that contain the element ytterbium (Yb). The chemical behavior of ytterbium is similar to that of the rest of the lanthanides. Most ytterbium compounds are found in the +3 oxidation state, and its salts in this oxidation state are nearly colorless. Like europium, samarium, and thulium, the trihalides of ytterbium can be reduced to the dihalides by hydrogen, zinc dust, or by the addition of metallic ytterbium. The +2 oxidation state occurs only in solid compounds and reacts in some ways similarly to the alkaline earth metal compounds; for example, ytterbium(II) oxide (YbO) shows the same structure as calcium oxide (CaO).

Hafnium compounds are compounds containing the element hafnium (Hf). Due to the lanthanide contraction, the ionic radius of hafnium(IV) (0.78 ångström) is almost the same as that of zirconium(IV) (0.79 angstroms). Consequently, compounds of hafnium(IV) and zirconium(IV) have very similar chemical and physical properties. Hafnium and zirconium tend to occur together in nature and the similarity of their ionic radii makes their chemical separation rather difficult. Hafnium tends to form inorganic compounds in the oxidation state of +4. Halogens react with it to form hafnium tetrahalides. At higher temperatures, hafnium reacts with oxygen, nitrogen, carbon, boron, sulfur, and silicon. Some compounds of hafnium in lower oxidation states are known.

Erbium(III) selenide is an inorganic compound with a chemical formula of Er2Se3.

References

  1. 1 2 Shi, N.; Fort, D. (1985). "Preparation of samarium in the double hexagonal close packed form". Journal of the Less Common Metals. 113 (2): 21. doi:10.1016/0022-5088(85)90294-2.
  2. Vohra, Y.; Akella, Jagannadham; Weir, Sam; Smith, Gordon S. (1991). "A new ultra-high pressure phase in samarium". Physics Letters A. 158 (1–2): 89. Bibcode:1991PhLA..158...89V. doi:10.1016/0375-9601(91)90346-A.
  3. 1 2 Leger, J.; Yacoubi, N.; Loriers, J. (1981). "Synthesis of rare earth monoxides". Journal of Solid State Chemistry. 36 (3): 261. Bibcode:1981JSSCh..36..261L. doi: 10.1016/0022-4596(81)90436-9 .
  4. 1 2 3 Gouteron, J.; Michel, D.; Lejus, A. M.; Zarembowitch, J. (1981). "Raman spectra of lanthanide sesquioxide single crystals: Correlation between A and B-type structures". Journal of Solid State Chemistry. 38 (3): 288. Bibcode:1981JSSCh..38..288G. doi:10.1016/0022-4596(81)90058-X.
  5. 1 2 Taylor D. (1984). Br. Ceram. Trans. J. 83: 92–98.{{cite journal}}: CS1 maint: untitled periodical (link)
  6. Daou, J.; Vajda, P.; Burger, J. (1989). "Low temperature thermal expansion in SmH2+x". Solid State Communications. 71 (12): 1145. Bibcode:1989SSCom..71.1145D. doi:10.1016/0038-1098(89)90728-X.
  7. Dolukhanyan, S. (1997). "Synthesis of novel compounds by hydrogen combustion". Journal of Alloys and Compounds. 253–254: 10. doi:10.1016/S0925-8388(96)03071-X.
  8. Zavalii, L. V.; Kuz'ma, Yu. B.; Mikhalenko, S. I. (1990). "Sm2B5 boride and its structure". Soviet Powder Metallurgy and Metal Ceramics. 29 (6): 471. doi:10.1007/BF00795346. S2CID   138416728.
  9. 1 2 Cannon, J.; Cannon, D.; Tracyhall, H. (1977). "High pressure syntheses of SmB2 and GdB12". Journal of the Less Common Metals. 56: 83. doi:10.1016/0022-5088(77)90221-1.
  10. Etourneau, J.; Mercurio, J.; Berrada, A.; Hagenmuller, P.; Georges, R.; Bourezg, R.; Gianduzzo, J. (1979). "The magnetic and electrical properties of some rare earth tetraborides". Journal of the Less Common Metals. 67 (2): 531. doi:10.1016/0022-5088(79)90038-9.
  11. 1 2 Solovyev, G. I.; Spear, K. E. (1972). "Phase Behavior in the Sm-B System". Journal of the American Ceramic Society. 55 (9): 475. doi:10.1111/j.1151-2916.1972.tb11344.x.
  12. Schwetz, K.; Ettmayer, P.; Kieffer, R.; Lipp, A. (1972). "Über die Hektoboridphasen der Lanthaniden und Aktiniden". Journal of the Less Common Metals. 26: 99. doi:10.1016/0022-5088(72)90012-4.
  13. 1 2 3 Spedding, F. H.; Gschneidner, K.; Daane, A. H. (1958). "The Crystal Structures of Some of the Rare Earth Carbides". Journal of the American Chemical Society. 80 (17): 4499. doi:10.1021/ja01550a017.
  14. 1 2 3 4 5 6 7 8 Greenwood, p. 1241
  15. 1 2 3 4 Greis, O. (1978). "Über neue Verbindungen im system SmF2_SmF3". Journal of Solid State Chemistry. 24 (2): 227. Bibcode:1978JSSCh..24..227G. doi:10.1016/0022-4596(78)90013-0.
  16. 1 2 Meyer, G.; Schleid, T. (1986). "The metallothermic reduction of several rare-earth trichlorides with lithium and sodium". Journal of the Less Common Metals. 116: 187. doi:10.1016/0022-5088(86)90228-6.
  17. Bärnighausen, H. (1973). Rev. Chim. Miner. 10: 77–92.{{cite journal}}: CS1 maint: untitled periodical (link)
  18. Zachariasen, W. H. (1948). "Crystal chemical studies of the 5f-series of elements. I. New structure types". Acta Crystallographica. 1 (5): 265. doi: 10.1107/S0365110X48000703 .
  19. Asprey, L. B.; Keenan, T. K.; Kruse, F. H. (1964). "Preparation and Crystal Data for Lanthanide and Actinide Triiodides" (PDF). Inorganic Chemistry. 3 (8): 1137. doi:10.1021/ic50018a015.
  20. Brown, R.; Clark, N. J. (1974). "Composition limits and vaporization behaviour of rare earth nitrides". Journal of Inorganic and Nuclear Chemistry. 36 (11): 2507. doi:10.1016/0022-1902(74)80462-8.
  21. 1 2 Meng, J.; Ren, Yufang (1991). "Studies on the electrical properties of rare earth monophosphides". Journal of Solid State Chemistry. 95 (2): 346. Bibcode:1991JSSCh..95..346M. doi:10.1016/0022-4596(91)90115-X.
  22. 1 2 Beeken, R.; Schweitzer, J. (1981). "Intermediate valence in alloys of SmSe with SmAs". Physical Review B. 23 (8): 3620. Bibcode:1981PhRvB..23.3620B. doi:10.1103/PhysRevB.23.3620.
  23. Greenwood, p. 1239
  24. Beaurepaire, Eric (Ed.) Magnetism: a synchrotron radiation approach, Springer, 2006 ISBN   3-540-33241-3 p. 393
  25. Emsley, John (2001). "Samarium". Nature's Building Blocks: An A–Z Guide to the Elements. Oxford, England, UK: Oxford University Press. pp.  371–374. ISBN   0-19-850340-7.
  26. Jayaraman, A.; Narayanamurti, V.; Bucher, E.; Maines, R. (1970). "Continuous and Discontinuous Semiconductor-Metal Transition in Samarium Monochalcogenides Under Pressure". Physical Review Letters. 25 (20): 1430. Bibcode:1970PhRvL..25.1430J. doi:10.1103/PhysRevLett.25.1430.
  27. Greenwood, pp. 1236, 1241
  28. Greenwood, p. 1240
  29. Baernighausen, H.; Haschke, John M. (1978). "Compositions and crystal structures of the intermediate phases in the samarium-bromine system". Inorganic Chemistry. 17: 18. doi:10.1021/ic50179a005.
  30. Beck, H. P. (1979). "Hochdruckmodifikationen der Diiodide von Sr., Sm und Eu. Eine neue PbCl2-Variante?". Zeitschrift für anorganische und allgemeine Chemie. 459: 81. doi:10.1002/zaac.19794590108.
  31. Beck, H. P.; Gladrow, E. (1979). "Zur Hochdruckpolymorphie der Seltenerd-Trihalogenide". Zeitschrift für anorganische und allgemeine Chemie. 453: 79. doi:10.1002/zaac.19794530610.
  32. 1 2 Nickerson, J.; White, R.; Lee, K.; Bachmann, R.; Geballe, T.; Hull, G. (1971). "Physical Properties of SmB6". Physical Review B. 3 (6): 2030. Bibcode:1971PhRvB...3.2030N. doi:10.1103/PhysRevB.3.2030.
  33. Nyhus, P.; Cooper, S.; Fisk, Z.; Sarrao, J. (1995). "Light scattering from gap excitations and bound states in SmB6". Physical Review B. 52 (20): 14308–14311. Bibcode:1995PhRvB..5214308N. doi:10.1103/PhysRevB.52.R14308. PMID   9980746.
  34. Sera, M.; Kobayashi, S.; Hiroi, M.; Kobayashi, N.; Kunii, S. (1996). "Thermal conductivity of RB6 (R=Ce, Pr, Nd, Sm, Gd) single crystals". Physical Review B. 54 (8): R5207–R5210. Bibcode:1996PhRvB..54.5207S. doi:10.1103/PhysRevB.54.R5207. PMID   9986570.
  35. Botimer, J.; Kim; Thomas; Grant; Fisk; Jing Xia (2013). "Robust Surface Hall Effect and Nonlocal Transport in SmB6: Indication for an Ideal Topological Insulator". Scientific Reports. 3 (3150): 3150. arXiv: 1211.6769 . Bibcode:2013NatSR...3E3150K. doi:10.1038/srep03150. PMC   3818682 . PMID   24193196.
  36. Zhang, Xiaohang; Butch; Syers; Ziemak; Greene; Paglione (2013). "Hybridization, Correlation, and In-Gap States in the Kondo Insulator SmB6". Physical Review X. 3 (1): 011011. arXiv: 1211.5532 . Bibcode:2013PhRvX...3a1011Z. doi:10.1103/PhysRevX.3.011011. S2CID   53638956.
  37. Wolgast, Steven; Kurdak, Cagliyan; Sun, Kai; et al. (2012). "Low-temperature surface conduction in the Kondo insulator SmB6". Physical Review B. 88 (18): 180405. arXiv: 1211.5104 . Bibcode:2013PhRvB..88r0405W. doi:10.1103/PhysRevB.88.180405. S2CID   119242604.
  38. Gladyshevskii, E. I.; Kripyakevich, P. I. (1965). "Monosilicides of rare earth metals and their crystal structures". Journal of Structural Chemistry. 5 (6): 789. doi:10.1007/BF00744231. S2CID   93941853.
  39. Smith, G. S.; Tharp, A. G.; Johnson, W. (1967). "Rare earth–germanium and –silicon compounds at 5:4 and 5:3 compositions". Acta Crystallographica. 22 (6): 940. doi: 10.1107/S0365110X67001902 .
  40. Yarembash, E. I.; Tyurin, E. G.; Reshchikova, A. A.; et al. (1971). Inorg. Mater. 7: 661–665.{{cite journal}}: CS1 maint: untitled periodical (link)
  41. Greenwood, p. 1248
  42. 1 2 3 Greenwood, p. 1249
  43. Evans, William J.; Hughes, Laura A.; Hanusa, Timothy P. (1986). "Synthesis and x-ray crystal structure of bis(pentamethylcyclopentadienyl) complexes of samarium and europium: (C5Me5)2Sm and (C5Me5)2Eu". Organometallics. 5 (7): 1285. doi:10.1021/om00138a001.