Next Article in Journal
Triple-Mode Switchable Terahertz Metamaterial Absorber with Tunable Absorption Characteristics
Previous Article in Journal
Ambient Pressure Synthesis of Re-Substituted MnGe and Its Magnetic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Reduction Annealing on the Coloration Mechanism of Yellow Sapphire with High Iron Content

1
Laboratory of Gem Materials, College of Jewelry, Shanghai Jianqiao University, No. 1111, Huchenghuan Road, Shanghai 201306, China
2
Research Center of Analysis and Test, East China University of Science & Technology, No. 130, Meilong Road, Shanghai 200237, China
3
Key Laboratory of Transparent and Opto-Functional Inorganic Materials, Chinese Academy of Sciences, Shanghai 200050, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(9), 1257; https://doi.org/10.3390/cryst12091257
Submission received: 13 August 2022 / Revised: 2 September 2022 / Accepted: 3 September 2022 / Published: 5 September 2022
(This article belongs to the Section Mineralogical Crystallography and Biomineralization)

Abstract

:
The color of yellow sapphire from Africa characterized by high iron content and low levels of other transition metal elements was changed from yellow to grayish-blue after high-temperature reduction annealing. Before reduction annealing, the optical absorption spectra showed that the outer dd electron transitions of Fe3+ were the main coloring cause of yellow sapphires, but the charge transfer between O2− and Fe3+ may have a greater contribution. The change in lattice parameter indicates that Fe3+ is reduced to Fe2+ during reduction annealing, and adjacent Fe2+ and Fe3+ form an Fe2+-Fe3+ ion pair. The absorption caused by intervalence charge transfer of Fe2+-Fe3+ is the essential reason for the grayish-blue appearance of yellow sapphires after reduction annealing. The charge compensation mechanism of Fe2+-Fe3+ in natural sapphire is also discussed, and oxygen vacancy is considered to be the most suitable charge compensator for Fe2+-Fe3+.

1. Introduction

The abundance of iron is highest among transition metal element; iron is also a common impurity in minerals such as corundum, beryl, tourmaline and spinel. Iron substitutes Al3+ ions in corundum (α-Al2O3), generally showing two valence states of Fe2+ and Fe3+. Both of these valence states have an important influence on the color of corundum crystals. The isolated Fe2+ ion has exhibits no absorption in the visible region and has a minimal effect on the color of the corundum; however, when it forms an ion pair with Ti4+ or Fe3+, the corundum exhibits a remarkable color. The Fe2+-Ti4+ heteronuclear ion pair was first proposed by Townsend [1], producing two broad intervalence charge transfer (IVCT) absorption bands at 17,800 and 14,200 cm−1, respectively, giving the corundum a beautiful blue color. An Fe2+-Fe3+ homonuclear ion pair is formed by adjacent Fe2+ and Fe3+. Lehmann and Harder [2] studied natural yellow, green and blue sapphires and found that as the proportion of Fe2+/Fe3+ increased, the color of corundum gemstones gradually changed from yellow to blue. However, Lehmann and Harder [2] attributed the 11,400 cm−1 (870 nm) absorption band in the optical absorption spectrum to the 5T25E2 transition of Fe2+. Subsequently, Faye [3] pointed out that the correct assignment of the 11,400 cm−1 absorption band should be the IVCT of Fe2+-Fe3+. In the same year, Ferguson and Fielding [4] also discussed the assignment of the 11,400 cm−1 absorption band in the process of studying iron-doped corundum synthesized by the flux method and attributed the absorption band to Fe2+-O2-Fe3+, which was consistent with Faye’s [3] point of view. Ferguson and Fielding [4] speculated that the charge compensation of Fe2+-Fe3+ in the corundum synthesized by flux method was completed by F, as suggested by Krebs [5]. Ferguson and Fielding [6] also proposed that the charge compensation would be effected by the creation of vacancies in the oxygen sublattice or additional interstitial Al3+ ions in the absence of foreign ions in natural corundum gemstones. Therefore, it is still important to study the charge compensation mechanism of Fe2+-Fe3+ in natural corundum gemstones. In addition to Fe2+-Fe3+ and isolated Fe3+ ions, Fe3+ also interacts with adjacent Fe3+ and O2− in corundum. Exchange-coupled Fe3+ pairs have been shown to have an important effect on the intensity of the 4A1, 4Ea and 4Eb bands of isolated Fe3+ [5]. The charge transfer between O2 and Fe3+ causes three absorption bands in the ultraviolet region of the optical absorption spectrum of synthetic corundum doped with low-content iron (0.02 at% Fe3+, flux-grown) [7]. The tails of the absorption bands extend even further to the violet region of the visible region as the Fe3+ content increases [8]. Therefore, when studying the coloring mechanism of yellow sapphire with high iron content, the charge transfer absorption of O2 and Fe3+ should also be considered.
Yellow is an important color type of natural corundum gemstones, and natural yellow sapphires are produced in various mining areas around the world, such as Myanmar, Sri Lanka, Madagascar, Tanzania, Mozambique, etc. [9,10,11,12]. It is well-known that pure corundum (α-Al2O3) is inherently colorless and exhibits different colors when various transition metal ions replace Al3+ ions [7,13]. The coloration of iron in corundum is relatively complex, and it also easily interacts with other kinds of transition metal ions in natural gemstones. In our research, we obtained a class of yellow sapphires from unknown mining areas in Africa characterized by high iron content and very low levels of other transition metal elements (Figure 1). This type of yellow sapphire is suitable for studying the coloration of iron ions in corundum. In 1993, Fritsch [14] reported two grayish-blue sapphires from Rwanda. The color of these two sapphires was mainly caused by the optical absorption of IVCT of Fe2+-Fe3+, so he proposed that Fe2+-Fe3+ is the second coloring mechanism of blue sapphire, in addition to Fe2+-Ti4+. In this paper, the use of reduction annealing to form new Fe2+-Fe3+ ion pairs in yellow sapphire with high iron content is further investigated with respect to Fritsch’s viewpoint. To the best of our knowledge, research on the formation of Fe2+-Fe3+ in natural sapphire has not been reported to date. Furthermore, the effect of reduction annealing on the coloration mechanism of yellow sapphire is discussed in this paper.

2. Materials and Methods

Yellow sapphire samples were provided by Lin Shaobo (Thailand, Bangkok) from an unknown mining area in Africa and characterized by high iron content and low levels of other transition metal elements. Three oriented yellow sapphire samples were cut, ground and polished on two sides parallel to the c axis of the crystal with a thickness of 1.80, 1.93 and 0.95 mm, respectively. Annealing was carried out twice in a silicon molybdenum rod electric furnace using a reducing atmosphere produced by carbon powder, and the temperature was kept at 1450 °C for 4 h. The sapphire slices were brownish-yellow before reduction annealing and grayish-blue after reduction annealing.
Elemental analysis was performed by laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) on a Resolution M50 deep ultraviolet 193 nm ArF excimer laser ablation instrument from Resonetics, USA, and an Agilent 7900 ICP-MS. Three points were selected for each sample for testing. The diameter of the ablated beam spot was 60 μm, and the signal acquisition time was 60 s. The element content is reported as the average of three measured points.
The optical absorption spectrum and polarized absorption spectrum were scanned with step Δλ = 1 nm in the range of 300–1600 nm with a PerkinElmer Lambda 950 UV-vis-NIR spectrometer with a 150 mm integrating sphere and small spot-focusing attachments. Polarized spectra were acquired using a PE Model B0505284 polarizer matched Lambda 950. The transmission method was used to test the absorbance of the sapphire slices. The slit width of the instrument was set to 2 nm across the entire studied range. Peakfit v4.12 software was used for spectral decomposition, and nonlinear Gaussian fitting was applied. The XRD pattern was recorded on a Philips X’pert PRO X-ray diffractometer equipped with Cu Kα radiation. The diffraction data of 2θ were collected from 10° to 90° with a scan step of 0.026°. The crystal structure of sapphire samples was refined in terms of Rietveld analysis.

3. Results and Discussion

3.1. Transition Metal Elements

The analysis results of transition metal elements of yellow sapphires are shown in Table 1. All transition metal elements are represented in the form of oxides, and T-Fe2O3 represents the total content of FeO and Fe2O3. The content of iron oxide was the highest, at approximately 1 wt.% (10,017 ppm); titanium dioxide, chromium oxide and vanadium trioxide were present at a concentration of 10 ppm or lower; and the content of manganese, nickel, cobalt and other elements was lower than the detection limit of the instrument.

3.2. Reduction Annealing

The three polished slices of yellow sapphires were annealed twice in a reducing atmosphere, each annealing at a temperature of 1450 °C for 4 h. The color of the yellow sapphire slices changed from yellow to grayish-blue after reduction annealing (Figure 2). The thickness of YS-1, YS-2 and YS-3 slices was 1.80, 1.93 and 0.95 mm, respectively. The color of the slice labeled YS-3 was the best after reduction annealing twice. Thickness and polishing quality may have an effect on the color rendering of sapphire slices. One scale of the straightedge placed below the samples in Figure 2 as a reference represents 1 mm.

3.3. Optical Spectroscopy

The analysis of optical absorption spectra of the three yellow sapphire slices before and after reduction annealing is exemplified by slice YS-3 with the best color effect.
Before reduction annealing, the optical absorption spectrum of yellow sapphire mainly has three relatively sharp absorption peaks at 377, 387 and 453 nm (Figure 3a). These three absorption peaks are assigned to the dd electron transitions: 6A14E(4D), 6A14T2(D) and 6A14E14A1(4G) of Fe3+ [12,15]. The dd electron transitions, 6A14E(4D) and 6A14E14A1(4G), are temperature- and concentration-dependent, and their absorption is enhanced by exchange-coupled Fe3+-Fe3+ ion pairs [5]. These absorption peaks have a strong absorption background as a result of charge transfer absorption between O2 and Fe3+ [7], similar to the decomposition spectrum shown in Figure 3b. Although the peak of charge transfer absorption between O2 and Fe3+ is located in the ultraviolet region, its tail extends to the visible light region in the absorption spectrum of the yellow sapphire sample, with an important influence on the coloration of sapphire. Therefore, the color of yellow sapphire with high iron content is mainly contributed by the charge transfer absorption of O2 and Fe3+ and the absorption of the dd electron transition of Fe3+.
In addition to the three absorption peaks of Fe3+, a new strong, broad absorption band with a peak at ~855 nm appears in the near-infrared region in the optical absorption spectrum (Figure 3a) after reduction annealing, which is similar to the absorption caused by the IVCT of Fe2+-Fe3+ [14,16]. The optical absorption spectrum of YS-3 after reduction annealing was fitted, as shown in Figure 3b. The intensity and full width at half maximum (FWHM) of each absorption band obtained by fitting are shown in Table 2. An absorption band at 855 nm (11,690 cm−1) observed in blue sapphire from Thailand was mentioned as early as Lehmann and Harder’s article [2] but was assigned to Fe2+. Iron-doped corundum crystals grown by the flux method showed uneven blue color [5], and the optical absorption spectra of these crystals also had a strong absorption band (11,400 cm−1) in the near-infrared region but was assigned to Fe2+ by the author, which is consistent with the viewpoint of Lehmann and Harder [2]. In the same year, Ferguson and Fielding [4] pointed out that this band should be attributed to Fe2+-O2-Fe3+ instead of Fe2+. Then, in 1993, Fritsch [14] observed this absorption band in two grayish-blue sapphires from Rwanda and proposed that Fe2+-Fe3+ is the second coloring mechanism, in addition to Fe2+-Ti4+, that makes sapphire blue. In the polarized absorption spectra (Figure 3c), the intensity of the absorption band at 855 nm is greater in the E⊥C direction than in the E∥C direction, indicating that it is polarized along the vector between the interacting cations. Large band half-width, intense polarization dependence along the metal-to-metal bond direction and high absorbance are features of the 855 nm absorption band, consistent with the identifying characteristics of the IVCT absorption band proposed by Mattson and Rossman [17]. Therefore, it can be inferred that the 855 nm absorption band is caused by the intervalence charge transfer of Fe2+-Fe3+, meaning that Fe2+-Fe3+ is newly formed in sapphire samples during the reduction annealing process and causes the color of sapphire to change from yellow to grayish-blue.
The center position of the IVCT absorption band of Fe2+-Fe3+ differs depending on the reference: for example, 11,400 cm−1 (877 nm) [2], 11,300 cm−1 (885 nm) [5] and 890 nm [14]. In α-Al2O3 crystals, adjacent [AlO6] octahedrons have two relative positional relationships: a face-sharing relationship in the direction parallel to the c axis and an edge-sharing in relationship in the direction perpendicular to the c axis. The distance between Al3+ and Al3+ ions is not equal in all directions. The calculation results show that the IVCT energies of Fe2+-Fe3+ are 1.43 ev (867 nm) for face sharing and 1.30 ev (953 nm) for edge sharing [18]. Therefore, the IVCT absorption band of Fe2+-Fe3+ should have two different center positions, similar to the IVCT absorption band of Fe2+-Ti4+. We hypothesize that this is the reason why the center position of the IVCT absorption band of Fe2+-Fe3+ is differs depending on the reference.
In addition to the absorption caused by IVCT of Fe2+-Fe3+ and charge transfer between O2 and Fe3+, there are two regions in the fitted spectrum with no further fitting, represented by the dashed line in Figure 3b. First, yellow sapphire contains a small amount of titanium, some of which can be attributed to the IVCT absorption band of Fe2+-Ti4+. This means that Fe2+ formed during reduction annealing interacts with both Fe3+ and Ti4+. Second, yellow sapphire also contains a minute amount of Cr3+ and Ti3+, which may form during the reduction annealing process. Therefore, the unfitting absorption of the area marked by the dashed line in Figure 3b may be caused by Fe2+-Ti4+, Cr3+ and Ti3+ [1,19,20]. Third, these unfit absorption regions may also be co-contributed by F, F+, F2, F 2 + and F 2 2 +   [21]—a possibility that cannot be ruled out.

3.4. Analysis of Lattice Parameter

The space symmetry group of corundum crystal (α-Al2O3) is R 3 c (No.167), and the lattice parameter is a(Å) = 4.76 095, c(Å) = 12.9 962. Figure 4 shows the results of the crystal structure refined in terms of Rietveld analysis of YS-3 before and after reduction annealing, where represents the difference between the measured and calculated intensities of XRD patterns. All of the diffraction peaks of sapphire YS-3 are in agreement with those of pure α-Al2O3 (ICSD #10-0173), that is, the structure of yellow sapphire did not change before and after reduction annealing (Table 3). The average fitted lattice parameters of YS-3 are a(Å) = 4.76 502, c(Å) = 13.0 038 before reduction annealing and a(Å) = 4.76 245, c(Å) = 13.0 017 after reduction annealing. Compared with those of pure α-Al2O3, the average lattice parameters of YS-3 are larger as a result of the substitution of Al3+ by Fe2+ and Fe3+ having a larger ionic radius (Table 4).
The ionic radius of Fe3+ in a six-coordinate octahedral crystal field is 0.055 nm, corresponding to the LS state, whereas a radius of 0.0645 nm corresponds to the HS state [22]. In the octahedral crystal field, Dq/B represents the intensity of the crystal field. Dq is one-tenth of splitting energy, and B is the Racah parameter in the crystal field theory. Dq/B > 2.5 is considered a strong field, and Dq/B < 2.5 is a weak field [23]. The optical absorption spectrum of α-Al2O3:Fe3+ crystal shows that Dq = 1510cm−1, B = 660 cm−1 and Dq/B = 2.32 < 2.5, so Fe3+ is in a weak field and takes a high-spin state [5]. Therefore, the ionic radius of Fe3+ should be 0.0645 nm instead of 0.055 nm [22]. Fe2+ is generally in a strong field and a low-spin state, and its ionic radius is 0.61 nm [22,23] (Table 4). Refinement results of single-crystal X-ray diffraction data of YS-3 show that lattice distortion is reduced after reduction annealing, so it can be inferred that a portion of Fe3+ is reduced to Fe2+ during the reduction annealing process. Furthermore, the ionic radius of iron is changed from 0.065nm (Fe3+) to 0.061nm (Fe2+). Then, Fe2+-Fe3+ and Fe2+-Ti4+ are generated by adjacent Fe2+ and Fe3+ and by Fe2+ and Ti4+, respectively, which is consistent with the test results of the above-mentioned optical absorption spectra.

3.5. Charge Compensation

The positive and negative charges in corundum crystal must be balanced. In the Fe2+-Fe3+ ion pairs, a charge imbalance occurs after Fe2+ replaces Fe3+. Therefore, the formation of Fe2+-Fe3+ is inevitably accompanied by the generation of a new charge compensator. Ferguson and Fielding [4] proposed that the charge compensation mechanism for corundum crystals synthesized by flux method with PbF2 and B2O3 as fluxes is F substitution of O2−; however, this explanation does not apply to synthetic corundum with PbO and B2O3 as fluxes. Nikolskaya [24] discussed the charge compensation mechanism for Fe2+ at the Al3+ site in natural sapphire: (i) oxygen vacancy (VO), (ii) H+ or Ti4+ and (iii) interstitial ion Fe i 2 + . The H in corundum may be present when the crystal is formed, or it may be introduced through the heat treatment process [25]. H forms OH groups with neighboring O, which can sometimes be observed via IR spectroscopy, and the introduction and escape of OH groups are reversible [26,27]. The IR spectra of OH in corundum are very complex as a result of many factors, such as the lattice occupancy of H and its adjacent transition metal ions [26,27,28,29]. Some scholars have proposed that the critical absorption peak of OH at 3310 cm−1 is associated with redox reactions involving iron, whereas more studies have shown that this infrared absorption peak is mainly related to titanium ions or defect clusters containing titanium ions [27,28,30,31,32]. The infrared absorption peaks of OH are not obvious with lower concentrations of titanium is [32]. The titanium content was also very low in the African yellow sapphire samples we tested. Furthermore, only trace amounts of H+ are present in the form of OH- in natural corundum gemstones [30]. In addition, the crystal lattice structure of the corundum crystal is tight, and Fe2+ cannot easily exist in the form of interstitial ions. Our heat treatment experiments were carried out in a reducing atmosphere created with carbon powder without the use of hydrogen. Therefore, we propose that the oxygen vacancies should be the most important charge compensator for Fe2+ in high-iron-content African yellow sapphire after reduction annealing. Xiang [33] studied the stability of each defect in corundum under oxygen-poor conditions and reported the order as VO > VAl > Ali > Oi > AlO > OAl, supporting our inference. The reactions that occur during reduction annealing are shown in Formulas (1) and (2):
O O X V ö + 2 e + 1 2 O 2  
2 ( Fe   Al 3 + Fe   Al 3 + ) + V ö + 2 e ( Fe   Al 2 + Fe   Al 3 + ) V ö ( Fe   Al 2 + Fe   Al 3 + )
Under reducing conditions, oxygen vacancies VO are the most suitable charge compensators formed during the reduction annealing process, possibly forming a (Fe2+-Fe3+) VO (Fe2+-Fe3+) defect cluster with Fe2+-Fe3+ ion pairs; therefore, it is beneficial for the Fe2+-Fe3+ ion pair to remain stable.

4. Conclusions

In African yellow sapphires with high iron content, the absorption of charge transfer between O2 and Fe3+ and dd electron transitions of Fe3+ are the two main causes of the sapphire’s yellow appearance. After high-temperature reduction annealing, a portion of Fe3+ in yellow sapphire is reduced to Fe2+; then, Fe2+-Fe3+ ion pairs are formed by the adjacent Fe2+ and Fe3+. The oxygen vacancy (VO_ is a charge compensator of Fe2+-Fe3+ ion pairs in natural sapphire, forming a (Fe2+-Fe3+) VO (Fe2+-Fe3+) defect cluster with Fe2+-Fe3+, which helps to maintain the stability of Fe2+-Fe3+. The IVCT absorption band of Fe2+-Fe3+ is centered at 855 nm in the near-infrared region and is characterized by a large band half-width, intense polarization dependence along the metal-to-metal bond direction and high absorbance. The absorption of IVCT of Fe2+-Fe3+ covering the entire visible region and the absorption of charge transfer between O2 and Fe3+ are the two main causes of the grayish-blue appearance of African yellow sapphires after reduction annealing.
Although the 855 nm band has been studied for decades, scholars have differing views. In the early studies, some scholars held the viewpoint that the absorption band at 810 nm in beryl was generated by the octahedral site Fe2+ and once believed that the band near 855 nm in corundum was also generated by the spin-allowed transition of Fe2+. Recently, some scholars used this viewpoint to explain the coloring mechanism of green sapphire. In recent years, few articles have been published focusing on the study of Fe2+-Fe3+ ion pairs in sapphire, and the misassignment of the 855 nm band is still an issue. Through our heat treatment experiment of natural sapphire with high iron content, we observed the color change of sapphire before and after heat treatment and the appearance of a strong absorption band at 855 nm, effectively supporting the viewpoint proposed by Fritsch in 1993 from the perspective of experimental data and actual phenomena. In addition, Fe3+ in corundum is in a weak field and takes a high-spin state, so its ion radius should be 0.065 nm, not 0.055 as commonly cited.
We know that dark-blue sapphires in many mines, such as Changle in Shandong Province, Thailand, and Australia, have a very high iron content, and their color is mainly caused by Fe2+-Fe3+. Over the years, the value of dark-blue sapphire has not been fully developed due to the lack of superior color improvement processes. Our study of the reduction annealing of yellow sapphire with high iron content may provide a useful reference with respect to color improvement of these dark-blue sapphires.

Author Contributions

Conceptualization, X.H. and X.F.; experimental studies, data acquisition and analysis, and writing and editing, X.W. and X.H.; data acquisition, Y.K.; review, X.F. and S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Townsend, M.G. Visible charge transfer band in blue sapphire. Solid State Commun. 1968, 6, 81–83. [Google Scholar] [CrossRef]
  2. Lehmann, G.; Harder, H. Optical spectra of di-and trivalent iron in corundum. Am. Mineral. 1970, 55, 98–105. [Google Scholar]
  3. Faye, G.H. On the optical spectra of di-and trivalent iron in corundum: A discussion. Am. Mineral. 1971, 56, 344–348. [Google Scholar]
  4. Ferguson, J.; Fielding, P.E. The origin of the colours of yellow, green and blue sapphires. Chem. Phys. Lett. 1971, 10, 262–265. [Google Scholar] [CrossRef]
  5. Krebs, J.J.; Maisch, W.G. Exchange effects in the optical-absorption spectrum of Fe3+ in Al2O3. Phys. Rev. B 1971, 4, 757–769. [Google Scholar] [CrossRef]
  6. Ferguson, J.; Fielding, P.E. The origins of the colours of natural yellow, blue, and green sapphires. Aust. J. Chem. 1972, 25, 1371–1385. [Google Scholar] [CrossRef]
  7. Tippins, H.H. Charge-transfer spectra of transition-metal ions in corundum. Phys. Rev. B 1970, 1, 126–135. [Google Scholar] [CrossRef]
  8. Marfunin, A.S.; Egorova, N.G. Physics of Minerals and Inorganic Materials: An Introduction; Li, G.S., Translator; Geological Publishing House: Beijing, China, 1984; pp. 170–177. (In Chinese) [Google Scholar]
  9. Zwann, P.C. Sri-Lanka: The gem island. Gems. Gemol. 1982, 18, 62–71. [Google Scholar] [CrossRef]
  10. Kiefert, L.; Schmetzer, K. Blue and yellow sapphire from Kaduna Province, Nigeria. J. Gemmol. 1987, 20, 427–442. [Google Scholar] [CrossRef]
  11. Rakotondrazafy, A.F.M.; Giuliani, G.; Ohnenstetter, D.; Fallick, A.E.; Rakotosamizanany, S.; Andriamamonjy, A.; Ralantoarison, T.; Razanatseheno, M.; Offant, Y.; Garnier, V.; et al. Gem corundum deposits of Madagascar: A review. Ore Geol. Rev. 2008, 34, 134–154. [Google Scholar] [CrossRef]
  12. Sutherland, F.L.; Zaw, K.; Meffre, S.; Yui, T.F.; Thu, K. Advances in trace element “fingerprinting” of gem corundum, ruby and sapphire, Mogok area, Myanmar. Minerals 2014, 5, 61–79. [Google Scholar] [CrossRef]
  13. McClure, D.S. Optical spectra of transition-metal ions in corundum. J. Chem. Phys. 1962, 36, 2757–2779. [Google Scholar] [CrossRef]
  14. Fritsch, E. Blue color in sapphire caused by Fe2+/Fe3+ intervalence charge transfer. Gems. Gemol. 1993, 29, 151. [Google Scholar]
  15. Schmetzer, K.; Bosshart, G.; Hänni, H.A. Natural coloured and treated yellow and orange-brown sapphires. J. Gemmol. 1983, 18, 607–622. [Google Scholar] [CrossRef]
  16. Han, X.Z.; Guo, S.G.; Kang, Y.; Feng, X.Q.; Li, Y.S. Colorizing Role of Iron and Titanium in Dark Blue Sapphire from Changle, Shandong. J. Chin. Ceram. Soc. 2018, 46, 1483–1488. (In Chinese) [Google Scholar]
  17. Mattson, S.M.; Rossman, G.R. Identifying characteristics of charge transfer transitions in minerals. Phys. Chem. Miner. 1987, 14, 94–99. [Google Scholar] [CrossRef]
  18. Bristow, J.K.; Tiana, D.; Parker, S.C.; Walsh, A. Defect chemistry of Ti and Fe impurities and aggregates in Al2O3. J. Mater. Chem. A 2014, 2, 6198–6208. [Google Scholar] [CrossRef]
  19. Garcia-Lastra, J.M.; Barriuso, M.T.; Aramburu, J.A.; Moreno, M. Origin of the different color of ruby and emerald. Phys. Rev. B 2005, 72, 113104. [Google Scholar] [CrossRef]
  20. Moulton, P.F. Spectroscopic and laser characteristics of Ti:Al2O3. J. Opt. Soc. Am. B 1986, 3, 125–133. [Google Scholar] [CrossRef]
  21. Ramírez, R.; Tardío, M.; González, R.; Santiuste, J.E.M. Optical properties of vacancies in thermochemically reduced Mg-doped sapphire single crystals. J. Appl. Phys. 2007, 101, 123520. [Google Scholar] [CrossRef]
  22. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta. Cryst. B. 1969, 25, 925–946. [Google Scholar] [CrossRef]
  23. Tian, Y.J.; Cao, M.S.; Cao, C.B. Introduction of advanced materials; Harbin Institute of Technology Press: Harbin, China, 2014; p. 271. (In Chinese) [Google Scholar]
  24. Nikolskaya, L.V.; Terekhova, V.M.; Samoilovich, M.I. On the origin of natural sapphire color. Phys. Chem. Miner. 1978, 3, 213–224. [Google Scholar] [CrossRef]
  25. Müller, R.; Günthard, Hs. H. Spectroscopic study of reduction Nickel and Cobalt ions in sapphire. J. Chem. Phys. 1966, 44, 365–373. [Google Scholar] [CrossRef]
  26. Volynets, F.K.; Vorobev, V.G.; Sidorova, E.A. Infrared absorption bands in corundum crystals. J. Appl. Spectrosc. 1969, 10, 665–667. [Google Scholar] [CrossRef]
  27. Eigenmann, K.; Günthard, H.H. Hydrogen incorporation in doped α-Al2O3 by high temperature redox reactions. Chem. Phys. Lett. 1971, 12, 12–15. [Google Scholar] [CrossRef]
  28. Moon, A.R.; Phillips, M.R. Defect clustering in H, Ti:α-Al2O3. J. Phys. Chem. Solids 1991, 52, 1087–1099. [Google Scholar] [CrossRef]
  29. Beran, A. Trace hydrogen in Verneuil-grown corundum and its colour varieties—An IR spectroscopic study. Eur. J. Mineal. 1991, 3, 971–976. [Google Scholar] [CrossRef]
  30. Beran, A.; Rossman, G.R. OH in naturally occurring corundum. Eur. J. Mineral. 2006, 18, 441–447. [Google Scholar] [CrossRef]
  31. Phlayrahan, A.; Monarumit, N.; Boonmee, C.; Satitkune, S.; Wathanakul, P. Fe oxidation state in heat-treated basaltic blue sapphire samples and its implication to the 3309 cm−1-series peaks in infrared absorption spectra. In Proceedings of the Siam Physics Congress 2018: A Creative Path to Sustainable Innovation, Pitsanulok, Thailand, 21–23 May 2018. [Google Scholar]
  32. Phlayrahan, A.; Monarumit, N.; Satitkune, S.; Wathanakul, P. Role of Ti content on the occurrence of the 3309-cm−1 peak in FTIR absorption spectra of ruby samples. J. Appl. Spectrosc. 2018, 85, 385–390. [Google Scholar] [CrossRef]
  33. Xiang, X.; Zhang, G.K.; Tang, T. Fe effect on the process of intrinsic point defects in α-Al2O3. J. Mater. Sci. 2018, 53, 11194–11203. [Google Scholar] [CrossRef]
Figure 1. Yellow sapphires from unknown mining areas in Africa.
Figure 1. Yellow sapphires from unknown mining areas in Africa.
Crystals 12 01257 g001
Figure 2. Yellow sapphire slices before and after reduction annealing. Left, before annealing; right, after annealing.
Figure 2. Yellow sapphire slices before and after reduction annealing. Left, before annealing; right, after annealing.
Crystals 12 01257 g002
Figure 3. Optical absorption spectra of sapphire samples: (a) comparison of optical absorption spec-tra before and after reduction annealing; (b) fitting spectra of the optical absorption spectrum of sapphire after reduction annealing; (c) polarized absorption spectra of sapphire after reduction an-nealing.
Figure 3. Optical absorption spectra of sapphire samples: (a) comparison of optical absorption spec-tra before and after reduction annealing; (b) fitting spectra of the optical absorption spectrum of sapphire after reduction annealing; (c) polarized absorption spectra of sapphire after reduction an-nealing.
Crystals 12 01257 g003aCrystals 12 01257 g003b
Figure 4. Rietveld refinement results of yellow sapphire obtained from XRD data: (a) before reduction annealing; (b) after reduction annealing.
Figure 4. Rietveld refinement results of yellow sapphire obtained from XRD data: (a) before reduction annealing; (b) after reduction annealing.
Crystals 12 01257 g004
Table 1. Test results of transition metal elements of yellow sapphire (ppm wt).
Table 1. Test results of transition metal elements of yellow sapphire (ppm wt).
OxideTiO2V2O3Cr2O3MnOT-Fe2O3CoONiOCuOGa2O3
ContentYS-19.22.327.1<110,425<1<1<111.5
YS-211.31.921.6<111,018<1<1<112.3
YS-38.82.616.4<110,017<1<1<111.2
Table 2. Intensity and FWHM of absorption bands obtained by fitting.
Table 2. Intensity and FWHM of absorption bands obtained by fitting.
Band center / cm−126,55025,84022,17011,690
Absorption coefficient / cm−11.7802.8873.2233.249
FWHM / cm−16831,0771,4926,153
Table 3. Diffraction data of yellow sapphire from Rietveld refinement results.
Table 3. Diffraction data of yellow sapphire from Rietveld refinement results.
(h k l)Before Reduction AnnealingAfter Reduction Annealing
d/nm2θ/od/nm2θ/o
(0 1 12)0.3490625.4970.3489125.508
(1 0 4)0.2555335.0880.2554735.098
(1 1 0)0.2383337.7130.2382937.72
(1 1 3)0.2082943.4090.2090743.239
(0 2 4)0.1741452.5050.1741152.516
(1 1 6)0.1604157.3980.1601957.481
(2 1 1)0.1549759.6110.1548659.658
(0 1 8)0.1511861.2640.1512761.221
(2 1 4)0.1407366.3690.1406266.429
(3 0 0)0.1375568.1110.1374668.164
(1 2 5)0.1335370.4620.1337370.338
(1 0 10)0.1240676.7660.1240276.796
(1 1 9)0.1234377.230.1235877.116
(2 2 0)0.1191480.5640.1190880.611
(3 0 6)0.1161183.1190.1159883.236
(2 2 3)0.1148684.2280.114884.289
(0 2 10)0.1100188.880.1099688.934
Table 4. The state and ionic radius of iron ions in corundum (α-Al2O3).
Table 4. The state and ionic radius of iron ions in corundum (α-Al2O3).
Valence StateCoordination NumberCrystal Field StateSpin StateIonic Radius (nm)
Fe3+(3d5)6weak-fieldhigh-spin0.065
Fe2+(3d6)6strong-fieldlow-spin0.061
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, X.; Han, X.; Kang, Y.; Feng, X.; Guo, S. Effect of Reduction Annealing on the Coloration Mechanism of Yellow Sapphire with High Iron Content. Crystals 2022, 12, 1257. https://doi.org/10.3390/cryst12091257

AMA Style

Wu X, Han X, Kang Y, Feng X, Guo S. Effect of Reduction Annealing on the Coloration Mechanism of Yellow Sapphire with High Iron Content. Crystals. 2022; 12(9):1257. https://doi.org/10.3390/cryst12091257

Chicago/Turabian Style

Wu, Xiao, Xiaozhen Han, Yan Kang, Xiqi Feng, and Shouguo Guo. 2022. "Effect of Reduction Annealing on the Coloration Mechanism of Yellow Sapphire with High Iron Content" Crystals 12, no. 9: 1257. https://doi.org/10.3390/cryst12091257

APA Style

Wu, X., Han, X., Kang, Y., Feng, X., & Guo, S. (2022). Effect of Reduction Annealing on the Coloration Mechanism of Yellow Sapphire with High Iron Content. Crystals, 12(9), 1257. https://doi.org/10.3390/cryst12091257

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop