Next Article in Journal
Effect of Solid/Liquid and Eutectic Front Velocities on Microstructure Evolution in Al-20%Cu Alloys
Previous Article in Journal
Effect of Acetylation on the Morphology and Thermal Properties of Maize Stalk Cellulose Nanocrystals: A Comparative Study of Green-Extracted CNC vs. Acid Hydrolysed Followed by Acetylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gemological Characteristics of Blue-Violet Cordierite

School of Gemology, China University of Geosciences, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(7), 637; https://doi.org/10.3390/cryst14070637
Submission received: 11 June 2024 / Revised: 2 July 2024 / Accepted: 2 July 2024 / Published: 10 July 2024
(This article belongs to the Section Mineralogical Crystallography and Biomineralization)

Abstract

:
Cordierite is a violet-blue gem mineral primarily composed of magnesium aluminum silicate. This study employed three samples of Mg-cordierite and conducted tests on their gemological characteristics, spectroscopic features, and chemical composition using Fourier transform infrared spectroscopy (FTIR), Raman spectroscopy, ultraviolet spectrophotometry, X-ray photoelectron spectroscopy (XPS), and electron microprobe. The study also explored and analyzed the polychroism and coloration mechanisms of the samples. The results indicate that the lattice vibrations of the Mg-cordierite samples differ along the directions parallel to the a, b, and c crystallographic axes, leading to certain variations in spectral characteristics among these directions. The article provides experimental evidence for the reasons for the polychroism of cordierite in different crystal axes, which is of great significance in the quality evaluation of cordierite.

1. Introduction

Purple-series gemstones have always been widely popular in the gem world for their rarity and mystery. However, research on purple gemstones has largely focused on amethyst, purple sapphires, and others. Comparatively, less attention has been paid to the gemological study of blue-violet cordierite. Cordierite has numerous sources, with gem-quality cordierite primarily distributed in India, Sri Lanka, Madagascar, the United States (California, Idaho, and Wyoming), Canada, and Greenland [1]. Cordierite is an orthorhombic crystal system mineral with a vitreous luster, potentially containing trace elements such as Na, K, Ca, Fe, Mn, and H2O. Among these, Mg and Fe can form a complete isomorphous series. As the Fe content increases, the color deepens, along with corresponding increases in refractive index and density. Cordierite is found in metamorphic rocks, granulites, and gneisses, and cordierite coexists with other magnesium facies, such as garnet, biotite, orthopyroxene, spinel, or hornblende. It also coexists with aluminum-containing minerals, such as sillimanite and corundum [2]. Cordierite with blue-violet color is the most valuable among cordierites, and it shares many similarities in appearance and gemological properties with amethyst [3], sapphire [4], and tanzanite [5], making them easily confusable. The consensus regarding the bulk color and strong trichroism of cordierite attributes it to the charge transfer between Fe3+ and Fe2+ [6]. However, the valence state distribution of Fe ions in cordierite and the connection between the crystallographic axis direction and polychroism have not yet been explained.
Current research on cordierite is largely focused on the fields of materials and ceramics [7,8], with fewer studies on its gem mineralogic characteristics [9,10]. Charles and others have explained the coloration mechanism of cordierite, stating that its blue color originates from the charge transfer between Fe2+ in the octahedral sites and Fe3+ in the T11 tetrahedral sites [11]. Dudka et al. [12] conducted precise X-ray diffraction analysis on cordierite, identifying the locations of metal cations Na+, H2O, and CO2 within the channels of cordierite. Pollak [13] explained the sites of charge transfer in cordierite by examining the differences between infrared absorption bands along different crystallographic axes, providing a deeper explanation for its polychroism. However, specific studies on the spectral differences between different crystal axes are lacking. Duncan et al. [14] measured the Mössbauer spectra of cordierite single crystals in different directions, determining the positions and occupancy rates of Fe2+ and Fe3+ ions, suggesting that the exchange interaction between Fe3+ ions on the T1 sites and the six-fold sites is the source of cordierite’s polychroism. Raphaël et al. [15] evaluated the spectroscopic properties of cordierite through TD-DFT, assessing the simulation of polychroism and the variation of color with different light sources. However, studies on the spectral differences along different crystallographic axes of cordierite and the content ratio of Fe2+ to Fe3+ are still lacking.
This paper investigates the coloration mechanism and the causes of polychroism in purple cordierite through conventional gemological testing, Fourier Transform Infrared Spectroscopy (FTIR), Raman Spectroscopy, Ultraviolet-Visible Spectrophotometry (UV–vis), X-ray Photoelectron Spectroscopy (XPS), and Electron Probe Microanalysis (EPMA). It elucidates the spectral differences of cordierite along different crystallographic axes and their impact on its polychroism. Additionally, the study determines the valence state and content ratio of Fe in cordierite through XPS, enhancing the related research on cordierite and providing experimental evidence for its gemological characteristics and identification.

2. Materials and Methods

2.1. Sample Materials

Figure 1 displays three samples of blue and purple cordierite from unknown locations in Madagascar, which were selected and purchased from local suppliers. All samples exhibit a fragmented appearance with evident polychroism. Notable cracks are present internally, and distinct inclusions can be observed.

2.2. Experimental Methods

Conventional Gemological Testing: The refractive index, specific gravity, ultraviolet fluorescence characteristics, and microscopic features of the samples were tested using a refractometer, hydrostatic weighing method, ultraviolet light, and gemological microscope, respectively. The testing was conducted at the Gemology Experimental Teaching Center of the Gemology College, China University of Geosciences, Beijing (CUGB).
The samples O1, O2, and O3 were cut into cubic blocks along their lattice directions and polished. The three central lines of the cube pass through the center point of the squared polished surface and are perpendicular to each other, with the directions of the three central lines parallel to crystallographic axes a, b, and c, respectively. By systematically testing and analyzing the UV-visible spectroscopy, infrared spectroscopy, and Raman spectroscopy of the cordierite samples in the three crystallographic axis directions, combined with XPS analysis and electron microprobe data, the study aims to reveal the polychroism and coloration causes of cordierite.
Raman spectra were obtained using the HR-Evolution micro-Raman spectrometer, manufactured by HORIBA, Japan, at the Gem Testing Laboratory within the School of Gemmology at CUGB. A microscope equipped with ×50 objectives was used to focus on the samples under precise test conditions: aperture set to 0.865 μm; power output maintained at 100 mW; the spectral range of 200–4000 cm−1; integration time of 10 s; and laser wavelength of 532 nm.
The infrared spectra of the samples were obtained using the Tensor 27 FT-IR spectrometer produced by Bruker, Berlin, Germany. The tests were carried out at the Gem Testing Laboratory of the Gemology College, CUGB, with a scanning range of 400–2000 cm−1 (reflection), a resolution of 4 cm−1, and a scan time of 32 s. In addition, the samples were tested using the powder transmission method. The samples were divided into O1, O2, and O3, respectively, and then refined to 200 mesh with an agate mortar after cleaning. The mineral powder is mixed with analytically pure potassium bromide powder, and the ratio is about 1:100. After being fully mixed, the tablet is pressed on the oil press, and then the test is carried out.
Ultraviolet-visible absorption spectra were collected using the UV-3600 UV-Vis-NIR spectrophotometer, produced by Shimadzu Corporation, Kyoto, Japan, at the Gem Testing Laboratory of the Gemology College, CUGB. The test conditions were as follows: transmission method, wavelength range of 300~800 nm, slit width of 20 nm, and a sampling interval of 1.0 s.
EMPA analysis JXA-8230 was used for qualitative and quantitative analysis of major elements in samples at the Geochemical Test Center of China University of Geosciences (Beijing). EMPA uses X-ray energy dispersive spectrometry (EDS) for qualitative and quantitative analysis of the major elements in the sample. Samples were prepared by mounting them on 25 × 75 mm silica glass sheets and polished to 30 μm. The accelerating voltage was 15 kV, the current was 10 nA, and the beam spot diameter was 5 μm. ZAF3 was used to correct the data, and the principal elements and standards tested were Na-Albite, Fe-Garnet, Mg-Olivine, Mn-Rhodonite, Si-Plagioclase, K-Sanidine, and Al-Plagioclase.
X-ray Photoelectron Spectroscopy (XPS) analysis was performed using a Thermo K-Alpha XPS instrument (Thermo Fisher, Waltham, MA, USA). The X-ray excitation source utilized monochromatic Al K-Alpha, with a sample test area of 500 μm and a scanning range of 0–1400 eV at a step size of 1 eV per step, with one scan repetition. The transmission rate of the energy analyzer was set to 100 eV. Data processing was carried out using the Avantage 5.99 software, and the binding energies were calibrated using the standard C1s peak at 284.8 eV.

3. Results and Discussion

3.1. Conventional Gemological Characteristics

The cordierite samples in this paper have a refractive index of 1.535–1.555, a birefringence of 0.015–0.020, and a relative density range of 2.574–2.618 (Table 1), which is consistent with the theoretical range of optical parameters for cordierite, showing fluorescence inertia.
Crystal inclusions are common in cordierite, such as hematite, apatite, tourmaline, etc. [16]. Upon magnified observation, various mineral inclusions can be seen in the cordierite samples (Figure 2). Figure 2 shows the inclusions and their Raman spectra in samples O1, O2, and O3. In sample O1, golden platy inclusions (Figure 2a) and irregular blocky dark inclusions (Figure 2b) are observed. By comparing with the standard Raman spectra in the RUFF database, it is identified that the golden platy inclusions are phlogopite, and the dark inclusions are apatite; Sample O2 contains numerous brownish-red inclusions of hematite and limonite, which appear as hexagonal platy crystals, and additionally, euhedral to subhedral prismatic crystalline inclusions are observed, confirmed by Raman analysis to be tourmaline (Figure 2c); In sample O3, dark striped inclusions are visible, often exposed on the surface (Figure 2d), and Raman testing determined them to be actinolite.

3.2. Main Chemical Composition

The samples O1, O2, and O3 to be measured were prepared as electron microprobe thin sections, with ten points selected on each thin section for electron microprobe measurement. The measurement results (Table 2) show that the cordierite samples have an average MgO content of 9.96–11.64 wt.%, an average SiO2 content of 51.38–56.66 wt.%, and an average FeO content of 1.68–4.60 wt.%. Since volatile substances such as H2O and CO2, as well as cations like Na and K, are present in the channels, the total measured chemical composition data are usually less than 100 wt.% [17]. In the chemical analysis of different samples (Table 2), the Mg content of the samples ranges from 1.494 to 1.725 a.p.f.u. The XMg values (XMg = Mg/(Mg + Fe)) of the current samples were calculated based on the a.p.f.u data, ranging between 0.794 and 0.925, indicating that the samples are Mg-cordierite.
Cordierite is a cyclic silicate mineral composed of different types of tetrahedra and octahedra, with its chemical formula expressed as (M)2(T11)2(T26)2(T23)2(T21)2(T16)O18,(H2O, CO2). The crystal structure is shown in Figure 3.
Six T2 tetrahedra (including four Si-O tetrahedra and two Al-O tetrahedra) are connected at their corners to form the hexagonal ring structure of cordierite, creating a wide channel along the c-axis. This channel often contains large radius cations such as Na+, as well as volatile components like H2O and CO2. The adjacent hexagonal rings are connected by T2 tetrahedra and are staggered by approximately 30°. Octahedra share edges with T1 tetrahedra while also sharing O2− with the T2 tetrahedra in the hexagonal rings [18].
In cordierite, the tetrahedral T1 and T2 sites are typically occupied by smaller radius cations such as Be, Al3+, and Si4+. Among these, Al3+ occupies both the T11 and T26 sites, while Si4+ occupies the T16, T21, and T23 sites. The octahedral M site in cordierite is usually occupied by medium-sized cations, such as Mg2+, Fe2+, and Mn2+ [14].
Using the oxygen atom method with O=18 as the basis, the chemical formula of the cordierite sample can be calculated (Table 3).

3.3. Ultraviolet-Visible Spectrum Analysis

To systematically study the color and pleochroism of cordierite, this paper processes the purple cordierite samples into nearly equal-sized cubes. The polished faces of the cubic blocks have their perpendicular lines parallel to the three crystallographic axes, respectively. The processed cordierite exhibits strong pleochroism (Figure 4), with the three samples appearing nearly colorless or light yellow in the parallel a-axis direction, light purple in the parallel c-axis direction, and uniformly dark purple in the parallel b-axis direction.
This paper conducts ultraviolet-visible light spectroscopic analysis from three crystallographic axes directions, respectively. The results show that the ultraviolet spectra (Figure 5) of the three samples are consistent with the spectral characteristics of standard cordierite samples, with certain differences in absorption intensity and band positions in different directions. We discovered that the ultraviolet spectral characteristics of cordierite are relatively consistent when measured in the parallel a-axis and c-axis directions, whereas the absorption intensity of the ultraviolet spectrum in the parallel b-axis direction differs significantly from those observed in the other two test directions.
A weak absorption band appears in the red region of the ultraviolet spectrum in the parallel b-axis direction, transmitting a small amount of red and blue light. A broad absorption band centered at 584 nm emerges in the orange-yellow region, which can absorb a large amount of yellow and orange-red light and transmit a significant amount of violet light, resulting in the cordierite exhibiting a rich blue-purple color in the parallel b-axis direction. The ultraviolet spectrum measured in the c-axis direction shows that the absorption band centered at 584 nm is much weaker than in the b-axis direction, hence presenting a light purple appearance. In contrast, the ultraviolet spectrum measured in the a-axis direction, compared to the b-axis direction, does not have a noticeable broad absorption band, thus absorbing more violet light and transmitting more yellow and red light, causing the gemstone to appear light yellow or nearly colorless when viewed with the naked eye in this direction. In summary, the ultraviolet spectrum results from the three directions are consistent with the pleochroic characteristics of the samples (Table 4).
Through ultraviolet spectroscopic analysis, it can be observed that the broad absorption band centered at 584 nm in cordierite samples varies in intensity among different crystallographic axis directions. This variation is a key reason for the strong pleochroism exhibited by the cordierite samples. The intensity of this absorption band is closely related to the pronounced degree of purple hues in the samples; the stronger the absorption band, the more pronounced the purple hues presented in the sample. This broad absorption band results from the charge transfer between Fe2+ on the octahedral sites and Fe3+ on the tetrahedral T11 sites in cordierite [19]. This is also the main reason cordierite exhibits a blue-purple appearance.
The absorption near 494 nm in the blue-green region is related to the spin-forbidden transition of Fe2+. In the 300–450 nm range, three weak absorption bands are present in all three blue-purple samples. Among them, the absorption band at 441 nm is caused by the d-orbital electronic transition of Fe3+ from the ground state 6A1g level to the excited states 4A1g and 4Eg. The absorption at 390 nm corresponds to the spin-forbidden d–d-orbital transition of Fe3+ (6A1g4T2g) [20].
In summary, the blue-purple body color and strong pleochroism of cordierite are primarily caused by the charge transfer between Fe2+ on the octahedral sites and Fe3+ on the tetrahedral sites. This charge transfer leads to d-orbital overlap, resulting in the variation of intensity in the broad absorption band at 584 nm across different directions, which in turn causes the strong pleochroism of cordierite [19]. Moreover, the blue-purple body color is more popular in the market compared to colorless or light yellow hues. Therefore, when cutting cordierite, the table should be made as parallel as possible to the (010) plane to maximize its rich blue-purple color. However, since cordierite has a medium cleavage along the (010) set, the table needs to be cut at a 5° angle to the (010) plane to prevent cleavage planes from flaking off.

3.4. X-ray Photoelectron Spectroscopy

The results of electron probe and ultraviolet-visible spectroscopy analyses indicate that the main coloring element in purple cordierite is (Fe), but the valence distribution of Fe in cordierite and the content of different valence states are unknown. This paper uses X-ray photoelectron spectroscopy to characterize the valence states and content of Fe. A wide spectrum scan was performed on the sample to determine the types of elements present in the sample, followed by a narrow spectrum detailed scan to determine the chemical states of transition metal elements in the sample.
The wide spectrum of the sample’s X-ray photoelectron spectroscopy is shown in Figure 6. The spectrum indicates that the cordierite sample contains elements such as O, Al, Si, Mg, Fe, and Be. It is noteworthy that the characteristic peak of the coloring element Fe has a relatively weak intensity in the wide spectrum scan, suggesting that the Fe content in the sample is low.
Under the principle of XPS, the Fe2p layer electrons in cordierite undergo energy level splitting due to spin-orbit coupling when excited by X-rays, resulting in the appearance of double peaks as 2p3/2 and 2p1/2 in the XPS spectrum (Figure 7). The 2p3/2 refers to the binding energy of Fe2+ at 709.2–709.6 eV, while the 2p1/2 refers to Fe3+, which has a higher binding energy of 711.4–711.9 eV.
The coloration of cordierite is primarily due to the charge transfer between Fe2+ and Fe3+. The characteristic energy spectrum line analysis results of the Fe ion are shown in Figure 5. The peak shape of the Fe2p energy spectrum line exhibits obvious asymmetry, with a significant shoulder peak appearing on the high binding energy side of Fe2p3/2, indicating that the Fe2p3/2 energy spectrum peak is a composite. Through peak fitting, it was found that this peak is a combination of peaks at 709.4 and 711.5 eV, corresponding to FeO and Fe2O3 compounds, respectively. This indicates that two valence states of Fe exist in the cordierite sample, with both Fe2+ and Fe3+ being coordinated with O to form bonds. Among them, the peak area ratio of Fe3+ to Fe2+ for sample O1 is 4.26:1, for sample O2, it is 2.02:1, and for sample O3, it is 3.05:1. Based on the calculated areas of the fitted peaks, the contents of Fe2+ and Fe3+ are approximately 25.55% and 74.45%, respectively.

3.5. Infrared Spectrum Analysis

The infrared spectrum of cordierite exhibits a distinctive strong band near 1193 cm−1, accompanied by a weak reflection band at 1145 cm−1. The reflection band near 1193 cm−1 is related to the stretching mode of SiO4, while the satellite band at 1140 cm−1 is caused by the asymmetric stretching vibration of Si-O-Si in the structure. In the parallel b-axis direction, this strong reflection band differs from the other two directions, with a higher wavenumber (located at 1199 cm−1) and a higher reflectance.
There is a strong reflection band at 961 cm−1 accompanied by multiple weak reflection bands. Among them, the strong band at 904 cm−1 and weak reflection bands such as 961 cm−1 and 1049 cm−1 are attributed to the stretching vibration of the Al-O tetrahedron [21]. The band intensity of the satellite band at 904 cm−1 is significantly weaker in the parallel b-axis direction compared to the other two directions.
The reflection band near 770 cm−1 is a characteristic band resulting from the vibration of the cordierite’s six-membered ring structure, and the band intensity in the parallel b-axis direction is significantly higher than in the other two crystal axes. Distinct reflection bands can be observed near 671 and 617 cm−1 in the infrared spectrum, both attributed to the vibration of the octahedron. The former is weak to an almost invisible parallel to the a-direction, while the latter is an invisible parallel to the b-direction. The reflection band at 584 cm−1 is attributed to the vibration of the octahedron.
In the infrared spectrum, several reflection bands between 400 and 500 cm−1 exhibit a finger-like band shape, specifically at 484, 446, and 421 cm−1. Notably, the reflection band near 482 cm−1 is associated with the bending vibration of the Si-O tetrahedron. Additionally, the band intensities of multiple reflection bands below 500 cm−1 are significantly stronger in the parallel b-axis direction compared to the other two directions (Figure 8).
The sample particles were selected and tested using powder transmission methods, revealing distinct water-related spectral bands in the range of 3300–3900 cm⁻1. Comparing the results with the studies conducted by Goldman et al. [22], the spectral band at 3687 cm−1 is attributed to the type I stretching vibration mode of H2O. The spectral bands observed at 3627 cm⁻1 in sample O1 and 3579 cm⁻1 in samples O2 and O3 are assigned to the type II stretching vibration of H2O [23,24]. Through infrared spectroscopic analysis, it is confirmed that both Type I H2O and Type II H2O coexist in the cordierite samples [25].

3.6. Raman Spectrum Analysis

The Raman spectral results of cordierite in three crystallographic directions show that there is a significant difference in Raman intensity among different crystallization directions within the same sample.
The Raman spectra (Figure 9) of the samples largely correspond to the characteristic Raman spectra of standard cordierite. The Raman spectra of the three samples are similar in different crystallization directions but are distinct in the parallel b-axis direction compared to the parallel a-axis and c-axis directions. Bands at 1011, 973, 670, 556, 577, 260, and 238 cm−1 are present in all three directions of the three samples, yet the spectral band intensities obtained from different crystallization directions vary.
As shown in Figure 10. In the parallel b-axis direction, the Raman bands at 1181 and 368 cm−1 are sharp, while in the other two directions, they are weak to the point of being almost invisible. Both Raman bands at 973 and 1011 cm−1 are present in all three directions of the sample, but the band at 1011 cm−1 is very weak in the parallel b-axis direction, whereas in the parallel a-axis and c directions, the band at 1011 cm−1 is high and significantly stronger than the band at 973 cm−1. The Raman bands at 260 cm−1 and its companion band at 238 cm−1 are sharper in the parallel b-axis, with the companion band at 238 cm−1 being very weak in the other directions. The band at 1381 cm−1 position is noticeably sharp in the parallel b-axis, while it appears somewhat duller in the other directions. The Raman band at 327 cm−1 is duller in the parallel b-axis direction and sharper in the parallel a-axis and c directions.
This study found that within the range of 1100–200 cm−1, the diffraction intensity parallel to crystallographic axes a and c is stronger than the parallel b-axis direction. This indicates that the Si-O symmetric stretching vibration diffraction is weakest along the parallel b-axis direction. The density of silicon-oxygen groups per unit area can explain this phenomenon. It is inferred that cordierite is most difficult to compress along the b-axis direction due to anisotropic compression, and thus, compared with other crystallographic directions, there are more SiO4 tetrahedra or Si5O18 groups per unit length in the parallel b-axis direction. In other words, on the parallel b-axis plane, the number of SiO4 or Si5O18 per unit area is the lowest, which is why the Si-O vibration diffraction obtained along the parallel b-axis direction is the weakest.
The band at 1383 cm−1 may be attributed to the symmetric stretching vibration of CO2 molecules, which often align parallel to the b-axis within the channel cavities. Hence, the Raman band parallel to the b-axis is stronger. This may relate to the relatively longer cavity length parallel to the b-axis. Both 1181 and 973 cm−1 are attributed to the stretching vibrations of SiO4 [23].
The band at 1011 cm−1 is caused by the stretching of the T21 and T23 tetrahedra in the hexagonal ring and the T16 tetrahedron connected to the M site, while 973 cm−1 may also be related to the stretching of the T21 and T16 sites [24]. The band position at 670 cm−1 is caused by the Al-O tetrahedral stretching vibration at the T11 site. The strong Raman bands between 530 and 600 cm−1 split into 556 and 577 cm−1, related to the bending vibrations of the T26, T21, and T23 sites, as well as the stretching of the M octahedron and the T26 site. The Raman band at 482 cm−1 may be the result of a combination of bending vibrations and other modes, sharp in the parallel c-axis direction and duller in the parallel b-axis and directions. The weak band at 459 cm−1 is caused by the bending modes of the T21 and T23 sites. The 430 cm−1 is attributed to the bending vibration of the T26, and 368 cm−1 is due to the bending vibration of the M octahedron, sharp in the parallel b-axis direction, and weakening to almost disappearance in the parallel a-axis and c-axis directions [26]. The band positions at 295, 260, and 238 cm−1 are attributed to the bending vibrations of the Si-O tetrahedra, Al-O tetrahedra, and M-O octahedra.
In the hexagonal ring structure of cordierite, the interaction between the alkali metal ions located in the [000] direction and H2O molecules determines the type of H2O. H2O that does not interact with alkali metal ions is classified as Type I H2O, while H2O that interacts with alkali metal ions belongs to Type II H2O [24,27].
It is noteworthy that the content of alkali metal ions in cordierite is related to the spectral band intensity caused by Type II H2O. As the Na content in cordierite increases, the spectral band intensity of Type II H2O in both infrared and Raman spectra also correspondingly strengthens [22,28]. The infrared and Raman spectra of the sample correspond to the Na content of the sample; the Na content of O1 is lower than that of O2 and O3, and its spectral band intensity related to Type II H2O is also lower. Raman analysis confirms that the sample contains a certain amount of Type I H2O and a small amount of Type II H2O, which is consistent with the results of infrared spectroscopy analysis (Figure 11).
To determine the species of the cordierite samples, the sample’s Raman bands were compared with the Raman spectrum of Mg-cordierite, calculated by Kaindl et al. (2011) [23] to be closest to the pure endmember composition.
Previous studies have shown that with the increase of Fe, most of the Raman modes of cordierite will shift to lower wavenumbers [29,30]. The Raman band positions of the cordierite samples in this study are slightly offset to lower wavenumbers compared to pure Mg-cordierite (Table 5). Combined with the results of the electron probe and XPS analysis, it is inferred that the samples in this study are Mg-cordierite.

4. Conclusions

This article presents the gemological characteristics, spectroscopic characteristics, and chemical composition of cordierite through fundamental gemological testing, ultraviolet spectroscopy, infrared spectroscopy, electron probe analysis, and Raman spectroscopy. The sample’s refractive index and relative density are lower compared to iron cordierite, with an XMg range between 0.79 and 0.96. The Raman spectral characteristics are closer to those of pure Mg-cordierite, indicating that the samples are Mg-cordierite. The cordierite samples contain both Fe2+ and Fe3+ valence states, with an average ratio of 25.55% to 74.45%. The ultraviolet spectral characteristics of cordierite indicate that the difference in the broad absorption band at 584 nm is the key to cordierite’s strong pleochroism (deep purple/light purple/light yellow), caused by charge transfer between Fe2+ in the octahedron and Fe3+ in the tetrahedron. Infrared and Raman spectroscopy results reveal that differences in the Si-O group density in different directions cause variations in the spectral characteristics among the three optical principal axis directions of cordierite, with the greatest difference along the parallel b-axis, while the spectral characteristics of the other two directions are similar. Both infrared and Raman spectroscopy also prove the presence of a certain amount of Type I H2O and a small amount of Type II H2O in Mg-cordierite, with the band intensity of Type II H2O being directly proportional to the sample’s Na content. This preliminary study focuses on the basic gemological characteristics of cordierite, explores spectral differences along various crystallographic directions, and undertakes a preliminary investigation into the coloration mechanism and causes of pleochroism in cordierite. Such insights not only enhance our understanding of cordierite but also offer a new approach to its identification and classification.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14070637/s1, Table S1: EMPA data.

Author Contributions

Performing the experiment, W.Y.; analysis, W.Y.; writing the original manuscript, W.Y.; review and editing, Q.G.; translation, Z.Z. and Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Science and Technology Infrastructure, The National Infrastructure of Mineral, Rock and Fossil Resources for Science and Technology (http//www.nimrf.net.cn, accessed on 25 December 2021), as well as the Program of the Data Integration and Standardization in the Geological Science and Technology from MOST, China, grant number 2013FY110900-3.

Data Availability Statement

The original contributions presented in the study are included in the Supplementary Material, further inquiries can be directed to the corresponding authors.

Acknowledgments

Thanks to the School of Gemology, China University of Geoscience, Beijing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Belley, P.M. Iolite from the Thor-Odin Dome, British Columbia, Canada: Geology, Chemical Composition, Inclusions, and Cause of Chatoyancy. Gems Gemol. 2023, 59, 340–355. [Google Scholar] [CrossRef]
  2. Das, S.; Mohanty, J. Characteristics of Cordierite (Iolite) of Bandarguha-Orabahala Area, Kalahandi District, Odisha, India. J. Geol. Geophys. 2017, 6, 4. [Google Scholar]
  3. Cheng, R.P.; Guo, Y. Study on the effect of heat treatment on amethyst color and the cause of coloration. Sci. Rep. 2020, 10, 14927. [Google Scholar] [CrossRef]
  4. Zhang, Y.Y.; Chen, M.H. Identification of Heat-Treated Sapphires from Sri Lanka: Evidence from Three-Dimensional Fluorescence Spectroscopy. Crystals 2022, 12, 293. [Google Scholar] [CrossRef]
  5. Pluthametwisute, T.; Wanthanachaisaeng, B.; Saiyasombat, C.; Sutthirat, C.; Bolshov, M. Cause of Color Modification in Tanzanite after Heat Treatment. Molecules 2020, 25, 3743. [Google Scholar] [CrossRef] [PubMed]
  6. Parkin, K.M.; Loeffler, B.M.; Burns, R.G. Mössbauer spectra of Kyanite, aquamarine, and cordierite showing intervalence charge transfer. Phys. Chem. Miner. 1977, 1, 301–311. [Google Scholar] [CrossRef]
  7. Chen, W.; Garofalo, A.C.; Geng, H.; Liu, Y.; Wang, D.W.; Li, Q. Effect of high temperature heating on the microstructure and performance of cesium-based geopolymer reinforced by cordierite. Cem. Concr. Compos. 2022, 129, 104474. [Google Scholar] [CrossRef]
  8. Olena, K.; Aleksandr, Z.; Aleksandra, A. Low-temperature cordierite ceramics with porous structure for thermal shock resistance products. Open Ceram. 2024, 17, 100520. [Google Scholar]
  9. Zwaan, P.C. Enstatite, Cordierite, Kornerupine, and Scapolite with Unusual Properties from Embilipitiya, Sri Lanka. Gems Gemol. 1996, 32, 262–269. [Google Scholar] [CrossRef]
  10. Liu, X.; Guo, Y. Feasibility Study on Color Grading of Blue Iolite Based on GemDialogue Color Comparison Charts. Appl. Sci. 2023, 13, 6475. [Google Scholar] [CrossRef]
  11. Geiger, C.A.; Rager, H.; Czank, M. Cordierite III: The site occupation and concentration of Fe3+. Contrib. Mineral. Petrol. 2000, 140, 344–352. [Google Scholar] [CrossRef]
  12. Dudka, A.P.; Belyanchikov, M.A.; Thomas, V.G.; Bedran, Z.V.; Gorshunov, B.P. Localization of Small Impurities of Water and Carbon Dioxide in Channels of the Structure of Natural Cordierite. J. Surf. Investig. X-ray Synchrotron Neutron Tech. 2020, 14, 718–721. [Google Scholar] [CrossRef]
  13. Pollak, H. Charge Transfer in Cordierite. Phys. Status Solidi 1976, 74, K31–K34. [Google Scholar] [CrossRef]
  14. Duncun, J.F.; Johnson, J.H. Single-crystal 57Fe Mössbauer studies of the site positions in cordierite. Aust. J. Chem. 1974, 27, 249–258. [Google Scholar] [CrossRef]
  15. Raphaël, R.; Pauline, C.; Quentin, D.; Stephan, N.S.; Tangui, L.B. Modeling the polychromism of oxide minerals: The case of alexandrite and cordierite. J. Comput. Chem. 2024, 45, 834–842. [Google Scholar]
  16. Smith, T.; Jones, E. Yellow Radiation Halos in Iolite (Cordierite). Gems Gemol. 2023, 59, 504–505. [Google Scholar]
  17. Bertoldi, C.; Proyer, A.; Schönberg, D.G.; Behrens, H.; Dachs, E. Comprehensive chemical analyses of natural cordierites: Implications for exchange mechanisms. Lithos 2004, 78, 389–409. [Google Scholar] [CrossRef]
  18. Wallace, J.H.; Wenk, H.R. Structure variation in low cordierites. Am. Mineral. 1980, 65, 96–111. [Google Scholar]
  19. Faye, G.H.; Manning, P.G.; Nickel, E.H. The polarized optical absorption spectra of tourmaline, cordierite, chloritoid and vivianite: Ferrous-ferric electronic interaction as a source of pleochroism. Am. Mineral. 1968, 53, 1174–1201. [Google Scholar]
  20. Liu, X.; Guo, Y. Study on the Color-Influencing Factors of Blue Iolite. Minerals 2022, 12, 1356. [Google Scholar] [CrossRef]
  21. Langer, K.; Schreyer, W. Infrared and powder X-ray diffraction studies on the polymorphism of cordierite, Mg2(Al4Si5O18). Am. Mineral. 1969, 54, 1442–1459. [Google Scholar]
  22. Goldman, D.S.; Rossman, G.R.; Dollase, W.A. Channel constituents in cordierite. Am. Mineral. 1977, 62, 1144–1157. [Google Scholar]
  23. Bulbak, T.A.; Shvedenkova, S.V. Solid solutions of (Mg, Fe2+)-cordierite: Synthesis, water content, and magnetic properties. Geochem. Int. 2011, 49, 391–406. [Google Scholar] [CrossRef]
  24. Kolesov, B.A.; Geiger, C.A. Cordierite II: The role of CO2 and H2O. Am. Mineral. 2000, 85, 1265–1274. [Google Scholar] [CrossRef]
  25. Ventura, G.D.; Radica, F.; Bellatreccia, F.; Cavallo, A.; Capitelli, F.; Harley, S. Quantitative analysis of HO and CO in cordierite using polarized FTIR spectroscopy. Contrib. Mineral. Petrol. 2012, 164, 881–894. [Google Scholar] [CrossRef]
  26. Kaindl, R.; Többens, D.M.; Haefeker, U. Quantum-mechanical calculations of the Raman spectra of Mg- and Fe-cordierite. Am. Mineral. 2011, 96, 1568–1574. [Google Scholar] [CrossRef]
  27. Majumdar, A.S.; Mathew, G. Raman-infrared (IR) spectroscopy study of natural cordierites from Kalahandi, Odisha. J. Geol. Soc. India 2015, 86, 80–92. [Google Scholar] [CrossRef]
  28. Armbruster, T. Role of Na in the structure of low-cordierite; a single-crystal X-ray study. Am. Mineral. 1986, 71, 746–757. [Google Scholar]
  29. Haefeker, U.; Kaindl, R.; Tropper, P.; Krüger, H.; Kahlenberg, V.; Orlova, M. Structural investigations of the two polymorphs of synthetic Fe-cordierite and Raman spectroscopy of hexagonal Fe-cordierite. Mineral. Petrol. 2014, 108, 469–478. [Google Scholar] [CrossRef]
  30. Haefeker, U.; Kaindl, R.; Tropper, P. Semi-quantitative determination of the Fe/Mg ratio in synthetic cordierite using Raman spectroscopy. Am. Mineral. 2012, 97, 1662–1669. [Google Scholar] [CrossRef]
Figure 1. Some of the samples used in this study were under standard D65 light. All samples are taken from the tabletop.
Figure 1. Some of the samples used in this study were under standard D65 light. All samples are taken from the tabletop.
Crystals 14 00637 g001
Figure 2. Raman spectra of inclusions of three samples. (a) O1 indicates phlogopite. (b) Apatite in O1. (c) Tourmaline in O2. (d) Actinolite in O3.
Figure 2. Raman spectra of inclusions of three samples. (a) O1 indicates phlogopite. (b) Apatite in O1. (c) Tourmaline in O2. (d) Actinolite in O3.
Crystals 14 00637 g002
Figure 3. Structure model of cordierite. Figure (ac) show the crystal structure of cordierite in the direction of three crystal axes, a, b, and c, respectively. Oxygen atoms are in red. Mg, Fe, and Mn occupy the orange octahedron. Si occupies the dark blue tetrahedron (T16, T21, and T23). Al occupies the light blue tetrahedron (T11 and T26).
Figure 3. Structure model of cordierite. Figure (ac) show the crystal structure of cordierite in the direction of three crystal axes, a, b, and c, respectively. Oxygen atoms are in red. Mg, Fe, and Mn occupy the orange octahedron. Si occupies the dark blue tetrahedron (T16, T21, and T23). Al occupies the light blue tetrahedron (T11 and T26).
Crystals 14 00637 g003
Figure 4. Observed colors of cordierite samples at different axis directions.
Figure 4. Observed colors of cordierite samples at different axis directions.
Crystals 14 00637 g004
Figure 5. (a) UV–vis spectroscopy of the sample O3; (b) the UV–vis spectroscopy of different crystal axes of the three samples.
Figure 5. (a) UV–vis spectroscopy of the sample O3; (b) the UV–vis spectroscopy of different crystal axes of the three samples.
Crystals 14 00637 g005
Figure 6. X-ray photoelectron spectra of cordierite.
Figure 6. X-ray photoelectron spectra of cordierite.
Crystals 14 00637 g006
Figure 7. (a) XPS spectra of sample O1; (b) XPS spectra of sample O2; (c) XPS spectra of sample O3.
Figure 7. (a) XPS spectra of sample O1; (b) XPS spectra of sample O2; (c) XPS spectra of sample O3.
Crystals 14 00637 g007
Figure 8. Infrared spectra of cordierite in the direction of three crystal axes.
Figure 8. Infrared spectra of cordierite in the direction of three crystal axes.
Crystals 14 00637 g008
Figure 9. IR transmission spectra of the cordierite in the energy range of the H2O stretching vibrations.
Figure 9. IR transmission spectra of the cordierite in the energy range of the H2O stretching vibrations.
Crystals 14 00637 g009
Figure 10. Raman spectra of three crystal axes of cordierite.
Figure 10. Raman spectra of three crystal axes of cordierite.
Crystals 14 00637 g010
Figure 11. Raman spectra of cordierite in the range of H2O tensile vibrational energy.
Figure 11. Raman spectra of cordierite in the range of H2O tensile vibrational energy.
Crystals 14 00637 g011
Table 1. Common gemological characteristics of blue-violet cordierite.
Table 1. Common gemological characteristics of blue-violet cordierite.
SampleLusterTransparencyFluorescenceRefractiveBirefringenceRelative Density
O1VitreousTransparentInert1.535–1.5550.0202.618
O2VitreousTransparentInert1.539–1.5550.0162.585
O3VitreousTransparentInert1.535–1.5500.0152.574
Table 2. Elemental composition of the sample (wt.%).
Table 2. Elemental composition of the sample (wt.%).
ComponentO1O2O3
SiO252.1751.3856.66
MnO0.060.53bdl
Na2O0.170.270.51
MgO11.649.9611.63
FeO2.914.601.68
K2Obdlbdlbdl
Al2O332.0532.1127.34
Total99.0298.8497.83
Oxygen181818
Si5.1885.1695.638
Al3.7573.8073.206
Fe0.2420.3870.140
Mn0.0050.0450.001
Mg1.7261.4941.725
Na0.0330.0530.098
K0.0030.0010.001
Total10.95010.95410.808
XMg0.8770.7940.925
bdl = below detection limit.
Table 3. Ideal chemical formula of cordierite.
Table 3. Ideal chemical formula of cordierite.
SampleChemical Formula
O1chNa0.033(Mg1.726, Fe0.242, Mn0.005)1.973Al3[Al0.757Si5.188O18]
O2chNa0.053(Mg1.494, Fe0.387, Mn0.045)1.926Al3[Al0.807Si5.169O18]
O3chNa0.098(Mg1.725, Fe0.140)1.865Al3[Al0.206Si5.638O18]
Table 4. Parameters of cordierite Fe2p peak.
Table 4. Parameters of cordierite Fe2p peak.
SamplePeak PositionPeak AreaHalf-Height Width (eV)
O1 Fe2p3709.6458.091.88
Fe2p3711.41953.233.37
Fe2p1723.4200.961.88
Fe2p1725.2856.843.37
O2Fe2p3709.131517.273.34
Fe2p3711.893074.753.37
Fe2p1722.93665.593.34
Fe2p1725.481348.833.37
O3Fe2p3709.37414.433.37
Fe2p3711.401265.963.37
Fe2p1723.17181.803.37
Fe2p1724.73 555.353.37
Table 5. Raman spectra of cordierite samples and Kaindl et al. (2011) calculated the vibration mode deviation of pure Mg-cordierite and its attribution.
Table 5. Raman spectra of cordierite samples and Kaindl et al. (2011) calculated the vibration mode deviation of pure Mg-cordierite and its attribution.
Mg-CrdO1Dev.O2Dev. O3Dev.Assignment
244238−6238−6238−6b(M)
2602600258−2258−2b(T11,T26)
297295−230710294−3b(T16,T21,T23)
330327−3325−5325−5b(M)
370368−2366−4366−4b(M)
424428442624284b(T26)
55655605560554−2s(T26,M),b(T16,T26)
574577357735773s(M),b(T21,T23,T26)
672670−2670−2670−2s(T11)
964951−13951−13949−15s(T21,T16)
978973−5971−7971−7s(T21,T23)
10151011−41011−41011−4s(T16,T21,T23)
1176118151181511815s(T21,T23)
Note: Mg-Crd = calculated Raman band for pure Mg-cordierite by Kaindl et al. 2011; Dev. = deviation of the observed band from the calculated band for Mg-cordierite (Mg-Crd); s = stretching; b = bending.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yan, W.; Zhou, Z.; Rao, Y.; Guo, Q. Gemological Characteristics of Blue-Violet Cordierite. Crystals 2024, 14, 637. https://doi.org/10.3390/cryst14070637

AMA Style

Yan W, Zhou Z, Rao Y, Guo Q. Gemological Characteristics of Blue-Violet Cordierite. Crystals. 2024; 14(7):637. https://doi.org/10.3390/cryst14070637

Chicago/Turabian Style

Yan, Wenjie, Zhiyi Zhou, Yinghua Rao, and Qingfeng Guo. 2024. "Gemological Characteristics of Blue-Violet Cordierite" Crystals 14, no. 7: 637. https://doi.org/10.3390/cryst14070637

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