Next Article in Journal
Pulsed Magnetic Treatment of Cobalt for Enhanced Microstructures and Mechanical Properties
Next Article in Special Issue
Novel Method of Bauxite Treatment Using Electroreductive Bayer Process
Previous Article in Journal
Degassing of Aluminum Alloy Melts by High Shear Melt Conditioning Technology: An Overview
Previous Article in Special Issue
Effect of Al Dross Addition on Temperature Improvements in Molten Steel by Blowing Dry Air
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Transformations and Tellurium Recovery from Technical Copper Telluride by Oxidative-Distillate Roasting at 0.67 kPa

Institute of Metallurgy and Ore Beneficiation JSC, Satbayev University, Almaty 050010, Kazakhstan
*
Author to whom correspondence should be addressed.
Metals 2022, 12(10), 1774; https://doi.org/10.3390/met12101774
Submission received: 12 August 2022 / Revised: 30 September 2022 / Accepted: 19 October 2022 / Published: 21 October 2022
(This article belongs to the Special Issue Recovery of Valuable Metals from Industrial By-Products)

Abstract

:
This paper presents the results of a study of phase transformations occurring in copper-telluride by-products during its processing of oxidation-distillate roasting at low pressure. The results show that copper telluride is oxidized through intermediate compounds to the most stable tellurate (Cu3TeO6) at low temperatures. The increase in the roasting temperature above 900 °C and the presence of an oxidizer favor the copper orthotellurate decomposition. Thus, the tellurium extraction rate is 90–93% at a temperature of 1000 °C, the oxidant flow rate is 2.2 × 10−2 m3/m2·s, and the roasting time is 60–90 min. One of the decomposition products is copper oxide alloy, which is the basis of the residue. The second product is tellurium in oxide form, which evaporates and then condenses in the cold zone of the condenser in crystalline form. The main constituent phase of the condensate is tellurium oxide (TeO2), which can be further processed during one operation to elemental chalcogen by thermal reduction or electrolytic method.

1. Introduction

The main raw material source of tellurium is slime from electrolysis production of copper, a multicomponent material that contains copper from 1 to 50, silver up to 29, gold up to 2.18, selenium from 2 to 15, and tellurium up to 22 wt%. As a rule, copper in slime is represented in elemental form, and selenium and tellurium in the form of copper and noble metal chalcogenides. The variety of chemical and phase compositions of slimes determines quite a wide range of methods for their processing. Some developed methods are described in [1,2,3,4,5].
To date, both hydrometallurgical and pyrometallurgical methods have been used to recover tellurium from copper electrolyte sludge [6,7,8].
A schematic diagram of classical tellurium production during the processing of copper electrolyte slime [6,7,9] is shown in Figure 1. Full implementation of the process enables selective recovery of target commercial metals from slime: gold, silver, selenium, tellurium, as well as copper.
As can be seen, the slime undergoes leaching in the first stage. The insoluble residue containing gold and silver is sent to the noble metals’ workshop. Tellurium concentrates in a solution that is first directed to the precipitation of selenium and silver, then to the precipitation of tellurium itself. The solution obtained at this stage is returned to copper production, and the precipitate is used for the production of elemental tellurium. Vacuum distillation and/or zone melting are used to obtain high-purity tellurium (up to 6 N) at the final stage of the process flow. The combined use of these two pyrometallurgical methods enables the obtaining of tellurium with semiconductor purity. It should be noted that, in the case of low-quality extraction of selenium at the beginning of the process, an azeotropic Te-Se mixture will form during the distillation of tellurium [10,11], which will adversely affect the quality of elemental tellurium.
Frequently, the cementation of metallic copper is used for tellurium precipitation from the solution. As a result, insoluble copper telluride containing phases of both stoichiometric (Cu2Te) and nonstoichiometric compositions (Cu2−xTe) are formed. Copper telluride, in addition to tellurium with varying purity, is a valuable product of tellurium production [12,13,14,15]. Some plants prefer to accumulate or sell copper telluride at relatively low prices rather than to further separate tellurium and copper. This approach is associated with technological and technical difficulties due to the amount of stages, the significant consumption of reagents, the generation of large amounts of wastewater containing heavy metals, etc.
Therefore, the development of a cost-effective and environmentally friendly process for the recovery of Te and Cu from technical copper telluride is of great importance to the metallurgical industry.
Information in the field on the developments of telluride recovery from copper telluride is practically absent in the literature. We found only two works by Chinese scientists [15,16] that proposed to improve the preparatory stage of metallic tellurium electrolysis.
Traditionally, the middling is first dissolved by oxidative alkaline leaching with the addition of NaOH [7,9]. Then, the tellurium-containing solution is sent for electrolysis, while the copper-containing residue is for copper recovery.
The authors studied the possibility of oxidative leaching of copper telluride with the addition of NaOH at both an increased and atmospheric pressure in one of their works [15]. Oxygen was used as an oxidizing agent. The best results were achieved with autoclave leaching; more than 95% of tellurium was transferred into the solution in the form of Na2TeO3. The precipitate exclusively contained crystalline phase Cu2O that enabled it to be returned to copper production. The technology developed for tellurium deposition in the form of TeO2 involved the neutralization of a tellurium-containing solution with sulfuric acid. Then, the obtained tellurium dioxide, containing up to 97% TeO2, could be sent to the electrolytic workshop for metallic tellurium separation.
The second method [16] suggested the use of H2O2 as an oxidizer. The process also consisted of two stages: oxidative-alkaline leaching and precipitation of tellurium from the solution in the form of its oxide. The tellurium recovery rate after the alkaline leaching stage that was performed in two stages was 93%. The overall recovery of tellurium was almost 90%; the TeO2 content in the resulting residue was 96%. The authors also proposed sending the tellurium-containing residue for electrolytic production of elemental tellurium.
As can be seen from the above data, the developed methods do not help to reduce the number of traditional drawbacks that accompany hydrometallurgical processes. Pyrometallurgical methods are well-combined with the schemes intended to obtain the pure element and are the most preferred. However, they have not found application in practice or research development in this area because of the high temperatures in the process. An effective way to reduce the process temperature by 100–200 °C is to conduct the process at reduced pressure. In addition, the use of a vacuum also contributes to the improvement of working conditions for the personnel as a consequence of the fact that the process is performed in a hermetic and compact apparatus.
Obtaining tellurium by Cu2Te decomposition in actual conditions by the vacuum-thermal method is not possible due to a small dissociation pressure of liquid copper telluride: 0.7 kPa at 1780 °C. However, the saturated vapor pressure of TeO2 is about 0.02 kPa and is already at 733 °C. Therefore, this study was performed to develop a method for the recovery of tellurium in the form of an oxide with the use of the pyrometallurgical method at low pressure with its further processing to elemental tellurium without any technological and technical difficulties. In particular, phase transformations were defined, and the influence of temperature and process duration on tellurium recovery at oxidative-distillation roasting of commercial copper telluride was determined.

2. Materials and Methods

2.1. Methodology

The laboratory installation (Figure 2) consisted of a horizontal tubular electric furnace RT 50/250/13 (Nabertherm, Germany), where the reactor used for this study was placed. The reactor was a quartz tube that had channels at the ends for air supply and gas phase evacuation. A 2HVR-5DM UHL4 vacuum pump (Vacuummash, Kazan, Russia) was used to create a vacuum in the system. The pressure in the reactor was controlled by a barometer-aneroid and a McLeod monometer. The oxidizer flow rate was controlled by a PC-3A rotameter (Teplopribor, Moscow, Russia).
The experimental procedure was as follows: A copper telluride sample (weight, 3 g) was loaded into an alundum boat. Then, the boat was placed in a split (lengthwise) alundum condenser. The condenser, in turn, was placed in a quartz reactor. The ready reactor was placed in a furnace that was heated to the required temperature so that the copper telluride sample was in the isothermal zone. A vacuum system and an oxidizer feeding system were connected to the reactor. The beginning of the experiment was considered the moment when the pressure and flow rate of the oxidant reached the specified values. The systems of gas supply and evacuation were disconnected at the end of the experiment. The reactor was removed from the furnace and cooled in the air. The obtained copper alloy and tellurium condensate were weighed and analyzed.
The process temperature varied in the range from 500 to 1000 °C, and the exposure time was from 5 to 90 min.
Pressure in the system was maintained at 0.67 kPa. The mentioned pressure is necessary for transferring a tellurium oxide into a gas phase in the decomposition of copper orthotellurate (Cu3TeO6), which formed during the oxidation of copper telluride.
Air oxygen was used as an oxidizer. The airflow rate was chosen to be constant and maximal at 2.2 × 10−2 m3/m2·s, considering the capacity of the evacuating vacuum system that provided the reactor pressure of 0.67 kPa.

2.2. Characterizations

The material composition was studied by X-ray fluorescence analysis using a wave dispersive combined spectrometer, Axios PANalytical (Malvern Panalytical Ltd., Malvern, UK).
An X-ray diffractometer D8 Advance Bruker and Cu-Kα radiation and the ICDDPDF-2 reference database (2020) were used to identify the phase composition.

2.3. Materials

The tellurium-containing middling (Cu2Te) used in our study was obtained from copper electrolyte slime at the refinery of Kazakhmys Smelting LLP (Balkhash, the Republic of Kazakhstan).
Copper telluride (Figure 3), in its appearance, was a pelletized, wet-black material and odorless at the time of delivery. There were inclusions of malachite color on the surface of agglomerates. The moisture content was 29%.
A large proportion of amorphous halo (78.3%) due to the scattering from disordered phases was established under the X-ray phase-analysis results. The crystalline part of the middling amounted to 21.7% and mainly consisted of copper telluride phases of stoichiometric (Cu2Te, PDF 00-057-0477) and no-stoichiometric (Cu7Te4, PDF 02-1222 Cu5Te3, PDF 00-057-0196) compositions. In addition to copper tellurides, crystalline phases of oxidative reaction products of the copper were found in the material (Cu5(SO3)2(OH)6·5H2O—PDF 51-0321, Cu3(SO4)(OH)4—PDF 07-0408). Apparently, the mentioned hydroxosulfates are the main phase of malachite-colored phenocrysts on the surface of copper telluride [14]. The probable cause of material oxidation is insufficient washing of copper telluride from CuSO4 solution after tellurium cementation operation on copper from tellurous acid.
Differences in composition and structure were determined when comparing the appearance and phase composition of the two samples of the copper telluride [14,17], which were acquired at different times but analyzed at the same time. Apparently, oxidative processes proceed over time in the material under the influence of its humidity, the presence of sulfate ions, and storage conditions. As a result, the mentioned differences were observed.
The sample of copper telluride was further analyzed before the experiments intended to study the phase transformations occurring in the material during oxidative-distillate roasting. It was found from the analysis results that the copper telluride structure was quite strongly amorphized. Both copper hydroxosulfates and copper tellurides of variable composition were found among the crystalline phases. The humidity of the middling was 9%.
The material composition of copper telluride samples is shown in Table 1: sample 1 at the time of delivery and sample 2 before the study. More detailed information on the physicochemical properties of the product we studied can be found in [13,17].
For this study, a representative sample of copper telluride (sample 2) was crushed to a fineness of less than 0.315 mm.

3. Results and Discussion

3.1. Low-Temperature Roasting (500 and 700 °C)

It is known that the volatility of tellurium dioxide is low up to its melting temperature (733 °C) and is the only stable tellurium oxide at temperatures above 430 °C [18]. The value of TeO2-saturated vapor pressure at its melting temperature of 733 °C is 0.026 kPa according to [19], and about 0.019 kPa according to [6]. Therefore, it is obvious that the distillation process at temperatures below 900 °C is inexpedient.
However, the phase transitions occurring at low temperatures are more expressed than at high temperatures. It should be expected, based on the literature sources, that strandbergite (Cu5(SO4)2(OH)6·5H2O) and posnjakite (Cu4(SO4)(OH)6·H2O) will eventually transform into the brochantite form (Cu4(SO4)(OH)6) (at a low content of SO2/SO42−) or in the form of anthlerite (Cu3(SO4)(OH)4) (at a high content of SO2/SO42−), depending on the SO2/SO42− content in the reaction space [20]. The formation of copper oxides and elemental tellurium is the first assumed of the oxidative reactions at low temperatures. Then, it will be oxidized to TeO2, which will further react with copper oxides to form CuTe2O5 and/or CuTeO3 if enough tellurium accumulates [21]. The formed copper tellurites are also expected to be oxidized to orthotellurate (Cu3TeO6) [18].
The results of qualitative (Figure 4) and semiquantitative (Table 2) analyses of the phase composition of the cinders obtained at 500 °C were as follows.
The decomposition of Cu5(SO4)2(OH)6·5H2O and Cu4(SO4)(OH)6·H2O occurs quite intensively and is accompanied by the joint formation of anthlerite and brochantite. Apparently, anthlerite is initially formed. The concentration of SO2/SO42− decreases with the removal of sulfate ions by the oxidant flow, resulting in the formation of brochantite. Cu4(SO4)(OH)6 and anthlerite further decompose to form dolerophanite (CuOSO4), copper (I) oxide, and water. At the same time, brochantite exists for a short time, while the anthlerite decomposition process is prolonged and did not end in the studied time interval.
The formation of copper and tellurium oxides was noted during the copper telluride oxidation at the initial stages, which agrees quite well with the data [21]. However, tellurite phases of copper (CuTeO3 and CuTe2O5), elemental tellurium, and tellurium dioxide phases at 500 °C were not detected. This may have been due to the disordered structure of the material, as evidenced by the presence of an amorphous halo on the diffractograms of the samples.
For the oxide forms of tellurium, the presence of intermediate phases Te2O5 and Te4O9 were established that decompose into dioxide and free oxygen at 500–520 °C according to [22], and according to [23] at 595 °C (at atmospheric pressure). In addition, during the roasting process, the possible formation of telluric anhydride TeO3 dissociates to Te2O5 and Te4O9 when it is heated above 400 °C [19,24].
In our study, the crystalline products of the interaction between copper and tellurium oxides were intermediate phase X (d = 2.91340 Å, at 2θ = 30.66 Å) and copper orthotellurate (d = 2.748606 Å, 2θ = 32.55038 Å). The latter may have been a product of the interaction of both copper tellurites, with copper (I) oxide in the air atmosphere and the “initial” CuO and TeO3 [25]. The nature of the X phase should be studied further.
A similar mechanism of hydroxosulfates destruction and the oxidation of copper telluride was shown by the results of the study performed at 700 °C (Figure 5, Table 2).
The complete decomposition of copper hydroxosulfates within the first 15 min of roasting and the intensive formation of phase X and copper orthotellurate through intermediate compounds were established. Phases of metatellurite CuTeO4, tellurite CuTeO3, and tellurite copper Te3Cu2O7 were found in addition to these compounds in the residues. Despite the existence of congruent melting tellurates Te3Cu2O7 and TeCu2O3 in the Cu2O-TeO2 system [26], the formation of the Te3Cu2O7 compound in the air was questioned in [27]. According to the authors, the Te3Cu2O7 is most probably represented mainly by CuTe2O5 pyrotellurite molecules. In this case, the formation of tellurites and orthotellurate involves only tenorite, which is an oxidation product of cuprite. Sufficiently complete oxidation of the material to copper orthotellurate occured at roasting times of 45 min or more. The Cu3TeO6 phase content in the cinders was 93.1–95.5 wt%.

3.2. High-Temperature Roasting (900 and 1000 °C)

The copper orthotellurate Cu3TeO6 formed in the process of oxidative-distillate roasting is stable up to temperatures of 810–880 °C according to the literature data [25,28,29].
The results of X-ray phase analysis of cinders from high-temperature roasting showed (Figure 6 and Figure 7, and Table 3) that the total decomposition of copper hydroxosulfates occurred at an exposure time of 10 min. Additionally, in the case of low-temperature roasting, the formation of Cu3TeO6 was accompanied by a decrease in the amount of phase X. Copper orthotellurate did not completely decompose at 900 °C for the 90 min of exposure. At 1000 °C, 10 min of roasting was sufficient for the total decomposition of Cu3TeO6; cinders from longer roasting durations were represented by a mixture of copper oxides.
There are very scarce literature data regarding the issue of the decomposition products of copper orthotellurate. Thus, it is indicated in [30] only that it breaks down to CuTeO4 and copper oxide CuO. Additionally, in [28], the following mechanism was established in the study of the thermal behavior of copper metatellurite in the air:
Metatellurite is destroyed by the reaction at 540–630 and 700–725 °C: 3CuTeO4 → Cu3TeO6 + 2TeO2 + O2;
Then, at 840–880 °C, the interaction of copper orthotellurate and tellurium oxide occurs, resulting in the formation of copper tellurite with the emission of oxygen in the gas phase: Cu3TeO6 + 2TeO2 → 3CuTeO3 + 1/2O2.
The results of our studies showed that the decomposition of Cu3TeO6 in the air flow proceeded by the req: Cu3TeO6 → 3CuO + TeO2 + 1/2O2.
This difference from the literature data is due to the fact that the saturated vapor pressure of TeO2 reaches a value of about 1.33 kPa at a process temperature of 900 °C, according to [19], and at approximately 0.53 kPa, according to [6]. Therefore, there will be an accumulation of formed oxide in the solid product at atmospheric pressure [28]. In our study (where the system pressure was 0.67 kPa), the tellurium oxide was transferred into the gas phase and removed from the reaction zone by an oxidant flow. Such conditions contributed to the decomposition of orthotellurate rather than its transition into the copper tellurite form.

3.3. Influence of Roasting Temperature and Time on Tellurium Recovery at 0.67 kPa

The dependence of the tellurium extraction rate and speed on the roasting time at different process temperatures is shown in Figure 8. The extraction rate (%) was taken as the ratio of the amount of evaporated tellurium to its initial mass, and the amount of tellurium evaporated per unit time from per unit geometric area of evaporation (boats) was taken as the extraction speed (g/cm2∙s).
It can be seen from the given data that, at 500 and 700 °C, the degree of tellurium extraction was insignificant and did not exceed 20% during an hour and a half exposure. An increase in the oxidative-distillate temperature of up to 900–1000 °C had a positive effect on tellurium recovery that reached about 70% at 900 °C and about 90% at 1000 °C during an hour and a half of roasting. A sharp rise in the step curve during high-temperature roasting was observed within the first 15 min; then, its growth slowed down. At the same time, the speed curve tended to grow within the first 5 min. An increase in the process duration over 60 min at 700–1000 °C had little effect on the tellurium recovery from commercial copper telluride. It should be noted that an acceptable value (93–98%) of the recovery rate is reached at a process temperature of 1100 °C [31].
The obtained condensate is a dense crystalline powder of white color (Figure 9), well-separated from the condenser surface. The presence of crystalline phases of tellurium dioxide TeO2 (PDF 00-042-1365) in an amount of 67.7% and tellurium oxysulfate Te2O3(SO4) (PDF 01-070-0135) at 32.3% was established in the condensate by diffractometric analysis (Figure 10). The presence of tellurium oxysulfate can be explained by a violation of the technological process for washing technical copper telluride from sulfuric acid solution in the hydrometallurgical production scheme.
The condensate is the initial middling intended to obtain elemental tellurium by known technologies.

4. Conclusions

We found as a result of this work that copper telluride is oxidized to the most stable tellurate Cu3TeO6 through intermediate compounds in the process of oxidation roasting at low temperatures. Thus, one of the intermediate compounds is the unidentified phase, X. The increase in the roasting temperature above 900 °C at a pressure of 0.67 kPa and in an oxidant flow process contributed to the copper orthotellurate decomposition. Thus, the value of the tellurium extraction rate is 90–93% at 1000 °C (pressure—0.67 kPa, the oxidant flow speed is 2.2 × 10−2 m3/m2·s, and roasting time is 60–90 min). One of the decomposition products is copper oxide alloy, which is the basis of the residue. The second product is the oxide forms of tellurium that evaporate and then condense in crystalline form in the cold zone of the condenser.
The condensate is a dense, fine crystalline powder of white color that separates well from the surface of the condenser. The process of condensate collapsing from the condenser walls in a loose form can be arranged with an appropriate technical design.
The main constituent phase of the condensate is tellurium oxide, which can be processed into elemental chalcogen in one operation by thermal reduction or electrolytic method.

Author Contributions

Conceptualization, A.N.; methodology, A.N. and V.V.; investigation, X.L., F.T., N.B. and G.R.; data curation, S.T. and V.V.; writing—original draft preparation, A.N. and V.V.; writing—review and editing, A.N. and S.T.; visualization, A.N. and X.L.; project administration, A.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education and Science of the Republic of Kazakhstan, grant number AP08052016.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the funding support from the Ministry of Education and Science of the Republic of Kazakhstan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mastyugin, S.A.; Naboichenko, S.S. Processing of copper-electrolyte slimes: Evolution of technology. Russ. J. Non-Ferr. Met. 2012, 53, 367–374. [Google Scholar] [CrossRef]
  2. Liu, G.; Wu, Y.; Tang, A.; Pan, D.; Li, B. Recovery of scattered and precious metals from copper anode slime by hydrometallurgy: A review. Hydrometallurgy 2020, 197, 105460. [Google Scholar] [CrossRef]
  3. Xing, W.D.; Lee, M.S. Leaching of gold and silver from anode slime with a mixture of hydrochloric acid and oxidizing agents. Geosyst. Eng. 2017, 20, 216–223. [Google Scholar] [CrossRef]
  4. Xiao, L.; Wang, Y.L.; Yu, Y.; Fu, G.Y.; Han, P.W.; Sun, Z.H.I.; Ye, S.F. An environmentally friendly process to selectively recover silver from copper anode slime. J. Clean. Prod. 2018, 187, 708–716. [Google Scholar] [CrossRef]
  5. Ding, Y.; Zhang, S.; Liu, B.; Li, B. Integrated process for recycling copper anode slime from electronic waste smelting. J. Clean. Prod. 2017, 165, 48–56. [Google Scholar] [CrossRef]
  6. Chizhikov, D.M.; Shchastlivyi, V.P. Tellurium and Tellurides; Collet’s Publishers Ltd.: London, UK, 1970; 279p. [Google Scholar]
  7. Hoffman, J.E. Recovering selenium and tellurium from copper refinery slimes. JOM 1989, 41, 33–38. [Google Scholar] [CrossRef]
  8. Mahmoudi, A.; Shakibania, S.; Mokmeli, M.; Rashchi, F. Tellurium, from copper anode slime to high purity product: A review paper. Met. Mater. Trans. B 2020, 51, 2555–2575. [Google Scholar] [CrossRef]
  9. Mastyugin, S.A.; Volkova, N.A.; Naboichenko, S.S.; Lastochkina, M.A. Slime from Electrolytic Refining of Copper and Nickel; Ural Federal University: Ekaterinburg, Russia, 2013; 256p. [Google Scholar]
  10. Volodin, V.N.; Trebukhov, S.A.; Burabaeva, N.M.; Nitsenko, A.V. Melt-gas phase equilibria and state diagrams of the selenium-tellurium system. Russ. J. Phys. Chem. 2017, 91, 800–804. [Google Scholar] [CrossRef]
  11. Volodin, V.N.; Trebukhov, S.A.; Kenzhaliyev, B.K.; Nitsenko, A.V.; Burabaeva, N.M. Melt-vapor phase diagram of the Te-S system. Russ. J. Phys. Chem. 2018, 92, 407–410. [Google Scholar] [CrossRef]
  12. Wang, S. Tellurium, its resourcefulness and recovery. JOM 2011, 63, 90–93. [Google Scholar] [CrossRef]
  13. Shibasaki, T.; Abe, K.; Takeuchi, H. Recovery of tellurium from decopperizing leach solution of copper refinery slimes by a fixed bed reactor. Hydrometallurgy 1992, 29, 399–412. [Google Scholar] [CrossRef]
  14. Nitsenko, A.V.; Burabaeva, N.M.; Tuleytay, F.K.; Seisembaev, R.S.; Linnik, X.A.; Azlan, M.N. Study of physical and chemical properties of tellurium-containing middlings. Kompleks. Ispolz. Miner. Syra 2020, 315, 49–56. [Google Scholar] [CrossRef]
  15. Xu, L.; Xiong, Y.; Song, Y.; Zhang, G.; Zhang, F.; Yang, Y.; Hua, Z.; Tian, Y.; You, J.; Zhao, Z. Recycling of copper telluride from copper anode slime processing: Toward efficient recovery of tellurium and copper. Hydrometallurgy 2020, 196, 105436. [Google Scholar] [CrossRef]
  16. Xu, L.; Xiong, Y.; Zhang, G.; Zhang, F.; Yang, Y.; Hua, Z.; Tian, Y.; You, J.; Zhao, Z. An environmental-friendly process for recovery of tellurium and copper from copper telluride. J. Clean. Prod. 2020, 272, 122723. [Google Scholar] [CrossRef]
  17. Nitsenko, A.V.; Linnik, X.A.; Tuleytay, F.K.; Burabaeva, N.M.; Seisembaev, R.S. Physico-chemical characteristics of tellurium-containing industrial middling Kazakhmys Smelting LLP. Theory Process Eng. Metall. Prod. 2021, 38, 10–16. [Google Scholar]
  18. Dimitriev, Y.; Gatev, E.; Ivanova, Y. High-temperature X-ray study of the oxidation of CuTeO3. J. Mater. Sci. Lett. 1989, 8, 230–231. [Google Scholar] [CrossRef]
  19. Kindyakov, P.S.; Korshunov, B.G.; Fedorov, P.I.; Kislyakov, I.P. Chemistry and Technology of Rare and Trace Elements, 2nd ed.; Bolshakov, K.A., Ed.; 3rd Part; Vysshaja Shkola: Moscow, Russia, 1976; 320p. [Google Scholar]
  20. Krätschmer, A.; Odnevall Wallinder, I.; Leygraf, C. The evolution of outdoor copper patina. Corros. Sci. 2002, 44, 425–450. [Google Scholar] [CrossRef]
  21. Kukleva, T.V.; Fedorova, T.B.; Vishnyakov, A.V.; Kovtunenko, P.V. Features of low-temperature oxidation of copper (I) telluride. Inorg. Mater. 1988, 24, 1469–1471. [Google Scholar]
  22. Itkin, V.P.; Alcock, C.B. The O-Te (Oxygen-Tellurium) system. J. Phase Equilibria 1996, 17, 533–538. [Google Scholar] [CrossRef]
  23. Pashinkin, A.S.; Dolgikh, V.A. Some questions about chemistry and thermodynamics of oxygen compounds of tellurium in connection with the oxidation of tellurides. Russ. J. Inorg. Chem. 1997, 42, 190–195. [Google Scholar]
  24. Ahmed, M.A.K.; Fjellvåg, H.; Kjekshus, A. Synthesis, structure and thermal stability of tellurium oxides and oxide sulfate formed from reactions in refluxing sulfuric acid. J. Chem. Soc. Dalton Trans. 2000, 24, 4542–4549. [Google Scholar] [CrossRef]
  25. Zhu, X.; Wang, Z.; Su, X.; Vilarinho, P.M. New Cu3TeO6 ceramics: Phase formation and dielectric properties. ACS Appl. Mater. Interfaces 2014, 6, 11326–11332. [Google Scholar] [CrossRef]
  26. Kozhukharov, V.; Marinov, M.; Pavlova, J. On the phase equilibrium in the TeO2–Cu2O system. Z. Anorg. Allg. Chem. 1980, 460, 221–227. [Google Scholar] [CrossRef]
  27. Kozhukharov, V.; Marinov, M.; Pavlova, J. On the phase equilibrium in the TeO2–Cu2O (CuO) system. Mater. Chem. Phys 1984, 10, 401–412. [Google Scholar] [CrossRef]
  28. Gospodinov, G.G. Phase states of copper orthotellurates in an aqueous medium and in thermolysis. J. Mater. Sci. Lett. 1992, 1, 1460–1462. [Google Scholar] [CrossRef]
  29. Gospodinov, G.G. Synthesis, crystallographic data and thermostability of some metal ortho-tellurates of the type Me3TeO6 and Me2TeO6. Thermochim. Acta 1985, 83, 243–252. [Google Scholar] [CrossRef]
  30. The Materials Project. Available online: https://materialsproject.org/materials/mp-21861/ (accessed on 10 August 2022).
  31. Nitsenko, A.V.; Volodin, V.N.; Linnik, X.A.; Tuleutay, F.K.; Burabaeva, N.M. Distillation recovery of tellurium from copper telluride in oxide forms. Russ. J. Non-Ferr. Met. 2022, 63, 284–291. [Google Scholar] [CrossRef]
Figure 1. Copper electrolyte slime processing diagram.
Figure 1. Copper electrolyte slime processing diagram.
Metals 12 01774 g001
Figure 2. Horizontal vacuum unit with a split condenser: (1) rotameter, (2) temperature controller in the reaction zone, (3) control thermocouple, (4) electric furnace, (5) boat, (6) isothermal zone, (7) sample, (8) reactor, (9) split condenser, (10) condensation zone, (11) condensate, (12) filter, (13) barometer-aneroid, (14) vacuum pump, and (15) furnace controller.
Figure 2. Horizontal vacuum unit with a split condenser: (1) rotameter, (2) temperature controller in the reaction zone, (3) control thermocouple, (4) electric furnace, (5) boat, (6) isothermal zone, (7) sample, (8) reactor, (9) split condenser, (10) condensation zone, (11) condensate, (12) filter, (13) barometer-aneroid, (14) vacuum pump, and (15) furnace controller.
Metals 12 01774 g002
Figure 3. X-ray diffraction patterns of a tellurium-containing middling of Kazakhmys Corporation LLC: (1) at the time of delivery and (2) before studies.
Figure 3. X-ray diffraction patterns of a tellurium-containing middling of Kazakhmys Corporation LLC: (1) at the time of delivery and (2) before studies.
Metals 12 01774 g003
Figure 4. X-ray patterns of residues were obtained at 500 °C, 0.67 kPa, and different roasting durations.
Figure 4. X-ray patterns of residues were obtained at 500 °C, 0.67 kPa, and different roasting durations.
Metals 12 01774 g004
Figure 5. X-ray patterns of residues were obtained at 700 °C, 0.67 kPa, and different roasting durations.
Figure 5. X-ray patterns of residues were obtained at 700 °C, 0.67 kPa, and different roasting durations.
Metals 12 01774 g005
Figure 6. X-ray patterns of residues were obtained at 900 °C, 0.67 kPa, and different roasting durations.
Figure 6. X-ray patterns of residues were obtained at 900 °C, 0.67 kPa, and different roasting durations.
Metals 12 01774 g006
Figure 7. X-ray patterns of residues were obtained at 1000 °C, 0.67 kPa, and different roasting durations.
Figure 7. X-ray patterns of residues were obtained at 1000 °C, 0.67 kPa, and different roasting durations.
Metals 12 01774 g007
Figure 8. Dependence of the rate (a) and speed (b) of tellurium extraction on the duration of exposure at different temperatures.
Figure 8. Dependence of the rate (a) and speed (b) of tellurium extraction on the duration of exposure at different temperatures.
Metals 12 01774 g008
Figure 9. Condensate from the copper telluride roasting on the surface of the condenser (a) and in the free state (b).
Figure 9. Condensate from the copper telluride roasting on the surface of the condenser (a) and in the free state (b).
Metals 12 01774 g009
Figure 10. X-ray pattern of obtained condensate.
Figure 10. X-ray pattern of obtained condensate.
Metals 12 01774 g010
Table 1. Composition of tellurium-containing middling.
Table 1. Composition of tellurium-containing middling.
SampleElements, wt%
OAlSiSClFeCuAsSeTePb
118.880.110.032.090.290.0247.190.110.0431.220.02
229.460.010.022.010.230.0241.500.100.0226.63
Table 2. Phase composition (wt%) of low-temperature roasting residues (for Figure 4 and Figure 5 (hereinafter, the quantitative analysis is given without considering the unknown phase, X.)).
Table 2. Phase composition (wt%) of low-temperature roasting residues (for Figure 4 and Figure 5 (hereinafter, the quantitative analysis is given without considering the unknown phase, X.)).
PhasePhases Content (%) at the Roasting Time, min
5101530456090
at 500 °C
Cu3(SO4)(OH)427.024.69.06.65.35.25.1
Cu4(SO4)(OH)627.013.0
Cu4(SO4)(OH)6·H2O10.9
CuOSO48.47.810.512.013.5
Cu2-xTe18.513.98.05.14.66.08.1
Cu3TeO610.015.420.627.031.5
CuO19.443.041.734.731.322.4
Cu2O12.220.05.1
Te2O54.49.18.05.25.3
Te4O914.018.218.918.014.2
at 700 °C
Cu3(SO4)(OH)45.5
CuOSO48.35.0
Cu3TeO643.847.171.380.094.195.594.5
CuTeO36.87.15.2
CuTeO42.913.29.68.5
Te3Cu2O79.16.0
CuO11.512.87.911.55.94.55.5
Te4O912.114.8
Table 3. Phase composition (wt%) of high-temperature roasting residues (for Figure 6 and Figure 7).
Table 3. Phase composition (wt%) of high-temperature roasting residues (for Figure 6 and Figure 7).
SamplePhases Content (%) at the Roasting Time, min
5101530456090
at 900 °C
Cu3(SO4)(OH)44.3
Cu3TeO627.259.15.147.777.255.430.3
CuTeO36.5
CuTeO419.3
Te3Cu2O711.5
CuO23.632.322.949.722.832.855.3
Cu2O4.18.614.32.611.814.3
Te4O93.54.7
at 1000 °C
Cu3TeO656.6
CuTeO32.9
Te3Cu2O74.3
CuO14.481.776.227.635.487.115.5
Cu2O6.312.020.360.956.112.982.7
Te4O94.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nitsenko, A.; Linnik, X.; Volodin, V.; Tuleutay, F.; Burabaeva, N.; Trebukhov, S.; Ruzakhunova, G. Phase Transformations and Tellurium Recovery from Technical Copper Telluride by Oxidative-Distillate Roasting at 0.67 kPa. Metals 2022, 12, 1774. https://doi.org/10.3390/met12101774

AMA Style

Nitsenko A, Linnik X, Volodin V, Tuleutay F, Burabaeva N, Trebukhov S, Ruzakhunova G. Phase Transformations and Tellurium Recovery from Technical Copper Telluride by Oxidative-Distillate Roasting at 0.67 kPa. Metals. 2022; 12(10):1774. https://doi.org/10.3390/met12101774

Chicago/Turabian Style

Nitsenko, Alina, Xeniya Linnik, Valeriy Volodin, Farkhat Tuleutay, Nurila Burabaeva, Sergey Trebukhov, and Galiya Ruzakhunova. 2022. "Phase Transformations and Tellurium Recovery from Technical Copper Telluride by Oxidative-Distillate Roasting at 0.67 kPa" Metals 12, no. 10: 1774. https://doi.org/10.3390/met12101774

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