Yuanjun Liua,
Guoxing Zhu*bc,
Jing Yangb,
Chunlin Baob,
Jing Wanga and
Aihua Yuan*a
aSchool of Environmental and Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212018, China. E-mail: aihuayuan@just.edu.cn
bSchool of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang, 212013, China. E-mail: zhuguoxing@ujs.edu.cn
cState Key Laboratory of Coordination Chemistry, Nanjing University, Nanjing, 210093, China
First published on 26th November 2015
Controlling the composition and crystal phase is an important issue to tune material physical/chemical properties. Herein, it was found that triphenyl phosphine (TPP) can be used as a phase transfer agent to transform CuS, Cu39S28 phases into pure low-sulfur Cu1.8S phase. When mixed phase copper sulfides were reacted with triphenyl phosphine under suitable temperature, sulfur was extracted to produce the low-sulfur Cu1.8S phase. It was also demonstrated that the Cu–S product can effectively catalyze a clock reaction between methylene blue and hydrazine in aqueous medium. In addition, the photothermal conversion properties of the Cu–S based products were studied. The results show that the purified Cu1.8S materials show enhanced or similar properties than the original mixed-phase Cu–S products.
The electrical conductivity breadth from metallic to semiconducting to superconducting and the defect chemistry owing to nonstoichiometry for copper sulfides render them attractive for various applications.8–22 Copper sulfides with various Cu/S ratios have demonstrated great potential for wide applications in fields such as solar cells, photoelectric transformers, nano-switches, thermoelectric materials, electrocatalysts and photocatalysts.23,24 The Cu–S system has at least nine kinds of different crystal phase with the varied x value in Cu2−xS (x = 0–1) including Cu2S (γ- and β-chalcocite),25 Cu1.98S (djurleite),26 Cu1.8S (digenite),27,28 Cu1.75S (anilite),29,30 and CuS (covellite),31,32 as shown in the X-ray diffraction pattern database. The energy band structure and physical/chemical properties of Cu2−xS are highly dependent on the stoichiometric factor, 2 − x. However, it is not always straightforward for synthesizing a copper sulfide product with targeting phase; usually, mixed phase copper sulfide products are often obtained through usual synthesis strategies. Up to date, some methods including phase transformation route have been devoted for the preparation of copper sulfide products with pure specific phase.26,33–35 For example, it was demonstrated that Cu7S4 and CuS can be obtained from the freshly formed Cu9S8 nanocrystals.36 With a solid-state annealing route, CuS also undergoes a reduction to tetragonal cuprous sulfide Cu2S.37 Recently, Fang et al. reported facile synthesis of Cu39S28 microcrystals via a solvothermal route.38 Alivisatos et al. has demonstrated the temperature-induced structural transformations of Cu2S nanorods from a low to a high chalcocite structure and investigated their size dependent phase transformation behaviour.39,40 Zhong et al. reported a phase transformation process from rhombohedral Cu1.8S nanocrystals to hexagonal CuS clusters.41 In spite of these valuable investigations, relative to highly studied PbS and CdS micro-/nanoscale materials, copper sulfides have not been extensively studied. In addition, as copper sulfide easily forms different but close stoichiometric phases, controlling of copper sulfide crystal phase with simple method is still a challenge.42–44
Among the various Cu2−xS materials, Cu1.8S, which is a useful p-type semiconductor with indirect bandgap of 1.5–1.6 eV, has got much attention due to its rich properties and potential applications.14,40,45,46 Cu1.8S is known to be used as thermo- or photoelectric transformers and high-temperature thermistors. The higher copper deficiency in Cu1.8S material makes it show unique localized surface plasmon resonance effect (LSPR).40 Due to the LSPR effect, Cu1.8S has shown excellent photothermal conversion effect and would be used as new type of photothermal agent.45,46 Especially, it was also demonstrated that Cu1.8S with higher ion deficiency can be used as catalyst inducing the formation of heterostructures.14
In this study, a simple and low cost hydrothermal method was developed for the synthesis of copper sulfide product. The obtained product contains two kinds of copper sulfides (CuS and Cu1.8S). Based on this, we described a new chemical extraction route with triphenyl phosphine (TPP) as extraction agent. This extraction process can transform CuS, Cu39S28 phases into pure low-sulfur metastable rhombohedral phase Cu1.8S. This chemistry has at least three important implications. First, sulfur-rich copper sulfides can be chemically transformed into Cu1.8S phase via TPP extraction of sulfur. This highlights the possibility of precise Cu1.8S phase targeting with this route. Secondly, impure samples contain mixtures of multiple copper sulfides including CuS, Cu1.8S, Cu39S28 can be purified by TPP extraction of sulfur, with the final product exclusively being metastable rhombohedral Cu1.8S phase. Third, it seems that the Cu1.8S product obtained with purification route shows similar properties to that of Cu1.8S sample obtained with direct preparation route.
We also investigate the morphology change of the copper sulfide products during the phase transformation process. Fig. 2 shows the SEM and TEM images of the mixture phase copper sulfides. It can be seen that the original mixture phase copper sulfide product is composed of flower-like micro-spheres with size of 3–5 μm (Fig. 2a and b). The flower-like microspheres are assembled by sheet-like units. Some irregular particles are loaded on the flower-like microstructure, suggesting the mixture phase. TEM image shows similar results (Fig. 2c). Inset of Fig. 2c shows the selected area electron diffraction pattern (SAED). The pattern can be indexed into CuS and Cu1.8S, although they have similar crystal plane spacing. Two kinds of clear lattice fringes with different spacing in the high-resolution TEM indicate the high crystallinity and two different crystal phases involved in it (Fig. 2d). The lattice fringe with spacing of 0.278 nm would be indexed to (10 10) plane of Cu1.8S, while that of 0.198 nm can be attributed to (008) plane of CuS. After phase purification with TPP, the morphology of the copper sulfide changed. The assembled flower-like microspheres are broken, leaving disorder sheet-like units (Fig. 3a and b). The corresponding SAED pattern and HRTEM shows pure Cu1.8S phase with high crystallinity (Fig. 3c and d). The morphology change indicates the chemical reaction between copper sulfide and TPP.
![]() | ||
Fig. 2 (a and b) SEM, (c) TEM, and (d) HRTEM images of the mixed phase copper sulfides. Inset of (c) shows the SAED pattern of the product. |
It was found that the formation of Cu1.8S phase by TPP treatment is highly phase-targeting. Increasing the treatment temperature to 200 °C, the obtained product is still Cu1.8S phase. Interestingly, other copper sulfide phase, for example Cu39S28, can also be transferred into the rhombohedral Cu1.8S phase by this phase transform route. The corresponding results are shown in Fig. 4. Cu39S28 sample was synthesized according to a reported method. Although the XRD patterns are difficult to distinguish owing to the relatively weak crystalline and the close XRD patterns between Cu39S28 and Cu1.8S phase, after TPP treatment, an obvious shift of the strongest peak is clearly observed, suggesting the crystal phase transformation.
These results suggest that TPP can extract sulfur from sulfur-rich copper sulfide, forming metal-rich Cu1.8S phase.
This extract process is quite different from the previous reported Cu(I) induced phase transformation route.43 In that case, Cu(I) specie was found to induce Cu1.1S phase to Cu1.1–1.5S phase. It was proposed that a fraction of Cu+ ions from the solution enters the Cu1.1S lattice, matched by a transfer of electrons from the solution. In our route, it seems that TPP extract sulfur from sulfur-rich covellite CuS product, causing the phase transformation. There are reports about crystal structure of hexagonal covellite CuS.47–50 Studies have shown that the valence of copper in covellite CuS is monovalent with formalism of (Cu+)3(S22−)(S−) or (Cu+)3(S2−)(S2−).43,47 Recent experiments and calculations have put the valence of Cu between 1 and 1.5 (1.33 for CuS from calculations).50 While, all studies indicate that sulfur in covellite phase exists with various valence or forms. Fig. 5 shows the crystal structure illustration of covellite CuS phase and digenite Cu1.8S. From the structural point of view, it is important to note the similarities of the Cu or S sub-lattice between covellite and digenite in the corresponding close-packed planes, i.e. (001) and (10−1), respectively. It would make it easily to transfer covellite CuS to digenite Cu1.8S under suitable reaction conditions.
In our case, the extract route is possibly driven by the large formation constant of the TPP–S complexes. The bonding of sulfur and TPP would be stronger, which cause the following reaction proceed.
CuS + TPP → Cu1.8S + TPP–S |
In the phase transformation process, solid copper sulfide reacts with liquid TPP causing the phase transformation. Thus the reaction rate would be controlled by the solid–liquid interface reaction.
In the presence of copper sulfide catalyst, hydrazine will gradually reduce MB at room temperature, forming colorless LMO (as shown in Scheme 1). While, upon further shaking reaction system in air, the colorless LMO will be oxidized again by air forming blue MB. The redox process demonstrates a simple “clock reaction”, which provides an engaging illustration of redox phenomena and reaction kinetics. As shown in Fig. 6, the original blue reaction solution gradually becomes weak with the increasing of duration time. With duration time of about 20 min, only the liquid–air interface shows blue color, while the solution body becomes colorless. This is reasonable since oxygen in air will oxidize LMO. Further prolonging the duration time to 3 h, the air in the sealed system is possibly exhausted, so the whole solution is colorless. Once shaking the reaction system in air, blue color quickly reappears.
![]() | ||
Fig. 6 Schematic representation of the “clock reaction” catalyzed by Cu1.8S, that is, blue color fading and regeneration of MB. |
Based on the fact that MB exhibits an intense absorption band in the region of 500–700 nm and LMB shows no absorption at this region, we monitored the process of the clock reaction by a UV-visible spectrophotometer. With the MB blue color bleaching, a steady decrease of the absorbance of MB was measured at same intervals, as shown in Fig. 7a. It should be noted that in the absence of copper sulfide catalyst, no such a decrease in the absorbance of MB was observed in the same experimental condition. Thus, the important role of the synthesized copper sulfide product in the “clock reaction” is no doubt concluded.
![]() | ||
Fig. 7 (a) Absorption spectra for successive decolorization of MB and (b) colour regeneration catalyzed by Cu1.8S. |
It seems that the synthesized Cu1.8S product has a strong catalytic ability for the reduction of MB by hydrazine at room temperature. We also investigated the catalytic activities of mixed phase copper sulfides and Cu39S28 product. In contrast, the catalytic activities of them are relatively low (Fig. SI-1 and SI-2, see ESI†). After complete reduction of MB, the solution containing colourless LMB can again regenerate the blue color in the presence of small amount aerial O2 on shaking of the reaction mixture openly and so the characteristic absorption band appears (shown in Fig. 7b).
We then investigated the influence of various experimental parameters on the reaction kinetics. It was found that without hydrazine, the reaction cannot proceed. More hydrazine is involved, the reaction rate is quickly. While, it seems that the catalyst dosage don't have obvious influence on the reaction rate in our investigated range. In the presence of copper sulfide catalysts, the plot of absorption factor as a function of time (Fig. 8a) shows a profile of exponential equation, At = A0e−kt, which is consistent with a pseudo-first-order reaction. Thus, pseudo-first-order reaction kinetics was applied for the evaluation of catalytic activity in our case. The relation of ln(At/A0) (at peak of 664 nm) versus time is shown in Fig. 8b. The slope of the straight line gives the rate constant of the catalytic reaction, 0.0323 min−1 for pure Cu1.8S. While, mixed phase copper sulfide sample shows similar and relatively lower rate constant, 0.0297 min−1. This indicates that the purified procedure don't change the catalytic activity of the copper sulfide product. In addition, the Cu39S28 product shows the poorest catalytic activity among the tested three samples. It is proposed that the tiny composition difference and the different crystal phase cause their different properties. Copper sulfides with different phases have different copper and sulfur ions arrangement in the crystal, which will cause the different chemical surroundings for the copper or sulfur ions. This finally induces the different catalytic or related properties.
![]() | ||
Fig. 8 Plots of (a) At and (b) ln(At/A0) of MB vs. reaction time in the presence of mixed copper sulfides, Cu39S28, and pure Cu1.8S. |
We then proposed a possible catalytic principle for the reaction. It is proposed that copper sulfide would promote the electron transfer process from hydrazine to MB.51 During the reaction, hydrazine and MB may be absorbed onto the copper sulfide surface due to the reasonable affinity. It is believed that hydrazine supplies electrons to MB via copper sulfide, causing the generation of colorless LMB. Hydrazine transfers electrons to Cu(I) to reduce it to Cu(0). Then, MB and dissolved oxygen in the system oxidizes Cu(0) back to Cu(I).51 During this process, MB was reduced forming colorless LMB. The electron-transfer processes can be supported by the redox potential values of φ(N2/N2H4) = −1.16 V, φ(Cu+/Cu) = 0.52 V, φ(MB/LMB) = 1.08 V.51,54 In the reaction system, excess hydrazine in turn decreases the dissolved oxygen concentration in water, which facilitates LMB formation. While, indeed, detailed catalytic mechanism needs further study.
In addition, it is recently reported that Cu–S system is a new promising semiconductor photothermal conversion platforms with relatively high photothermal conversion efficiency, good photostability, synthetic simplicity, low toxicity and low cost.42 It can effectively induce temperature elevation under NIR irradiation, which would provide a possible route for cancer treatment.
The NIR photothermal conversion property of the obtained Cu–S samples in the aqueous dispersions was then examined at a fixed concentration of 0.03 mg mL−1. The results are shown in Fig. 9. The pure water system was also tested for comparison. It is obvious that, for the pure water, the NIR irradiation (980 nm) caused a temperature increase of only about 4 °C after 8 min. For the aqueous dispersion of Cu–S products, the NIR irradiation induced temperature elevation is much higher than the pure water. For the mixtured sample with CuS and Cu1.8S phase, the temperature increase is 6 °C, while the pure phase Cu1.8S sample gives a higher temperature increase of 7 °C, suggesting the better photothermal conversion property of Cu1.8S than that of CuS. It should be noted that all the aqueous dispersions reach a temperature platform with irradiation for 8 min, suggesting a thermal equilibrium between laser energy input and thermal diffusion towards environment are obtained at this stage. After that, the laser is closed, the temperature then decreases to room temperature.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra17652b |
This journal is © The Royal Society of Chemistry 2015 |