Open Access Article
Zeng-Wen
Hu‡
,
Liang
Xu‡
,
Yuan
Yang
,
Hong-Bin
Yao
,
Hong-Wu
Zhu
,
Bi-Cheng
Hu
and
Shu-Hong
Yu
*
Division of Nanomaterials & Chemistry, Hefei National Laboratory for Physical Sciences at the Microscale, Collaborative Innovation Center of Suzhou Nano Science and Technology, CAS Center for Excellence in Nanoscience, Department of Chemistry, Hefei Science Center of CAS, University of Science and Technology of China, Hefei 230026, China. E-mail: shyu@ustc.edu.cn
First published on 9th March 2016
Two-dimensional inorganic nanomaterials have drawn much attention due to their excellent properties and wide applications associated with unique 2D structures. However, an efficient and versatile chemical synthesis method using ambient conditions for 2D nanomaterials, especially with secondary structures (e.g. mesopores), has still not been reported. Herein, we report a versatile method to synthesize a family of ultrathin and mesoporous nanosheets of metal selenides based on a precursor so-called “red Se remaining Zn” (RSRZ). The principle of our synthesis is based on a template-assisted chemical transformation process via acidification of inorganic–organic hybrid ZnSe(DETA)0.5 nanosheets (DETA: diethylenetriamine). An appropriate amount of acid was added into an aqueous dispersion of ZnSe(DETA)0.5 nanosheets under air for activation. The acidification induced chemical transformation mechanism was studied by tracking the acidification process. This acid controlled reactivity of lamellar hybrids allows it to be possible to capture the highly reactive intermediates, which will provide a new platform for the synthesis of various mesoporous metal selenides.
Herein, we report a simple and versatile chemical transformation method to prepare ultrathin and mesoporous metal selenide nanosheets starting from ZnSe(DETA)0.5 nanosheets using a acidification process. We find that highly reactive nanosheet intermediates called “red Se remaining Zn” (RSRZ) can be prepared by using hydrochloric acid to acidify the ZnSe(DETA)0.5 nanosheets, which can act as an excellent template for the chemical transformation reaction. Addition of hydrochloric acid to the amine-assisted hybrid precursors would result in depletion of the amine in the lamellar hybrid structure, followed by dissociation and oxidation of ZnSe. The obtained RSRZ nanosheets can be easily transformed into a family of metal selenide nanostructures including Ag2Se nanosheets, Cu2Se nanosheets, PtxSey alloy nanosheets, PdxSey alloy nanosheets, and Se nanowires under ambient conditions.
The lamellar structured ZnSe(DETA)0.5 hybrids were synthesized as previously reported26,31 and used as the starting materials. The as-prepared ZnSe(DETA)0.5 nanosheets were examined using scanning electron microscopy (SEM) and X-ray diffraction (XRD). The SEM image (Fig. 1a) showed that the ZnSe(DETA)0.5 nanosheets have a thickness of ∼50 nm. The typical XRD pattern (Fig. S1a, ESI†) further confirmed that the hybrid precursor is the same as our reported ZnSe(DETA)0.5 nanobelts.26 The red floccules precipitated within hours after addition of the hydrochloric acid into the aqueous solution of the ZnSe(DETA)0.5 precursor. SEM images (Fig. 1b and c) clearly demonstrated that the size of the RSRZ nanosheets was inherited from the size of the hybrid precursors, except for the thickness which ranged from several nanometers to 30 nanometers. Atomic force microscopy (AFM) images (in Fig. 1d and S1b†) clearly showed that a large part of the RSRZ nanosheets had a thickness of ∼5 nm, indicating that RSRZ nanosheets were successfully exfoliated from the bulk ZnSe(DETA)0.5 hybrid with the help of acid. Some incompletely exfoliated nanosheets showed a lamellar structure with a thickness larger than 20 nm (Fig. 1c and S2d, ESI†).
Microscopy characterization clearly showed that the formed RSRZ nanosheets were highly porous (Fig. 1e) and the surface was very rough (Fig. 1c and S1b, ESI†), indicating a potentially large specific surface area. Nitrogen adsorption–desorption isotherms of the RSRZ nanosheets are shown in Fig. S1c (ESI†), which revealed that the RSRZ nanosheets have a BET surface area of 78.76 m2 g−1 and a total pore volume of 0.15 cm3 g−1. The distribution curve of the pore sizes in Fig. S1d† shows that the RSRZ nanosheets have a narrow pore size distribution around 3 nm, indicating mesoporous properties of the as-obtained RSRZ nanosheets.
To investigate the chemical composition, it is necessary to proceed with elemental analysis. Energy dispersive spectroscopy (EDS) elemental mapping images (Fig. 1f) of the RSRZ nanosheets obtained after acidification for 10 h confirmed that Se occupied a majority of the elemental composition, which still contained a fraction of Zn elements. In addition, the FTIR spectra (Fig. S1e, ESI†) confirmed that there was no organic DETA in the RSRZ nanosheets at all after 10 hours of acidification, suggesting that the removal of the amine from the ZnSe(DETA)0.5 hybrid precursors was carried out very thoroughly. The peaks in the FTIR near 3430 and 1630 cm−1 could be attributed to the stretching and bending vibrations of –OH from absorbed water, indicating oxygen or water adsorption on the surface of the as-obtained RSRZ nanosheets. Raman spectra (ESI, Fig. S1f†) obtained for the RSRZ nanosheets generated from acidification for 10 h showed a sharp peak at 255 cm−1, which corresponds to disordered chain-like Se molecules32–35 or monoclinic selenium.36 The weak and broad peak near 495 cm−1 in the Raman spectra demonstrated that the RSRZ nanosheets after acidification for 10 h still contained a fraction of ZnSe.37 Hence, it was difficult to define the novel sheets as a single phase. We call the intermediates “red Se remaining Zn” (RSRZ) considering their colour (Fig. 1b) and composition.
H+ is so small that it can diffuse into a lamellar structure and attack amines in the ZnSe(DETA)0.5 nanosheets. So the pH value is especially important for the reaction system. As a result, the pH was varied from 0.1 to 1 at the same concentration of precursor (see Experimental section 2). TEM and SEM images of the RSRZ nanosheets formed under different pH conditions are shown in Fig. S2(a and b) for pH 0.1, Fig. S2(c and d)† for pH 0.5, and Fig. 1b, c and e for pH 1, respectively. As these images showed, the sheets obtained with pH 0.1 were the most broken. This indicated that the acidity contributes to the porosity of the RSRZ nanosheets. TEM images (Fig. S2(e and f), ESI†) demonstrated that feeding with oxygen or a long acidification time contributed to forming large particles. A higher concentration of precursor, higher temperature and constant stirring would result in nanoframes (Fig. S2g, ESI†). In addition, the phase transformation of the ZnSe(DETA)0.5 hybrid during the acidification process was studied using XRD (Fig. S2h, ESI†). The broad peaks of the RSRZ nanosheets after acidification for 2 h could be indexed as hexagonal ZnSe (JCPDS: 15-0105). With prolonging of the reaction time, the RSRZ sheets were gradually acidified by hydrochloric acid and post-oxidized into Se by dissolved oxygen in the solution. The XRD peaks of the sample after acidification for 24 h can be indexed as t-Se (JCPDS: 01-0853).
The aforementioned characterization of the acidification product of ZnSe(DETA)0.5 with reaction time showed that the hybrid became a metastable ZnSe phase first and then a stable Se phase by oxidation with oxygen in water, but the details of the transformation of the hybrid into metastable ZnSe were still unclear. To get a better understanding of the stability and reactivity of the ZnSe(DETA)0.5 lamellar hybrid during the acidification process, we explored the phase transformation mechanism by tracking the acidification process at time intervals (see Experimental section 3).
The real-time changes of the UV-vis absorption spectrum, the amount of oxygen, the conductivity and the pH value of the reaction system at 30 °C are summarized in Fig. 2a and b. The UV-vis absorption spectrum at 0 min in Fig. 2a is exactly consistent with the previously reported results for ZnSe(DETA)0.5.26 It is obvious that the peak position of the UV-vis absorption spectrum changed significantly from 0 min to 10 min, indicating that the inorganic–organic hybrid structure of ZnSe(DETA)0.5 was destroyed under the acidic conditions. The broad peak near 250 nm in the UV-vis spectrum for the sample obtained after acidification for 10 min matched with the bandgap of ZnSe.5,38 In addition, electron energy loss spectroscopy (EELS) was used to analyse the N and O elements of the sample after acidification for 10 min. The smooth curve in Fig. 2c demonstrates that there were hardly any N and O elements in the sheet. It was observed that just a little amount of N elements was distributed only on the edge of the sheet from the energy filtered transmission electron microscopy (EFTEM) images (Fig. 2c). It was also apparent that the content of the O element was larger than the N element using the EFTEM images. As we know, the O element could only occur from adsorption because the ZnSe(DETA)0.5 itself did not contain O elements. The small quantity of O elements existing in the sheet may come from absorbed O2 or H2O. These results illustrated that almost all DETA (C4H13N3) inside the hybrid sheet was successfully depleted and diffused into solution after acidification for 10 min. In other words, the ZnSe(DETA)0.5 inorganic–organic hybrid was transformed into inorganic ZnSe. As the reaction proceeded with time, the peak near 250 nm became weaker and weaker, suggesting that more and more inorganic ZnSe was destroyed. This indicated that oxidation had happened. After addition (at 0 min in Fig. 2b) of the hybrid precursors into water which was preheated and adjusted to a desirable pH value in advance, the dissolved oxygen, conductivity and pH of the solution showed an extreme change in the first few minutes because it took a few minutes to mix the solution homogeneously. Then, the dissolved oxygen amount went down slowly because of the consumption of oxygen due to oxidation in the solution being faster than diffusion of oxygen from the air. The phenomenon that the conductivity went down and the pH value went up slowly accounts for consumption of H+. All of these proved that oxidation of ZnSe occurred during the acidification process. Thus, the mechanism of the whole acidification included two stages: H+ attack on DETA (exfoliation) and oxidation. The related chemical reactions and apparent rate equation for each stage are shown below (the details for all the equations are shown in the ESI†):
Stage 1:
![]() | (1) |
Stage 2:
![]() | (2) |
k = A exp(−Ea/RT) | (3) |
ln xO2 = A + B/T*, T* = T/100 K | (4) |
| PO2 = HxO2 | (5) |
To further validate the mechanism, we compared the HRTEM images of a sample obtained after acidification for 120 min with that for 10 min (Fig. 2d and e and S3, ESI†). The crystal lattices corresponded to hexagonal ZnSe (0002) planes for the sample obtained after 10 min (the molar ratio of Zn
:
Se is 56
:
44 based on EDS elemental mapping analysis shown in Fig. S4a, ESI†). The lattice fringes of the sample after acidification for 120 min became so ambiguous that two lattices combined together, showing an average spacing of 0.667 nm. This means that the crystal structure had been destroyed to some degree during the acidification from 10 min to 120 min, owing to oxidation of ZnSe. Furthermore, real-time tracking of the acidification reaction at different temperatures was carried out (Fig. S4(c and d), ESI†). The process of exfoliation at 50 °C lasted only one minute, but it extended to ten minutes at 10 °C. On the contrary, the process of oxidation at 50 °C was much slower than that at 10 °C. The reason for this is that a high temperature is favourable for increase of the rate constant k but also lower oxygen levels in solution. According to eqn (1) and (2), we might draw the conclusion that the reaction rate of stage 1 went up while that of stage 2 went down at high temperature.
As is known, the redox potential is the criterion for a redox reaction to occur. The oxidation of ZnSe at stage 2 could be divided into three steps as shown in eqn (6)–(8). Using eqn (4) and (5), the molar concentration of dissolved oxygen in water was 2.36 × 10−4 mol L−1 when the temperature was 30 °C (303 K). Thus, the potential of stage 2 in solution was approximately 1.7 V according to the Nernst equation (eqn (9)) when the pH was 1.5 and the molar concentration of Se2− was
, indicating that the redox reaction would proceed quite easily and thoroughly. Moreover, a gas–solid reaction was extremely likely to happen in the acidic fluid medium. The stabilizer-depleted ZnSe layers were far from the equilibrium state and tended to absorb oxygen and release Zn2+, considering that they were ultrathin and porous. All of this means that the captured intermediates during the acidification process were always obtained along with poor stability and high reactivity.
| ZnSe ↔ Zn2+ + Se2−, KSP = 3.6 × 10−26 | (6) |
| 0.5O2 + 2H+ + 2e− ↔ H2O, Eθ = 1.229 V | (7) |
| Se + 2e− ↔ Se2−, Eθ = −0.924 V | (8) |
![]() | (9) |
Interestingly, the RSRZ sheets have a high chemical activity under ambient conditions making them suitable as a new platform for the synthesis of 2D metal selenides through a facile chemical transformation while maintaining the 2D structures. The RSRZ nanosheets can be easily transformed into Ag2Se and Cu2Se nanosheets through a chemical transformation process by simply adding AgNO3 or CuCl into a suspension containing the RSRZ nanosheets. The XRD patterns of the chemical transformation products are shown in Fig. 3a, which could be indexed as β-Ag2Se (JCPDS no. 71-2410) and cubic Cu2Se (JCPDS no. 88-2043), respectively. SEM images of the as-synthesized Ag2Se and Cu2Se are shown in Fig. 3b and S5a,† respectively. Compared to the RSRZ nanosheets, the pore size of the Ag2Se and Cu2Se nanosheets became much larger, as indicated by the SEM images. In addition, many particles were attached on the surface of the Cu2Se nanosheets, which might be due to an Ostwald ripening process during the reaction.
The unique properties and exorbitant price of noble metals like Pt draw intensive attention to Pt-based heteronanostructures or alloy catalysts.39–41 Xia’s group reported that hollow nanostructures of Pt could be synthesized by templating with Se nanowires and colloids in 2003.42 The RSRZ nanosheets reported here can certainly be used for templating the synthesis of PtxSey and PdxSey alloys, which is different from the Se@Pt reported previously.33
TEM, HRTEM and SAED images of the PtxSey alloy nanosheets are shown in Fig. 3c and 4a, b and e. Obviously, the PtxSey alloy nanosheets were quite rough, porous and polycrystalline. The lattice spacings of 0.224 nm and 0.194 nm are close to the (111) and (200) facets of Pt (JCPDS # 04-0802), which is consistent with the results from X-ray diffraction (Fig. 4d). EDS elemental mapping analysis (Fig. 4c) demonstrated that Pt and Se elements were uniformly distributed in a single nanosheet, but these nanosheets still contained a tiny minority of Zn elements. The accurate chemical composition was analyzed using inductively coupled plasma (ICP) emission spectroscopy. As a result, the composition was Pt3Se2, only containing 0.25% Zn. Pt 4f XPS spectra (Fig. S5b, ESI†) indicated that most of the Pt was in the metallic state. This template synthesis could also be applied to Pd, and the related details are shown in the ESI (Fig. S5(c–f) and S6(d and f)†). The difference was that the Pd63Se37 alloy nanosheets were amorphous and some large particles were attached to the sheets.
In addition, the RSRZ nanosheets can also be transformed into t-Se nanowires by naturally ageing in ethanol. Fig. 3d shows a SEM image of the Se nanowires transformed from RSRZ nanosheets by ageing for two days. The diffraction peaks of the Se nanowires in Fig. 3a can be indexed to trigonal selenium (t-Se, JCPDS # 01-0853). The as-prepared t-Se nanowires may also act as a good template for the synthesis of a variety of nanowires.42–44
Amorphous Se tends to form larger colloids.36,43,45 Many large particles were observed in our experiment if the RSRZ nanosheets were oxidized more thoroughly (Fig. S2(e and f), ESI†). In particular, the amorphous Se was unstable in alcohol whereas the follow-up transformation would be carried out in a solution containing alcohol. However, the formation of large particles was not desired. Thus, it is better to transform the Se into MxSey as soon as the nanosheets are partly oxidized to Se, because the remaining ZnSe would be oxidized in the follow-up transformation by keeping the reaction system acidic in the presence of air. Almost all nanomaterials are usually far from the equilibrium state on account of the increased Gibbs free energy.46 Poor stability and high reactivity were always found together as a double-edged sword for nanomaterials as we demonstrated previously in the case of ultrathin tellurium nanowires in solution.47 The conditions for the formation of different target nanostructures by the acidification of a ZnSe(DETA)0.5 hybrid are summarized in Table S1 (ESI).†
:
14
:
16 to form a homogenous solution under constant strong stirring. The mixed solution was then transferred into a 50 ml Teflon-lined autoclave (with a filling ratio of 80%). The sealed vessel was then maintained at 140 °C for 12 h, and allowed to cool down naturally. The samples were collected and washed three times with water.
Nanotechnol., 2011, 6, 147–150 CrossRef CAS PubMed.Footnotes |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6sc00674d |
| ‡ Zeng-Wen Hu and Liang Xu contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2016 |