Elemental anion thermal injection synthesis of nanocrystalline marcasite iron dichalcogenide FeSe2 and FeTe2

Ebin Bastola, Khagendra P. Bhandari, Anthony J. Matthews, Niraj Shrestha and Randy J. Ellingson*
Wright Center for Photovoltaics Innovation and Commercialization, School for Solar and Advanced Renewable Energy, Department of Physics and Astronomy, The University of Toledo, 2801 W. Bancroft Street, Toledo, Ohio 43606, USA. E-mail: randy.ellingson@utoledo.edu

Received 9th March 2016 , Accepted 14th July 2016

First published on 15th July 2016


Abstract

We report a hot-injection colloidal method for the synthesis of nanocrystalline (NC) iron diselenide (FeSe2), and iron ditelluride (FeTe2) derived from iron(II) bromide as the iron (Fe) precursor. These highly crystalline iron chalcogenides are synthesized by injecting elemental selenium (Se) or tellurium (Te), complexed with oleylamine, into a reaction flask containing the iron precursor. NC FeSe2 and FeTe2 films have been characterized by using X-ray diffraction (XRD), scanning electron microscopy (SEM), and Raman spectroscopy. Both FeSe2 and FeTe2 NC particles exhibit marcasite phase with orthorhombic crystal structure. The as-synthesized NC FeSe2 and FeTe2 show irregular shapes with consistent stoichiometric ratio. In addition, we discuss these iron chalcogenides with regard to their thin film electronic properties including sheet resistance, resistivity, majority carrier type, and their possible device applications.


Introduction

Semiconductor nanocrystalline materials offer many promising properties for electronic and photonic applications.1,2 Potentially simplified material preparation and solution-sourced thin film processing from nanocrystals enable consideration of inexpensive inkjet or screen printing on rigid or flexible substrates. In many cases, solution-sourced nanocrystals can be deposited in thin film form at room temperature to yield excellent opto-electronic performance.3–5 Semiconductor nanocrystals and nano-composites have great potential in the field of photovoltaics (PV),6–8 medicine,9 light emitting devices,10,11 and photo-catalysis.12 Researchers routinely search for alternative materials to improve upon those presently used commercially, in an effort to reduce material input or manufacturing cost, enhance performance, or reduce undesired environmental costs. For example, several studies in recent years have demonstrated the possible application of earth abundant chalcogenides13–15 which can reduce toxicity in photovoltaic and other optoelectronic devices. While much attention has been drawn recently to the understanding and development of iron pyrite (FeS2), here we discuss the colloidal synthesis and properties of nanocrystalline (NC) iron diselenide (FeSe2) and iron ditelluride (FeTe2), with grain sizes on the order of or less than 100 nm.

Several authors have reported the synthesis of iron dichalcogenides by methods including high temperature solid state reaction,16 high temperature and high pressure,17 molecular precursor solution synthesis,18 ultrasonic chemical synthesis,19 and metal organic chemical vapor deposition (MOCVD).20 However, the most commonly reported techniques for synthesizing FeSe2 and FeTe2 are hydrothermal,21–23 and solvothermal methods.24 Though the wet-chemical method is an easy route, very few reports exist for iron dichalcogenide NC synthesis by this approach. Iron dichalcogenides, especially iron pyrite (FeS2), have been studied for many decades with interest in their applications to photovoltaics and other energy conversion processes. Several groups have studied iron pyrite as the potential absorbing layer in solar cells, with limited success owing to the material's tendency for very high ionized defect densities and strongly limited photovoltage.25,26 Our previous work has shown that FeS2 NCs can serve as an interface material for a low barrier back contact to CdTe solar cells.5 Other iron dichalcogenides, FeSe2, and FeTe2, although their fabrication using colloidal methods has slowly begun, have not been well explored for possible applications as opto-electronic materials. Although both materials have been prepared at elevated temperature in the cubic pyrite structure, they readily form the marcasite orthorhombic structure. Early studies indicated that the marcasite forms of FeSe2 and FeTe2 are narrow band gap semiconductors with complicated temperature-dependent transport properties.27 In addition, FeSe2 and FeTe2 materials possess magnetic properties,22,23,28 and based on a theoretical study, FeSe2 offers distinct promise as a thermoelectric material.29 Furthermore, authors have reported the performance of solution-processed FeSe2 thin films as counter electrodes with performance comparable to the conventional platinum electrodes in dye-sensitized solar cells.30,31 Li et al. have reported the synthesis of FeSe2 nanoparticles exhibiting laminar morphology using triethylene glycol as solvent, triethyleneterramine as an assisting agent, and polyvinylpyrrolide as a dispersing agent.32 Recently, Qin et al. have also reported hot-injection synthesis of FeSe2 NCs by using iron(II) acetylacetonate, and elemental selenium,33 and recent work reported the synthesis of 30–100 nm FeSe2 NCs in the presence of ethylenediamine.34 However, these authors have not described the dependence of growth time on particle shape and size, and on the thin film electronic characteristics. Similarly, very few solution based methods have been reported to prepare FeTe2 nanoparticles. Oyler et al. have reported a solution based method to grow nanosheets of iron chalcogenides (FeSe, FeTe, Fe(Se,Te), and FeTe2),35 and Zhang et al. have synthesized FeTe2 NCs by dissolving elemental Te and the iron(II) complex Na2[Fe(EDTA)] in aqueous solution at 140 °C.36 The synthesis of NC FeTe2 by hot-injection colloidal method using iron(II) bromide as Fe source, elemental Te, and 1,2-hexanediol as surfactants, to the best of our knowledge, has not been previously reported.

We present a hot-injection colloidal synthesis of NC FeSe2, and FeTe2, using iron(II) bromide37 as the cation precursor, and elemental Se and Te as anion sources. Marcasite phase NCs with orthorhombic crystal structure are grown in the temperature range of 200 °C to 340 °C using 1,2-hexanediol as the surfactant to enhance the yield of NC FeX2 (X = Se, Te) control of nucleation and growth. Further, the shape and size of the NCs depend on the growth durations, and here, the influence of reaction time on shape and size of the NCs has been investigated by varying the growth interval following anion precursor injection.

Experimental

Chemicals

Materials utilized include iron(II) bromide (FeBr2, anhydrous 99.9%; Alfa-Aesar), oleylamine (OLA, 70%; Sigma-Aldrich, or S.A.), 1,2-hexanediol (S.A.), Se powder (S.A.), methanol (CH3OH, 99.5%; S.A.), chloroform (CHCl3; Fisher-Scientific), Te powder (anhydrous, 99.9%; S.A.), and trioctylphosphine (TOP, S.A.). The elemental Se and Te powders were stored in a nitrogen environment glove box to avoid the formation of any hazardous compounds such as H2Se and H2Te gases. All chemicals are used as obtained without further purification, and the reactions were performed under nitrogen atmosphere.

Synthesis procedure

Iron dichalcogenide (FeSe2 and FeTe2) syntheses were done under nitrogen atmosphere using standard Schlenk line technique. The typical method used to synthesize FeSe2 NCs was similar to that reported previously for iron pyrite (FeS2) NCs.37 Iron(II) bromide (FeBr2) was the source of Fe2+ ions, and elemental Se was used as the Se22− anion source. In a three neck flask, 0.5 mmol FeBr2, 0.4 mL 1,2-hexanediol, and 10 mL OLA were taken; the three-neck flask was evacuated and purged with nitrogen, repeated twice, to remove residual air. The mixture was then heated at 170 °C for 2 hours to dissolve FeBr2 into OLA. To prepare the Se source, 3 mmol Se was mixed with 10 mL OLA in a vial in the glove box. The Se mixture was sonicated for ∼15 minutes, and heated at ∼200 °C until the Se powder was completely dissolved in OLA. To support temperature control and stability during NC nucleation, the dissolved Se in OLA was held at 200 °C until the time of injection into the reaction flask, and the mixture was held at the same temperature for the growth of NCs over different time durations. After the desired growth time of NCs, the heating mantle was removed and the flask was allowed to cool down to room temperature. The NCs thus synthesized were cleaned by centrifugation at 2400 × g in chloroform with methanol as the non-solvent, and the purified NCs were stored dry in an N2 glove box for further characterization. Similarly, to synthesize FeTe2 NCs, the Fe2+ source in the reaction flask was prepared identically as described above, and the Te source was prepared by adding 3 mmol of elemental Te to 3 mL TOP. The Te source was sonicated ∼15 minutes to dissolve the Te, and then ∼7 mL OLA was added into it. The mixture was then injected to the reaction flask at 300 °C, and heated for different time durations to study the growth patterns of NCs at the same temperature. The FeTe2 NCs were purified as described above. For both FeSe2 and FeTe2 syntheses, the temperature was varied to determine optimal conditions for high yield and crystallinity of NC marcasite.

Thin film preparation

Nanocrystalline thin films were prepared using layer by layer drop-casting onto a sodalime glass substrate in the fume hood. For this, FeX2 NCs were dissolved in chloroform at a concentration of ∼6 mg mL−1. To disperse the NCs fully in the solvent, the solution was sonicated with Digital Ultrasonic Cleaner Model PS-10A, and ultrasonic power of 70 W.

Characterization

The X-ray diffraction pattern of the FeX2 films were taken using a Rigaku Ultima III X-ray Diffractometer fitted with small angle X-ray Scattering (SAXS) at 40 kV accelerating voltage, and 44 mA current. Scherrer analysis was used to determine the average grain sizes for each sample. A Hitachi S-4800 UHR scanning electron microscope (SEM) was used to take the SEM images of the sample. Energy dispersive X-ray spectroscopy (EDS) was measured at an accelerating voltage of 20 kV to study the composition/stoichiometry of the materials. Optical absorption spectra of the NC solutions and the films were taken using a PerkinElmer Lambda 1050 UV/Vis/NIR spectrophotometer. To study the Raman vibrational peaks of the material, a Jobin Yvon Horiba confocal Raman spectrometer fitted with HeNe laser source of wavelength 632.8 nm was employed. A Lucas 4 Probe was used to determine the sheet resistance of the thin films, and film thickness was measured using a Dektak 1050 Profilometer. Hot probe measurements employed a lab-constructed apparatus to determine the sign of the Seebeck coefficient as an indication of majority carrier type in FeX2 thin films.38

Results and discussion

Crystal structure and phase analysis

Fig. 1(a) displays the X-ray diffraction (XRD) pattern of as-synthesized NC FeSe2. The sharp peaks in the XRD pattern confirms that the material synthesized was highly crystalline in nature, and indicates that the NC FeSe2 exhibit orthorhombic crystal structure (abc, α = β = γ = 90°). The crystal structure belongs to the space group Pnnm (58), with a = 0.4799 nm, b = 0.5777 nm, c = 0.3576 nm.39 Hence, the synthesized NC FeSe2 exists in marcasite phase with orthorhombic crystal structure. The vertical lines in the graph are the reference lines of the standard FeSe2 marcasite material obtained from the MDI JADE software (PDF 97-004-2041). With no evident discrepancy, these vertical lines match the peaks of the measured XRD of the synthesized sample which illustrates that these NCs maintain high phase purity.
image file: c6ra06351a-f1.tif
Fig. 1 X-ray diffraction pattern of as-synthesized nanocrystalline (a) FeSe2, and (b) FeTe2. Both materials possess orthorhombic crystal structure, marcasite phase, and the sharp intensity peaks of the pattern indicate phase-pure crystalline material. The reference lines for FeSe2 and FeTe2 materials are PDFs # 97-004-2041 and # 97-063-3879 obtained from MDI JADE software respectively.

Keeping the temperature of the Se precursor solution as close to 200 °C as possible, the Se solution was injected into the reaction flask containing the Fe precursor solution. The NC materials obtained at growth temperatures of 200 °C, 240 °C, 280 °C and 320 °C were investigated. Immediately after the injection of the Se precursor, the temperature of the reactants solution dropped to the range of 185 °C (for 200 °C starting point) to 240 °C (for 320 °C starting point), and rapidly returned to the initial temperature. In all these cases, NC FeSe2 so obtained are of marcasite phase with orthorhombic crystal structure. Interestingly, the properties of the NC product are independent of the injection temperature within this reaction temperature range. The XRD pattern of all these samples are shown in Fig. SF1 in the ESI. The XRD pattern presented here are similar to the previous reports obtained from hydrothermal and solvothermal methods.23,24 The XRD peaks of NC FeSe2 samples match, independent of reaction temperature, and the sharp peaks of the pattern confirm that these materials are highly crystalline in nature. The only discernible temperature-dependent property of the reaction we note was that yield improved for 240 °C and higher growth temperatures.

The average grain sizes of the NC FeSe2 were estimated by using Debye–Scherrer analysis given by, image file: c6ra06351a-t1.tif where Dp is the grain size, K is the shape factor usually taken as close to unity, λ is the wavelength of the X-ray used, β is the full width at half maximum (FWHM) in radians, and θ is the Bragg's diffraction angle. Based on the XRD pattern presented in Fig. SF1 in the ESI, the Scherrer-analysis was completed to estimate the grain size of the FeSe2 synthesized at different temperature, and the results are summarized in the Table 1. For this calculation, K = 0.94, and λ = 0.154059 nm were used in the equation, β was calculated by fitting each diffraction peak by Gaussian function and was corrected by using X-ray broadening (0.08° = 1.4 × 10−3 rad). The value of the grain size obtained from calculation is rounded to the nearest nanometer. The grain sizes of the FeSe2 NCs synthesized in the temperature range of 200 °C to 320 °C are similar in spite of different reaction temperature. Grain size analysis of the XRD pattern indicates that the sizes of the synthesized FeSe2 are still nanocrystalline in order, and the larger features in SEM imaging indicate coalescence.

Table 1 Scherrer-analysis of NC FeSe2 synthesized at different temperatures
S.No. Synthesis temperature (°C) Grain size based on the peak at (nm) Average grain size (Dp) (nm)
(101) (111) (120) (211)
1 200 67 58 62 47 59 ± 9
2 240 63 64 65 52 61 ± 6
3 280 59 64 62 48 58 ± 7
4 300 53 62 63 54 58 ± 5


Similarly, the XRD pattern of as-synthesized FeTe2 also shows sharp peaks indicating that the material is highly crystalline, as shown in Fig. 1(b). The XRD pattern corresponds with the pattern of iron ditelluride (PDF 97-063-3879) of the JADE software. These NC FeTe2 belong to marcasite phase with orthorhombic crystal structure with a = 0.5265 nm, b = 0.6265 nm, c = 0.3869 nm, and the crystal structure belongs to space group Pnnm (58).40 In addition, these XRD patterns of NC FeTe2 are similar to the XRD pattern reported previously.23,35 As with the FeSe2, the phase and crystal structure of FeTe2 were found to be independent of the growth temperature in the range of 220 °C to 300 °C. However, the yield was found comparatively better at 300 °C than at 220 °C and 260 °C. FeTe2 could be synthesized using only TOP to dissolve the Te powder, though the FeTe2 yield clearly improved when OLA was added to TOP. The existence of OLA is expected to assist with elemental Te dissolution through binding by the amine group.41,42 The XRD pattern of the NC FeTe2 grown at different temperature are shown in the Fig. SF2 in ESI, and the sharp peaks of these patterns again illustrate that the product exhibits excellent crystallinity independent of growth temperature. Similar to the NC FeSe2, the Scherrer-analyses for FeTe2 NCs synthesized at different temperatures were carried out on the XRD patterns presented in Fig. SF2 in ESI and the results appear in Table 2 with average grain sizes from ∼78 to 100 nm.

Table 2 Scherrer-analysis of FeTe2 NCs synthesized at different temperatures
S.No. Synthesis temperature (°C) Grain size based on the peak at (nm) Average grain size (Dp) (nm)
(011) (111) (120) (211)
1 240 102 109 100 89 100 ± 8
2 280 85 86 84 68 81 ± 8
3 300 83 79 79 73 78 ± 4
4 320 115 108 99 77 100 ± 17


Surface morphology and EDS analysis

To study the surface morphology of FeSe2, films were prepared by drop-casting onto sodalime glass substrates. The SEM images of the FeSe2 film are shown in Fig. 2(a and b). The NC FeSe2 films show flake-like surface morphology with feature sizes in the range of ∼150 nm to ∼1 μm. Qin et al. reported flower-like micron-sized FeSe2 crystallites by hot-injection synthesis while Yuan et al. have reported flake-like particles 100 nm to 200 nm in diameter by a solvothermal method.24,33 The EDS analyses of these NC thin films are shown in Table 3. The NC FeSe2 is found to be nearly stoichiometric with Se to Fe atomic ratio 1.99 ± 0.03, and we find excellent stoichiometric reproducibility within different syntheses.
image file: c6ra06351a-f2.tif
Fig. 2 Scanning electron microscopy (SEM) images of as-synthesized (a) and (b) FeSe2 nanocrystals (c) and (d) FeTe2 nanocrystals.
Table 3 EDS analysis of as synthesized FeSe2 and FeTe2. The instrumental error is estimated at ±3%
FeSe2 NCs FeTe2 NCs
Elements Atomic% Elements Atomic%
Fe K 33.54 Fe K 32.75
Se L 66.36 Te L 67.25


SEM images of FeSe2 NCs synthesized with different growth durations such as 1 minute, 5 minutes, 15 minutes, and 30 minutes after the injection of Se precursor to the reaction flask at 240 °C are shown in Fig. SF3 (ESI). We found that the yield of FeSe2 for 1 minute growth duration was relatively low while other growth durations 5 minutes and higher produced a similar improved yield. The NCs grown for longer durations show slightly larger average size than the NCs grown for 1 minute. Based on the SEM images, these NCs grown for different time intervals have very similar shape, and surface morphology. Here, all of the FeSe2 NC products show the pure marcasite phase with orthorhombic crystal structure, and the ratio Fe to Se atoms is ∼1[thin space (1/6-em)]:[thin space (1/6-em)]2. Further, we did not observe any change in color of the FeSe2 thin films during the characterization process indicating the films are stable in ambient conditions.

Similarly, FeTe2 thin films were also prepared by using layer by layer drop-cast method, to investigate the morphology of the NC FeTe2 thin films using SEM and EDS measurements. The SEM images of as-synthesized FeTe2 are shown in Fig. 2(c and d). The feature sizes of the FeTe2 thin films vary from ∼200 nm to ∼500 nm with apparently widely-varied crystalline shape. Zhang et al. have reported marcasite FeTe2 synthesized by the hydrothermal route yielding shape-nonspecific nanoparticles36 as well as nanorods.43 Similar to FeSe2, the NC FeTe2 was also found to be close to stoichiometric where the average atomic ratio of Te to Fe atoms is 2.04 ± 0.03. In addition, SEM images of the NC FeTe2 thin films resulting from injection and growth at 300 °C with different reaction time such as 1 minute, 5 minutes, 15 minutes, and 30 minutes are shown in the ESI, Fig. SF4. The SEM image belonging to 1 minute growth time (SF4a) is quite different than the rest of the SEM images. The XRD pattern of this 1 minute grown sample showed unidentified peaks in addition to the peaks of the orthorhombic crystal structure, and from the EDS measurement, Fe to Te atomic ratio was found to be close to 1[thin space (1/6-em)]:[thin space (1/6-em)]1. This indicates that the initial nucleation and growth within ∼1 minute growth time favors formation of a different crystalline structure, and that for FeTe2 the marcasite phase forms after nucleation. These FeTe2 NCs having longer growth time have a similar surface morphology as shown in Fig. SF4(b–d). As was observed for the FeSe2, the size of FeTe2 NCs depends on growth time, with longer growth time yielding slightly larger NCs. Similar to FeSe2 films, we did not notice any change in color or smell on FeTe2 thin films during the preparation, and characterization of this material indicating good stability of the materials. In addition, typical EDS measurement spectra are shown in Fig. SF5 of the ESI for both NC FeSe2 and FeTe2 (Table 3).

Raman spectroscopy

Fig. 3 shows the results of Raman spectroscopy measurements carried out using a 632.8 nm HeNe laser source to further characterize the NC FeSe2 and FeTe2 thin films. The FeSe2 film shows distinct Raman active modes at 180 cm−1, 220 cm−1 and 256 cm−1. Lutz and Müller measured Raman spectra for FeSe2 marcasite, attributing peaks at 221 cm−1 and 264 cm−1 to Ag and B1g Se–Se stretching modes, and a peak at 183 cm−1 to a librational mode;44 these peaks agree reasonably well with our measurements, and we note that the Raman laser excitation wavelengths differ slightly (632.8 nm in our case vs. 676.4 nm). These active vibrational peaks of Se–Se atoms closely agree with the peaks which have been previously reported for nanocrystalline FeSe2.24,33 We observe Raman peaks for NC FeTe2 thin films at 119 cm−1 and 137 cm−1 and these peaks are attributed to the libration modes.44 Interestingly, we did not clearly discern a peak at 155 cm−1 due to the Te–Te stretching mode previously reported.44 However, based on the evidence provided by the XRD, EDS, and Raman analyses, we conclude that the NC FeTe2 synthesized here belong to marcasite phase with orthorhombic crystal structure.
image file: c6ra06351a-f3.tif
Fig. 3 Raman spectroscopy measurement of as-synthesized NC FeSe2 and FeTe2 using 632.8 nm excitation from a HeNe laser.

UV-Vis-NIR spectroscopy

The absorbance spectra of as-synthesized NC FeSe2 dispersed in chloroform were taken using a Perkin Elmer Lambda 1050 spectrophotometer in the spectral region from 350 nm to 1500 nm. Although the theoretical value for the band gap of bulk FeSe2 material has been reported to be 1 eV,45 the reported experimental values vary more widely within the range of 1.03 to 1.24 eV.24,32,46,47 A typical absorbance spectrum of NC FeSe2 taken in chloroform as a solvent is given in the ESI, Fig. SF6(a). Qin et al. have reported an absorption peak of FeSe2 NCs at 1135 nm; however, they did not present a wide bandwidth measurement of absorption from UV to the NIR.33 Unlike the report of Qin et al., our optical absorption measurements of FeSe2 NCs carried out in hexane, trichloroethylene, and in thin film form did not exhibit a clear peak. Instead, we observed a largely featureless spectrum for which optical extinction increased gradually with wavelength through most of the visible and the near-infrared region. As stated earlier, our NC FeSe2 range in size from ∼150 nm to ∼1 μm, with irregularly shaped NCs. Yuan et al. have reported an increase in absorbance as wavelength increases in FeSe2 particles synthesized at 180 °C; their measured absorbance spectrum is almost identical for 220 °C growth temperature, and they show a broad optical absorption peak for FeSe2 NCs synthesized by a solvothermal method at higher temperatures (260 °C, 300 °C).24 We find non-zero absorption over the entire range we measured. Absorption in longer wavelength region is unexpectedly higher (SF6a), similar to the UV-Vis-NIR spectra reported by Yuan et al. for FeSe2 particles synthesized at 180 °C,24 which indicates the presence of a high density of mid-gap defect states.

Similarly, the absorbance spectra of NC FeTe2 dispersed in chloroform were measured in the wavelength range 350 nm to 2200 nm. However, we did not observe any absorption peaks in the selected wavelength range as shown in the ESI, Fig. SF6(b). The nature of this absorption remained unchanged with respect to solvents used to disperse the NCs and also for the NC thin film. The reported band gap values of this material fall in the range of 0.2 to 0.5 eV.27,48 Compared with iron pyrite FeS2, the FeSe2 and FeTe2 NCs incorporate chalcogenides with increasing atomic radii, and hence the outer electrons ionize more readily indicating they possess more metallic character than sulfur. The reason behind the increase in absorbance in the IR region might be due to defect-induced generation of large free carriers populations (electrons and holes)24 corresponding to semi-metallic behavior of the material.

Electronic properties

The electronic properties of the as-synthesized FeSe2 and FeTe2 films were studied using thermal probe, and four point probe measurements, and the film thicknesses were measured using a Dektak profilometer. To find the majority carrier type for each material, we employed the thermal probe measurement method, which utilizes a simple two-probe test in which one probe is heated to drive diffusion of free carriers. The basic experimental setup is quite similar to a system used to measure the Seebeck coefficient, though in the case of the thermal probe we are concerned only with the sign of the signal as a method to identify the majority carrier type. For thin films of both FeSe2 and FeTe2, the deflection in the voltmeter was opposite (negative) to that of the n-type silicon wafer (positive). Thus the majority charge carriers in as-synthesized FeSe2 and FeTe2 NCs films are holes, indicating that both of the marcasite NC iron dichalcogenide materials are p-type. Our typical as-synthesized FeSe2 film of average thickness ∼6.2 μm shows an average sheet resistance of ∼3.0 × 104 Ω □−1, and the resistivity is 19.3 Ω cm. Similarly, in our typical FeTe2 film with an average thickness of ∼5.3 μm shows a measured average sheet resistance value of ∼8.0 × 103 Ω □−1, corresponding to a resistivity of 4.1 Ω cm. The previously reported resistivity values for bulk FeSe2 material fall in the range of 0.26–1 Ω cm (ref. 27, 47 and 49) and for FeTe2 ∼ 0.015 Ω cm.50 Our measured values are significantly higher, which we explain based on the presence of long chain organic molecules like 1,2-hexanediol, and oleylamine, in as-synthesized FeSe2 and FeTe2 NCs. For comparison, the resistivity of as-synthesized iron pyrite FeS2 NCs (containing the tri-octylphosphine oxide (TOPO) molecules used in synthesis) for a typical film of average thickness 3.5 μm was ∼120 Ω cm, and after hydrazine treatment to remove TOPO surfactant molecules the resistivity decreased to ∼12 Ω cm.37 Hence, these as-synthesized marcasite FeSe2 and FeTe2 NCs films are more conductive than as-synthesized iron pyrite FeS2 NCs film. The comparison of the resistivity of these FeSe2 and FeTe2 thin films with FeS2 films is made based on the (i) similar synthesis procedure using same iron precursor, (ii) similar thin film preparation method, and (iii) similar films thicknesses even though the particle feature sizes of NC FeSe2 (or NC FeTe2) are larger than FeS2 NCs. Also, we note that in all cases the resistivity/conductivity of these FeSe2 and FeTe2 thin films does not reflect the intrinsic properties of the grains or crystals of these materials, but rather is expected to be dominated by the inter-grain transport. We have previously reported that FeS2 NCs can serve as an effective back contact for CdTe solar cells;5 however, we have as yet been unable to test these FeSe2, and FeTe2 marcasite NCs in a similar role. In principle, through improved surface control and film formation, these marcasite phase films which show improved conductivity may find a role in electronic materials. The reported narrow band gap energies for the marcasite Se and Te iron dichalcogenides do not suggest a role as a light-absorbing layer in highly efficient solar cells. While the decreased band gap energy suggests that any sustained photovoltage would be low, additional measurements are also needed to determine the band edge energetics, defect state distributions, and free carrier concentrations.

Conclusion

The NC iron dichalcogenides of FeSe2 and FeTe2 have been successfully synthesized by hot-injection of elemental Se and Te, respectively, into a bromine-anion Fe precursor solution. For both FeSe2 and FeTe2, the as-synthesized NC particles have orthorhombic crystal structure, and are in marcasite phase with flower and flake like shape. Based on Raman spectroscopy, NC FeSe2 shows active modes at 180 cm−1, 220 cm−1, and 256 cm−1 while NC FeTe2 shows modes at 119 cm−1 and 137 cm−1. The straightforward synthetic process described here enables additional discovery of NC FeSe2 and FeTe2 properties with applications in magnetic, electronic, and thermoelectric applications. These as-synthesized FeSe2 and FeTe2 films show p-type conductivity based on hot-probe measurement, and they exhibit relatively low conductivity due to the presence of long-chain hydrocarbon surfactant molecules. We measured unexpectedly weak and featureless optical absorption spectra. Although the reported narrow band gap of these materials indicate that they likely will not find success as absorber layers in photovoltaic devices, we are optimistic that the facile colloidal synthesis route to NC FeSe2 and FeTe2 materials demonstrated here will enable continued exploration of their properties and potential application in opto-electronic devices.

Acknowledgements

The authors acknowledge support from the National Science Foundation's Sustainable Energy Pathways Program under Grant CHE-1230246; the United States Air Force Research Laboratory, Space Vehicles Directorate, under Contract #FA9453-11-C-0253; and the University of Toledo through startup funds and the Professional Science Masters program in Photovoltaics.

References

  1. H. Zhou, W.-C. Hsu, H.-S. Duan, B. Bob, W. Yang, T.-B. Song, C.-J. Hsu and Y. Yang, Energy Environ. Sci., 2013, 6, 2822–2838 CAS .
  2. D. V. Talapin, J.-S. Lee, M. V. Kovalenko and E. V. Shevchenko, Chem. Rev., 2009, 110, 389–458 CrossRef PubMed .
  3. C.-H. M. Chuang, P. R. Brown, V. Bulović and M. G. Bawendi, Nat. Mater., 2014, 13, 796–801 CrossRef CAS PubMed .
  4. V. Sukhovatkin, S. Musikhin, I. Gorelikov, S. Cauchi, L. Bakueva, E. Kumacheva and E. Sargent, Opt. Lett., 2005, 30, 171–173 CrossRef CAS PubMed .
  5. K. P. Bhandari, P. Koirala, N. R. Paudel, R. R. Khanal, A. B. Phillips, Y. Yan, R. W. Collins, M. J. Heben and R. J. Ellingson, Sol. Energy Mater. Sol. Cells, 2015, 140, 108–114 CrossRef CAS .
  6. N. Kholmicheva, P. Moroz, U. Rijal, E. Bastola, P. Uprety, G. Liyanage, A. Razgoniaev, A. D. Ostrowski and M. Zamkov, ACS Nano, 2014, 8, 12549–12559 CrossRef CAS PubMed .
  7. I. J. Kramer and E. H. Sargent, Chem. Rev., 2014, 114, 863–882 CrossRef CAS PubMed .
  8. G.-H. Kim, F. P. García de Arquer, Y. J. Yoon, X. Lan, M. Liu, O. Voznyy, Z. Yang, F. Fan, A. H. Ip, P. Kanjanaboos, S. Hoogland, J. Y. Kim and E. H. Sargent, Nano Lett., 2015, 15, 7691–7696 CrossRef CAS PubMed .
  9. P. K. Jain, X. Huang, I. H. El-Sayed and M. A. El-Sayed, Acc. Chem. Res., 2008, 41, 1578–1586 CrossRef CAS PubMed .
  10. P. Moroz, G. Liyanage, N. N. Kholmicheva, S. Yakunin, U. Rijal, P. Uprety, E. Bastola, B. Mellott, K. Subedi, L. Sun, M. V. Kovalenko and M. Zamkov, Chem. Mater., 2014, 26, 4256–4264 CrossRef CAS .
  11. M. V. Kovalenko, R. D. Schaller, D. Jarzab, M. A. Loi and D. V. Talapin, J. Am. Chem. Soc., 2012, 134, 2457–2460 CrossRef CAS PubMed .
  12. Z. Han, F. Qiu, R. Eisenberg, P. L. Holland and T. D. Krauss, Science, 2012, 338, 1321–1324 CrossRef CAS PubMed .
  13. D. B. Khadka and J. Kim, J. Phys. Chem. C, 2014, 118, 14227–14237 CAS .
  14. T. K. Todorov, K. B. Reuter and D. B. Mitzi, Adv. Mater., 2010, 22, E150–E159 CrossRef .
  15. P. S. Vasekar, T. P. Dhakal, Thin Film Solar Cells Using Earth-Abundant Materials, Solar cells – research and application perspectives, ed. A. Morales-Acevedo, InTech, 2013, p. 386, ISBN: 978-953-51-1003-3,  DOI:10.5772/3418 .
  16. P. R. Bonneau, R. F. Jarvis and R. B. Kaner, Nature, 1991, 349, 510–512 CrossRef CAS .
  17. T. A. Bither, C. T. Prewitt, J. L. Gillson, P. E. Bierstedt, R. B. Flippen and H. S. Young, Solid State Commun., 1966, 4, 533–535 CrossRef CAS .
  18. J. G. Brennan, T. Siegrist, Y. U. Kwon, S. M. Stuczynski and M. L. Steigerwald, J. Am. Chem. Soc., 1992, 114, 10334–10338 CrossRef CAS .
  19. J.-P. Ge and Y.-D. Li, J. Mater. Chem., 2003, 13, 911–915 RSC .
  20. M. Bochmann, Chem. Vap. Deposition, 1996, 2, 85–96 CrossRef CAS .
  21. A. Sobhani and M. Salavati-Niasari, J. Alloys Compd., 2015, 625, 26–33 CrossRef CAS .
  22. W. Shi, X. Zhang, G. Che, W. Fan and C. Liu, Chem. Eng. J., 2013, 215–216, 508–516 CrossRef CAS .
  23. A. Liu, X. Chen, Z. Zhang, Y. Jiang and C. Shi, Solid State Commun., 2006, 138, 538–541 CrossRef CAS .
  24. B. Yuan, X. Hou, Y. Han, W. Luan and S.-T. Tu, New J. Chem., 2012, 36, 2101–2105 RSC .
  25. M. Limpinsel, N. Farhi, N. Berry, J. Lindemuth, C. L. Perkins, Q. Lin and M. Law, Energy Environ. Sci., 2014, 7, 1974–1989 CAS .
  26. M. Cabán-Acevedo, D. Liang, K. S. Chew, J. P. DeGrave, N. S. Kaiser and S. Jin, ACS Nano, 2013, 7, 1731–1739 CrossRef PubMed .
  27. T. Harada, J. Phys. Soc. Jpn., 1998, 67, 1352–1358 CrossRef CAS .
  28. L. Chun-Rong, Y.-J. Siao, S.-Z. Lu and G. Chie, IEEE Trans. Magn., 2009, 45, 4275–4278 CrossRef .
  29. V. K. Gudelli, V. Kanchana, G. Vaitheeswaran, M. C. Valsakumar and S. D. Mahanti, RSC Adv., 2014, 4, 9424–9431 RSC .
  30. F. Liu, J. Zhu, L. Hu, B. Zhang, J. Yao, M. K. Nazeeruddin, M. Gratzel and S. Dai, J. Mater. Chem. A, 2015, 3, 6315–6323 CAS .
  31. W. Wang, X. Pan, W. Liu, B. Zhang, H. Chen, X. Fang, J. Yao and S. Dai, Chem. Commun., 2014, 50, 2618–2620 RSC .
  32. J. Li, Z. Jin, T. Liu, J. Wang, X. Zheng and J. Lai, CrystEngComm, 2014, 16, 6819–6822 RSC .
  33. Z. Qin, B. C. Yang, S. R. Zhang, K. Li, S. Q. Yan, X. W. Lv and Z. J. Li, Vacuum, 2015, 111, 157–159 CrossRef CAS .
  34. W. Chengrong, B. Yongxiao, D. Aiying and B. Yan, Nanotechnology, 2016, 27, 165702 CrossRef PubMed .
  35. K. D. Oyler, X. Ke, I. T. Sines, P. Schiffer and R. E. Schaak, Chem. Mater., 2009, 21, 3655–3661 CrossRef CAS .
  36. W. Zhang, Y. Cheng, J. Zhan, W. Yu, L. Yang, L. Chen and Y. Qian, Mater. Sci. Eng., B, 2001, 79, 244–246 CrossRef .
  37. K. P. Bhandari, P. J. Roland, T. Kinner, Y. Cao, H. Choi, S. Jeong and R. J. Ellingson, J. Mater. Chem. A, 2015, 3, 6853–6861 CAS .
  38. T. Kinner, K. P. Bhandari, E. Bastola, B. M. Monahan, N. O. Haugen, P. J. Roland, T. P. Bigioni and R. J. Ellingson, J. Phys. Chem. C, 2016, 120, 5706–5713 CAS .
  39. J. Pickardt, B. Reuter, E. Riedel and J. Söchtig, J. Solid State Chem., 1975, 15, 366–368 CrossRef CAS .
  40. S. Tengner, J. Inorg. Gen. Chem., 1938, 239, 126–132 CAS .
  41. S. Mourdikoudis and L. M. Liz-Marzán, Chem. Mater., 2013, 25, 1465–1476 CrossRef CAS .
  42. J. Zimdars and M. Bredol, New J. Chem., 2016, 40, 1137–1142 RSC .
  43. W. Zhang, Z. Yang, J. Zhan, L. Yang, W. Yu, G. Zhou and Y. Qian, Mater. Lett., 2001, 47, 367–370 CrossRef CAS .
  44. H. D. Lutz and B. Müller, Phys. Chem. Miner., 1991, 18, 265–268 CrossRef CAS .
  45. B. Ganga, C. Ganeshraj, A. G. Krishna and P. Santhosh, arXiv preprint arXiv:1303.1381, 2013.
  46. B. Ouertani, J. Ouerfelli, M. Saadoun, M. Zribi, M. B. Rabha, B. Bessaïs and H. Ezzaouia, Thin Solid Films, 2006, 511–512, 457–462 CrossRef CAS .
  47. N. Hamdadou, A. Khelil, M. Morsli and J. C. Bernàde, Vacuum, 2005, 77, 151–156 CrossRef CAS .
  48. D. Sharma, B. Joshi, K. Patel, V. Ganeshan and Y. Sharma, International Journal of Recent Research and Review, 2013, 6, 32–49 Search PubMed .
  49. Non-Tetrahedrally Bonded Binary Compounds II, ed. O. Madelung, U. Rössler and M. Schulz, Springer Berlin Heidelberg, Berlin, Heidelberg, 2000, pp. 1–3,  DOI:10.1007/10681735_558 .
  50. Non-Tetrahedrally Bonded Binary Compounds II, ed. O. Madelung, U. Rössler and M. Schulz, Springer Berlin Heidelberg, Berlin, Heidelberg, 2000, pp. 1–3,  DOI:10.1007/10681735_560 .

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra06351a

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.