1 s2.0 S0921510718300539 Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Materials Science & Engineering B 236–237 (2018) 104–124

Contents lists available at ScienceDirect

Materials Science & Engineering B


journal homepage: www.elsevier.com/locate/mseb

Pyrite FeS2 nanostructures: Synthesis, properties and applications T


a,⁎ a a a b,⁎ c
Haiying Qin , Junkang Jia , Longxia Lin , Hualiang Ni , Mudan Wang , Liang Meng
a
College of Materials and Environmental Engineering, Hangzhou Dianzi University, Hangzhou 310018, China
b
Shanghai Huali Microelectronics Corporation, Shanghai 201203, China
c
School of Materials Science and Engineering, Zhejiang University, Hangzhou 310027, China

A R T I C LE I N FO A B S T R A C T

Keywords: Nanostructured metal sulfides have been widely studied due to their unique electrical, chemical, magnetic and
FeS2 optical properties and the promising application prospects. Pyrite iron disulfide (FeS2) receives high attention
Nanostructure because of its suitable band gap, exceptionally high optical absorption coefficient in the visible region and the
Synthesis advantages of earth abundance, nontoxicity and low fabrication cost. It exhibits promising applications in
Photovoltaic
photovoltaics, rechargeable batteries, dye-sensitized solar cells, photodetectors, photocapacitors and photo-
Rechargeable batteries
catalysis. This paper provides a comprehensive review of the FeS2 nanostructures, including the preparation
methods, properties and applications. We begin from a short background of FeS2, describing the structure,
fundamental physical, chemical and electronic properties. Then a detailed investigation of FeS2 nanostructures
with different shapes and the corresponding methods are described. Finally, the application prospects of FeS2
nanostructures in plentiful novel devices are introduced.

1. Introduction In this article, we provide a comprehensive review of the FeS2 na-


nostructures, including the preparation methods, properties and appli-
Nanostructured metal sulfides show great potential for the devel- cations. We begin from a short background of FeS2, describing the
opment of various novel and smart devices. These metal sulfides have structure, fundamental physical, chemical and electronic properties,
been widely studied because of their unique electrical, optical, chemical and make a detailed investigation of FeS2 nanostructures with various
and magnetic properties and the promising applications in biological shapes and corresponding methods. Abundant structured FeS2, such as
labelling, photovoltaic solar cells and medical diagnostics [1]. nanoparticles, nanowires, nanorods, nanotubes, nanoplates, nanosheets
Pyrite iron disulfide (FeS2) is a typical representative and a popular and microspherolites were successfully prepared. Effects of synthetic
topic among the research on metal sulfides due to its unique properties conditions on the properties of nanostructures like physical, chemical
such as suitable band gap, exceptionally high optical absorption and, and photoelectrical properties are further discussed. Finally, we in-
most importantly, the advantages of earth-abundance, nontoxicity and troduce the applications the FeS2 structures in plentiful functional and
low cost [2]. Since pyrite FeS2 was firstly explored as a photovoltaic novel devices, including quantum-dot-sensitized solar cells (QDSSCs),
semiconductor in 1983, there has been continuous interest in that rechargeable batteries, dye-sensitized solar cells (DSSCs), photo-
material recently. It was employed as a high-performance electrode detectors, photocapacitors and photocatalysis.
material in rechargeable batteries [3]. The Shockley-Queisser limit of
FeS2 is around 31% for an ideal band gap of 1.34 eV [2]. Therefore 2. Fundamental properties of FeS2
pyrite is well positioned to address the challenges in solar harvesting
and the economic difficulty as the active material in photovoltaic FeS2, one of the transition-metal sulphides, has the most elemental
junctions. Moreover, extremely high optical absorption allows FeS2 film storage in the earth. FeS2 has two common crystal structures, cubic-
harvest 90% of incident light with a thickness of ∼40 nm [4]. It could system pyrite and orthorhombic -system marcasite. Fig. 1(a) and (b)
produce efficient carrier extraction and create novel heterojunction show the schematic view of pyrite and marcasite structured FeS2. The
device architectures such as thin absorber cells, which can be con- lattice parameters of pyrite are a = b = c = 0.54 nm, the bond length of
sidered “semiconductor-sensitized” in analogy to dye-sensitized cells Fe-S and S-S are 0.23 nm and 0.21 nm, respectively. The structure of
[5]. cubic FeS2 is similar to that of NaCl. Namely, Fe2+ and S2−2 displace


Corresponding authors.
E-mail addresses: [email protected] (H. Qin), [email protected] (M. Wang).

https://doi.org/10.1016/j.mseb.2018.11.003
Received 29 August 2017; Received in revised form 24 December 2017; Accepted 14 November 2018
Available online 27 November 2018
0921-5107/ © 2018 Elsevier B.V. All rights reserved.
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 1. Crystal structures of FeS2 (a) Pyrite and (b) marcasite, (c) calculated density of states of pyrite. The inset shows a zoomed-in region near the Fermi level [7].
Reprinted with permission from [7]. Copyright (2016) American Chemical Society.

the lattice sites of Na+ and Cl-, respectively. One Fe atom and six S2 onto molybdenum-coated glass substrate and then it was heated to
form octahedral coordination, one S atom and surrounding three Fe 370 °C over the course of 10 min. The obtained ink films were then
atoms and one S atom form tetrahedral coordination [6]. Pyrite FeS2 is annealed at 390 °C for 12 h with 1 atm of flowing H2S gas to produce
a semiconductor material and its band structure is given in Fig. 1(c) [7]. pyrite FeS2 nanoparticles (Fig. 2(b)).
There is a large density of valence band states consisting of mainly Fe d
character. The onset of optical absorption implied by the large density 3.1.2. Electrodeposition method
of states in the conduction band is around 1 eV above the Fermi level. There are few reports of FeS2 films fabricated by electrodeposition
The bandgap of marcasite FeS2 is only 0.34 eV due to its low structure method. Besides, the FeS2 is hardly obtained by one-step electro-
symmetry. Therefore, marcasite FeS2 is not suitable for application as deposition. FeS was formed via electrodeposition, and it was converted
photovoltaic materials. Almost all researches of FeS2-related systems into pyrite FeS2 after thermal sulfidation. Nakamura et al. [10] elec-
are built on pyrite structure, and this review also is based on the fact. trodeposited FeS film on titanium sheets in an aqueous solution of
FeSO4 and Na2S2O3 with Fe/S molar ratio 1:2. Dilute H2SO4 was used to
3. Synthesis of FeS2 nanostructures control the pH value of the solution system and the electrodeposition
time was 60 min. The as-synthesized FeS film was sulfurized at 500 °C
Abundant nanostructured FeS2, such as nanoparticles, nanowires, using N2 as the carrier gas, and a chemical conversion from FeS to FeS2
nanofilms and microspherolites have been synthesized under specific was finished after 20 min. Moreover, Gomes et al. [11] and Dong et al.
experimental conditions. According to the difference of their shapes, [12] also obtained FeS2 films by sulfurizing FeS films prepared via si-
FeS2 nanostructures are divided into 0D, 1D, 2D and 3D. The detailed milar electrodeposition method and studied the effect of annealing
investigation of FeS2 nanostructures with various shapes and corre- temperature on the structure of FeS2. When the annealing temperature
sponding synthesis methods is summarized in this section. was lower than 400 °C, the crystalline phases of sulfurization product
contain both pyrite FeS2 and marcasite FeS2. When the annealing
3.1. 0D nanostructures temperature was higher than 500 °C, single-phase pyrite FeS2 films
were formed. Meng et al. [13] obtained precursor Fe3O4 films using
3.1.1. Sol-gel method electrodeposition in combination with low-temperature heat treatment
In past decades, Meng et al. [8] has made various researches on and changed it into FeS2 films by sulfidation process. The effects of
pyrite FeS2 films synthesized by the sol–gel method. In the typical sulfidation time and sulfur vapor pressure on the structure of FeS2 films
synthesis, Fe2O3 films were grown on fluorine-doped tin oxide (FTO) were studied. It was found that uniform and dense FeS2 nanoparticles
substrate by sol–gel method; then it was transformed into FeS2 films by were formed at 40 kPa sulfur vapor pressure for 10 h.
subsequent sulfidation treatment. The specific preparation steps were
described as follow: (1) Fe(NO3)3·9H2O was dissolved in a mixture 3.1.3. Chemical vapor deposition (CVD)/metal organic CVD (MOCVD)
containing 2-methoxythanol and acetyl acetone; (2) the solution was CVD method is a direct synthesis, in which FeS2 can be formed by
continuously stirred for forming homogeneous sol solution; (3) pre- one step without subsequent sulfidation. It mainly takes advantage of
cleaned FTO substrates were immersed into the gel solution and were the vapor chemical reactions between selected iron and sulfur sources
coated a gel film and then were dried; (4) the dipping and drying on various substrates, such as glass, silicon and molybdenum. The dif-
process was repeated for several cycles; (5) the gel films were annealed ference for MOCVD is using metal–organic compound as reacting
for forming Fe2O3 films; (6) Fe2O3 films and sublimed sulfur powder source. Early MOCVD commonly adopted the combination of di-
were sealed in quartz ampoules and annealed to obtain FeS2 nano- tertbutyl disulfide with iron (III) acetylacetonate precursors [14]. Ac-
particles (Fig. 2(a)). cording to the specific combination of precursors, the reaction tem-
Seefeld et al. [9] used a method similar to sol–gel to form Fe(acac)3 perature varied from 130 °C to 600 °C. Recently developed CVD
ink (acac = acetylacetonate), which consisted of Fe(III) acetylacetonate preparation methods employed the reaction between FeCl3 and thioa-
and elemental sulfur dissolved in pyridine. The ink was spin-coated cetamide at 400–550 °C [15] or the reaction between iron (III)

105
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 2. SEM images of the FeS2 nanoparticles by sol–gel method (a) in ref. [8] and (b) in ref. [9]. Reprinted with permission from [8]. Copyright (2010) Elsevier.
Reprinted with permission from [9]. Copyright (2013) American Chemical Society.

Fig. 3. (a) Schematic diagram of CVD system using FeCl3 and ditert butyl disulfide as precursors; (b) SEM image of FeS2 film grown on CoS2 substrate; (c) Raman
spectra of FeS2 films grown on quartz, borosilicate and CoS2 substrates, respectively [17]. Reprinted with permission from [17]. Copyright (2015) American Chemical
Society.

acetylacetonate and ditertbutyl disulfide at 300 °C [16] to fabricate tin sulfide and iron disulfide [18]. Ouertani et al. [19] utilized 0.03 M
pyrite FeS2 films. Marcasite phase or other iron sulfide impurities were FeCl3·6H2O as a single-source precursor and sprayed it onto glass sub-
commonly observed in that synthesis. Recently, Samad et al. [17] pro- strates heated at 350 °C. The obtained product was amorphous iron
duced phase-pure iron pyrite films via CVD using FeCl3 and ditert butyl oxide, and it was further sulfurized at 450 °C for 6 h to produce pyrite
disulfide as precursors and cattierite CoS2 as the substrate. Fig. 3(a) and FeS2. High-density of defects existed in the as-prepared FeS2 films,
(b) shows the schematic diagram of CVD system and the scanning which generated a narrow band of localized states in the forbidden
electron microscopy (SEM) image of the FeS2 film, respectively. They band. Yamamoto et al. [20] took FeSO4 and (NH4)2Sx as Fe source and S
made a comparison experiment using quartz, borosilicate glass and source, respectively. The two solutions were sprayed alternately onto
cattiertite CoS2 as the substrates and tested their corresponding Raman substrates heated at 120 °C. The as-prepared films usually included
spectra, respectively. Hematite and marcasite phases were observed in other phases, such as FeS and marcasite. As a result, the films required
the sample using quartz as the substrate, and marcasite phase was also to be annealed in the H2S atmosphere for 30 min. Experiments found an
observed in the sample using borosilicate as the substrate (Fig. 3(c)). ideal sulfidation temperature was around 500 °C and sulfidation at this
High-purity pyrite FeS2 was deposited on cattierite CoS2 substrate. temperature could synthesize single-phase FeS2 films. The obtained
Results indicated that substrate should play an important role in the FeS2 was a p-type semiconductor and possessed high carrier con-
purity of FeS2 films synthesized via CVD method. centration and Hall mobility. Recently, Shukla et al. [21] chose
FeCl3·6H2O and NH2CSNH2 as Fe and S precursors, respectively. A
3.1.4. Spray pyrolysis 50 mL mixture of the two precursors was prepared with Fe/S molar
Spray pyrolysis technique provides the advantages of simple and ratio of 1:6. It was sprayed onto FTO substrates mounted on a hot plate
quick processing and non-vacuum technology. It has been successfully at 330 °C. The film was sulfurized at 400 °C for 30 min in Ar atmosphere
used for the fabrication of some photovoltaic materials including to obtain the FeS2 film. The FeS2/FTO configuration was directly used
copper indium sulfide, copper indium gallium sulfide and copper zinc as a counter electrode in dye-sensitized solar cells. Fig. 4 shows the

106
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 4. (a) Schematic illustration of the spray devices and sulfidation process, (b) SEM image of the obtained FeS2 nanoparticles [21]. Reprinted with permission from
[21]. Copyright (2014) American Chemical Society.

Table 1
Reports for FeS2 synthesis via hydrothermal and solvothermal methods.
Method Fe source S source Surfactant Reaction temperature (oC) Reaction time (h) Size (nm) Shape Refs.

Hydro-thermal FeCl3 Diethyl dithiophosphate ammonium salt CTAB 200 4–48 ∼150 cube [29]
FeCl2 S powder Gelatin 200 48 ∼20 particle [30]
FeSO4 (NH4)2S2O3 — 500 9 ∼50 particle [31]
FeCl3 Na2S S powder PEG-400 120 48 10–100 particle [32]
FeCl2 S powder PVA PVP 180 12 130 cube [33]
FeCl2 S powder PVP 200 24 ∼1000 cube [34]
Fe3O4 Na2S2O3 S powder 1-octylamine 220 12 15–35 particle [35]

Solvo-thermal FeF3 S powder DMF 190 3–22 5.5 particle [36]


FeCl2 S powder Gelatin 200 15–23 particle [28]
FeCl2 S PVP 180 14–18 particle [37]
FeSO4 Na2S2O3 — 200 20–500 particle [38]
FeCl2 S OLA 180 80–120 cube [39]
Fe2O3 S CTAB 290 ∼200 cube [40]

a
CTAB: hexadecyltrimethylammonium bromide; PEG: polyethylene glycol; PVA: polyvinyl Alcohol; PVP: polyvinyl Pyrrolidone; DMF: dimethyl Formamide; OLA:
oleylammine.

schematic illustration of the spray device and sulfidation process and sulfidation. Their previous research revealed that Fe3S4 was a natural
the obtained FeS2 nanoparticles. intermediate on the pathway from polysulfide to pyrite [26]. Therefore,
sulfidation of Fe3S4 instead of Fe could effectively suppress the for-
mation of FeS intermediate layer as well as FeS impurity phase in FeS2
3.1.5. Magnetron sputtering films.
In 1992, Willeke et al. [22] prepared pyrite FeS2 films by sulfur-
assisted magnetron sputtering in one step. The synthetic FeS2 powder
was used as a sputtering target. There was a small quantity of FeS phase 3.1.6. Hydrothermal/solvothermal method
in the target material. A certain amount of S vapor was loaded into the Hydrothermal and solvothermal methods are two common pre-
reaction chamber to ensure that the film composition can satisfy the- paration techniques among wet-chemical methods. The difference be-
oretical stoichiometry of FeS2 (Fe:S = 1:2). A series of experiments tween the two methods is the solution system used for the reaction. The
found that pyrite phase appeared at high temperature and films pre- hydrothermal method uses an aqueous solution and the solvothermal
pared at low temperature commonly contained marcasite phase. One- method uses an organic solvent. Many studies about the preparation of
step synthesis of FeS2 films by magnetron sputtering has rarely been FeS2 nanocrystals using hydrothermal/solvothermal methods have
reported. Many published reports firstly prepared precursor films by the been reported. A review of the shape controllable synthesis of pyrite
magnetron and then transformed it into FeS2 films by sulfidation. Meng crystals through various wet-chemical methods was provided by Xian
and co-workers [23] prepared FeS2 films on different substrates by et al. [27]. Various surfactants have been widely used in those methods,
sulfidation annealing of magnetron sputtered iron films at 400 °C for such as hexadecyltrimethylammonium bromide and gelatin. Surfactant
20 h. They found sufficient formation and growth of iron grains can plays an important role in the stabilization and size uniformity of FeS2
improve the crystallinity and continuity of FeS2 films, which was im- nanocrystals, and it also can effectively prevent the aggregation of
portant for the electrical properties of FeS2 films. Zhang et al. [24] also nanocrystals into large microparticles [28]. Table 1 summarizes some
prepared FeS2 films by thermally sulfurizing iron films obtained via representative publications of hydrothermal and solvothermal methods,
magnetron sputtering. Liu et al. [25] prepared precursor Fe3S4 films by which hold promise for synthesizing FeS2 nanocrystals. These results
magnetron sputtering and conversed it into FeS2 films by thermal indicate that all these factors: Precursor characteristics, the molar ratio

107
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

of Fe/S, reaction temperature and time, solvents and surfactants have injection route, making a comparison of several important parameters.
significant influence on controlling the morphology and phase purity of
synthesized FeS2 products.
3.1.8. Ultrasonic irradiation
Ultrasound method is an efficient way for the synthesis of pyrite
3.1.7. Hot-injection method nanoparticles. Khhabaz and Entezari [54] successfully synthesized the
Preparation of FeS2 nanocrystal inks has been a hot topic in recent pyrite nanoparticles by utilizing ultrasonic irradiation in 10 min, 70 °C
years due to applications in photovoltaics. Hot-injection method is the and 70% electrical power. The pyrite nanoparticles had an average size
most efficient technique to have successfully synthesized FeS2 nano- of 28.8 nm and contained both cubic and marcasite structures.
crystal inks. Hot injection is a method that injects S precursor solution
into Fe precursor solution under high temperature and protection at-
mosphere or injects one/all precursors into a pre-heated flask under 3.2. 1D nanostructures
water-free and air-free environment. Puthussery et al. [41] synthesized
phase-pure, single-crystalline and well dispersed colloidal FeS2 nano- 3.2.1. Solvothermal method
crystal inks by hot-injection route. A mixture of FeCl2·4H2O and octa- Kar and Chaudhuri [55] synthesized 1D FeS2 nanostructures using
decylamine was prepared in a three-neck flask. A mixture consisting of solvothermal process, and could well control the morphologies of FeS2
sulfur and diphenyl ether were used as injection solution. Under Ar nanostructures by changing molar concentrations of precursors, the
atmosphere, the injection solution was added into Fe precursor by iron source and reaction temperatures. They used FeSO4·7H2O or FeCl3
syringe, and the reaction mixture was kept at 220 °C for 3 h. When the or Fe(NO3)3·9H2O as iron sources, thiourea (NH2CSNH2) as a sulfur
reaction stage finished, the reaction system was cooled using a water source, ethylenediamine (NH2CH2CH2NH2) as a solvent, mixed these
bath. Anhydrous chloroform was injected into the solution once the reagents and kept them reacted at the different temperatures for 12 h.
temperature decreased to 100 °C to prevent the octadecylamine from Fig. 6 displays various 1D FeS2 nanostructures obtained by controlling
solidifying. Well-dispersed FeS2 nanocrystal ink was obtained after re- these experimental parameters. When using FeSO4·7H2O as iron source,
peated washing and centrifugation processes. Corresponding FeS2 films short nanorods with diameter ranging from 40 to 100 nm and length up
were fabricated by dipping substrates repeatedly in the nanocrystal ink to 500 nm were obtained; when using Fe(NO3)3·9H2O as iron source,
using a dip coater and sintering at 540 °C for 4 h. Fig. 5 shows the SEM uniform nanowires with a diameter of 40–60 nm were produced; when
and transmission electron microscopy (TEM) characterizations of the FeCl3 was used as iron source, both nanowire (greater than90%) and
FeS2 nanocrystal inks. The as-synthesized FeS2 nanocrystals have a micro-rod nanostructures were obtained. The morphology difference
uniform size distribution and show oblate and spheroidal shapes with can be ascribed to the solubility difference of the iron salts in the sol-
diameters of 5–20 nm (Fig. 5(a) and (b)). Lattice-resolved TEM images vent and the displacement of the corresponding anions [56]. Research
(Fig. 5(c)) reveals that the nanocrystals have high crystallinity with no on the reaction temperature parameter indicates that the dimension-
sign of an oxide coating or amorphous overlayer. The electron dif- ality of FeS2 nanostructures increases from 1D to 2D with the rise of
fraction results indicate that the nanocrystals are phase-pure pyrite FeS2 reaction temperature. The effects of molar concentration of precursors
without impurity phases, such as marcasite, greigite, pyrrhotite are similar to that of temperature variation on the shape of FeS2 na-
(Fig. 5(d)). Fig. 5(e) shows plane-view and cross-section SEM images of nostructures. Bai et al. [57] used FeCl2·4H2O dissolved in dimethyl
a 2 μm thick FeS2 nanocrystal film deposited on glass substrate. Sin- sulfoxide as an iron source, Na2S2O3·5H2O as a sulfur source, thiogly-
tering treatment results in significant grain growth, surface roughness colic acid (TGA) and ethylenediamine (EDA) as the costabilizer. When
and formation of voids. There are many relevant reports about FeS2 the molar ratio of [Fe2+]/[TGA] and [EDA]/[TGA] were kept as 1:3
nanocrystal ink by the similar hot-injection approach. Table 2 sum- and 2:1, respectively, FeS2 nanorods and nanowires were produced
marizes some typical reports of FeS2 nanocrystal ink using the hot- after reaction at 139 °C for 2–12 h.

Fig. 5. TEM and SEM characterization of the colloidal FeS2 nanocrystals and corresponding films [41]. (a) Low, (b) middle and (c) high magnification TEM images,
(d) electron diffraction patterns, (e) SEM image. Reprinted with permission from [41]. Copyright (2011) American Chemical Society.

108
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Table 2
Reports for FeS2 nanocrystal inks synthesized via hot-injection method.
Fe precursor solution S precursor solution Fe/S molar ratio T (oC) Size (nm) Shape Refs.

Fe source Solvent S source Solvent

FeCl2 OLA TOPO S OLA 1:6 220 60–200 Cubes [42]


FeCl2 OLA S OLA 1:6 220 150 Cubes dendrites [43]
10
FeCl2 OLATOPO S OLA 1:6 220 ∼70 Sphere [44]
FeCl2 DMSO Na2S2O3 DI water 1:5.2 139 2–9 Sphere [45]
FeCl2 ODA S DPE 1:3.8 220 ∼27 Sphere [46]
TOPO
FeCl2 ODA S DPE 1:6 220 5–20 Sphere [41]
Oval
FeCl2 ODA S DPE 1:8 120–220 40–80 Cubes [47,48]
FeCl2 HAD S HAD 1:6 200–250 30–50 Cubes [49]
OLA OLA
FeCl2 ODE S OLA 1:12 240 12–18 Sphere [50]
FeCl3 OLA S OLA 1:6 220 80 Cubes [51]
FeCl2 OLA S OLA 1:7.5 230 60 Cubes [52]
FeCl2 ODA S TOA 1:6 220 50–120 Cubes [53]
TOA
FeCl2 ODA S OLA 1:6 220 20–45 Sphere [53]
TOA

T: temperature; OLA: Oleylamine; TOPO: trioctylphosphineoxide; DMSO: dimethyl sulfoxide; DI water: deionized water; ODA: octadecylamine; DPE: diphenyl ether;
HAD: hexadecylamine; ODE: 1-octadecence; TOA: trioctylamine.

Fig. 6. SEM images of various 1D FeS2 nanostructures by controlling iron sources, the molar concentration of precursors and reaction temperature [55]. (a-c)
FeSO4·7H2O, FeCl3 and Fe(NO3)3·9H2O, respectively; (d-g) at 150 °C, 180 °C, 210 °C and 230 °C, respectively; (h, i) half and double of the precursors, respectively.
Reprinted with permission from [55]. Copyright (2004) Elsevier.

3.2.2. Direct thermal sulfidation temperature and sulfur vapor pressure. Series of research indicates that
Cabán-Acevedo et al. [58] prepared FeS2 nanorods via the sulfida- intermediate reaction temperature of 425 °C, appropriate sulfur super-
tion reaction between FeCl2 or FeBr2 precursor with flowing sulfur at- saturation and high pressure of sulfur vapor play key role in the
mosphere. Apart from nanorods, FeS2 nanobelts and nanoplates were synthesis of FeS2 nanostructures. Cabán-Acevedo and other co-workers
also obtained by changing the iron source and controlling the reaction [59] also reported the growth of vertically oriented single-crystalline

109
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 7. SEM and high-resolution TEM (HRTEM) characterizations for the single-
crystalline FeS2 nanowires prepared via sulfidation of steel foil [59]. (a) SEM
image; (b, c) HRTEM image and corresponding FFT of an individual nanowire
oriented along [95] ZA. Reprinted with permission from [59]. Copyright (2012)
American Chemical Society.

FeS2 nanowires via thermal sulfidation of steel foil, as shown in Fig. 7.


The process was completed in a home-built reactor which was con-
structed by a fused silica tube and pressure as well as argon gas flow
control devices. Pyrite FeS2 nanowires were obtained by sulfurizing
low-carbon steel foil at 350 °C for 2 h.

3.2.3. Template-directed synthesis


Template-directed synthesis has been extensively used in the con-
struction of abundant nanostructures, especially 1D nanostructures. The
most widely used and commercially available templates are anodic Fig. 8. SEM and TEM images for various 1D FeS2 nanostructures synthesized
aluminum oxide (AAO) membrane. Wan et al. [60] firstly electro- using AAO templates [60,61]. (a, b) Fe nanowires; (c, d) FeS2 nanowires; (e)
deposited Fe nanowires in the pores of AAO and then transformed it FeS2 nanowires; (f) FeS2 nanotubes. Reprinted with permission from [60,61].
into ordered FeS2 nanowire arrays after sulfidation at 450 °C for 8 h. Copyright (2005) Elsevier and (2014) Springer Science.
The obtained FeS2 nanowire arrays show high packing densities.
Fig. 8(a) and (b) show the precursor Fe nanowires; Fig. 8(c) and (d) are Fe3 + + 3H2 O→ Fe(OH)3 + 3H+ (1)
the corresponding SEM images of the FeS2 nanowires. It can be seen
that FeS2 nanowires keep similar shape and size to Fe nanowires, which ZnO + 2H+ → Zn2 + + H2 O (2)
can be controlled by changing the structure of AAO template. Li et al.
[61] fabricated FeS2 nanowire and nanotube arrays via AAO template- 4Fe(OH)3 + 11 S→ 4FeS2 + 6H2 O+ 3SO2 (3)
directed synthesis in combination with a sol–gel method. AAO template 3+ +
The hydrolysis of Fe produces Fe(OH)3 colloid and H during
was immersed into sol solution for 10 min and followed by annealing at
immersion. Fe(OH)3 covers onto ZnO surface. Simultaneously, the ZnO
300 °C for 10 min. Different 1D nanostructured precursors Fe2O3 can be
nanorod template is etched by H+, which preferentially proceeds in the
obtained by controlling the sol immersion and annealing times. FeS2
central part and along the axis. As a result, ZnO template is removed out
nanostructures were obtained by sulfurizing Fe2O3 at 500 °C for 1 h and
and finally pure FeS2 is obtained. The difference for the synthesis of the
subsequent template etching in 1 M NaOH. Nanowire structure was
FeS2 nanotube is the different solution treatment, which produces
obtained by repeating immersion and annealing process for several
precursor Fe2O3 nanotube array. Further sulfidation will form FeS2
times, nanotube structure was obtained by repeating the processes for
nanotube array. Fig. 10 shows the SEM images of the as-synthesized
only one time. The corresponding SEM images of FeS2 nanowires and
FeS2 nanorod and FeS2 nanotube arrays.
FeS2 nanotubes are shown in Fig. 8(e) and (f).
Besides the template mentioned above, the Fe-compound precursor
Recently, ZnO nanostructures have also been used as function
could also play as self-template. Morrish et al. [65] synthesized FeS2
templates to direct the preparation of 1D nanomaterials. The most
nanorod films using plasma-assisted sulfurization of Fe2O3 precursors.
important benefit of ZnO template is the in-situ dissolution due to its
The sulfidation reaction occurs according to:
amphoteric characteristic. Our group has prepared FeS2 nanorod and
FeS2 nanotube arrays based on ZnO nanorod templates [62]. Fig. 9 plasma
Fe2 O3 + 4H2 S → 2FeS2 + 3H2 O+ H2 (4)
shows the schematic illustration of the growth of FeS2 nanorod arrays.
The basic synthesis processes are described as follow: (1) ZnO nanorod Hematite Fe2O3 nanorods were grown by chemical bath deposition
array template was immersed into Fe(NO3)3 for 30 min; (2) precursor in combination with calcination. Pyrite FeS2 with S:Fe atomic ratio of
Fe(OH)3 nanoarrays were formed after immersion and drying treat- ∼2.0 was formed under the action of plasma exposure at 400 °C for
ments; (3) FeS2 nanorod array was obtained by sulfurizing Fe(OH)3 at 60 min. Various experiments show that the chalcogen feed pressure
350 °C for 3 h. The chemical reactions occurred in the whole process are used in that method is obviously lower than requirements for thermal
[63]: sulfur vapor exposure [16], and sulfur incorporation can be completed

110
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 9. Schematic illustration of the synthesis of FeS2 nanorod arrays [62]. Reprinted with permission from [62]. Copyright (2014) Royal Society of Chemistry.

quickly at 400 °C. There are some other similar researches about the converted into FeS2 nanowires after sulfurization at 500 °C for 2 h. The
fabrication of 1D FeS2 nanomaterials via sulfidation of Fe2O3 precursor sulfidation process occurs according to the chemical reaction [68]:
films, as listed in Table 3. In another case, Huang et al. [66] used
FeOOH nanorod array as a precursor and converted it into FeS2 na- 4FeF3·x H2 O(s) + 5S2 (g) → 4FeS2 (s) + 2SF6(g) + 4x H2 O(g) (6)
norod through sulfidation reaction at 500 °C for 60 min. The chemical
reaction equation is described as following [67]: The sulfidation process was completed in a home-made tube reactor
equipped with pressure controls and gas-flow. The schematic illustra-
2FeOOH + 7S → 2FeS2 + H2 S+ 2SO2 (5)
tion of the tube reactor is shown in Fig. 11a. Fig. 11(b–g) shows the
Besides Fe2O3 and FeOOH precursors, α-FeF3·3H2O can also be used SEM images of 1D α-FeF3·3H2O, FeOOH and Fe2O3 precursors, and their
as precursor and transformed into FeS2 via thermal sulfidation treat- corresponding sulfurization products FeS2 nanostructures. It can be
ment. Li and co-worker [68] firstly synthesized α-FeF3·3H2O nanowires seen that the 1D structure of final FeS2 materials is well preserved after
precursor by dislocation- driven solution growth method, and it sulfidation of precursors, which has the control over the morphology

Fig. 10. SEM images of FeS2 nanostructures prepared by ZnO templates.(a, b) Fe(OH)3 precursor and FeS2 nanorod arrays [62], respectively; (c, d) Fe2O3 precursor
and FeS2 nanotube arrays [64], respectively. Reprinted with permission from [62]. Copyright (2014) Royal Society of Chemistry. Reprinted with permission from
[64]. Copyright (2016) Elsevier.

111
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Table 3
Various 1D FeS2 nanostructures synthesized by sulfidation of different precursors and corresponding sulfidation parameters.
Precursor films Sulfurizing parameters FeS2 structure Refs.

o
Precursor Preparation method Sulfur pressure Temperature( C) Time(h)

Fe2O3 nanorod Chemical bath deposition — 400 1 Nanorod [65]


Fe2O3/Fe3O4 nanotube Electrochemical anodization — 400 2 Nanowire [69]
Fe2O3 nanotube Electrochemical anodization 80 kPa 400 5 Nanotube [70]
FeOOH nanorod Hydrothermal method — 500 1 Nanorod [66]
α-FeF3·3H2O nanowire Solution synthesis 473 Torr 500 2 Nanowire [68]

hexagonal and triangular nanoplates with a size 200–500 nm were


obtained at 240 °C. Hu et al. [71] selectively obtained FeS2 nanosheets
via a one-step hydrothermal method. FeS2 nanosheets with diameters of
∼2 μm and thickness of 30 nm were obtained after Fe foil with purity of
99.99% and sulfur powder dissolved in deionized water reacted at
160 °C for 12 h.

3.4. 3D nanostructures

Li et al. [72] applied a microwave-assisted polyol method to prepare


FeS2 microspherolites with 3-dimensional sizes in micron scale
(Fig. 13). They performed reaction-time dependence experiments to
confirm their aggregation-based mechanism. The formation process
involved nucleation, growth and aggregation of the particles. The FeS2
microspheres were also synthesized through the hydrothermal method
with the help of PVP [73]. A synthetic route involving the possible role
of PVP was illustrated in Fig. 14. The growth process included three
steps: nucleation of the S and Fe2O3, the transformation of FeS2 and the
growth and assembly of FeS2 crystals.

4. Properties of FeS2 nanostructures

4.1. Optical properties

Common parameters used to characterize the optical properties of


semiconductors includes optical absorption coefficient (α), energy
bandgap (Eg), extinction coefficient and refractive index. Absorption
coefficient and energy bandgap are the two most popular parameters
which are widely studied and applied. They can be calculated according
to the optical spectra tested using an ultraviolet–visible spectro-
photometer and photoelectrochemical characterization. Absorption
coefficient α can be obtained by the following equation [74]:
Fig. 11. SEM images of precursors and as-synthesized FeS2 nanostructures. (a-
2.303 lg(I0/I ) ln(1 − R) 2.303Abs ln(1 − R) 2.303Abs
c) schematic illustration of the reactor, precursor α-FeF3·3H2O and FeS2 nano- α = + = + ≈
wire [68], respectively; (d, e) FeOOH and FeS2 nanorod [66], respectively; (f, g)
d d d d d
Fe2O3 nanotube array and FeS2 nanowire array, respectively; the inset is TEM (7)
image for a single nanowire [69]. Reprinted with permission from [66] and
Abs = lg(I0/ I ) (8)
[68]. Copyright (2013, 2014) Royal Society of Chemistry. Reprinted with
permission from [69]. Copyright (2014) American Chemical Society. where I0 is the incident intensity; I is the transmitted light intensity; d is
the thickness of sample; R represents the reflectivity; Abs represents the
and composition of materials. It further indicates that chemical trans- absorbance. Combining Eqs. (7) and (8), the value of α can be obtained
formation and structure duplication of existing nanostructures is a using the measured Abs value from light absorption spectrum and d
versatile method for nanomaterial synthesis. value from SEM or other approaches. There are three characteristic
absorption areas in the optical absorption spectra of semiconductors:
3.3. 2D nanostructures weak absorption area, linear absorption area and saturated absorption
area [74]. Semiconductors can be divided into two types: direct and
So far, research about 2D FeS2 nanostructures has rarely been re- indirect bandgap semiconductors. It can be distinguished by the feature
ported. Fig. 12 shows the morphologies of developed 2D FeS2 nanos- of shoulder structure shown in the liner absorption area. Indirect
tructures, such as nanoplates and nanosheet. Kirkeminde et al. [48] bandgap semiconductor has shoulder structure while direct type has no
synthesized FeS2 nanoplates via the reaction between Fe(CO)5 pre- shoulder structure. The difference may be caused by the different en-
cursor and oleyamine coordinated elemental sulfur solution at 180 °C or ergy band structures.
higher temperature and aging more than 180 min. They found that Similarly, the energy bandgap can also be fitted with the liner ab-
increasing reaction temperature had important effects on the planar sorption area in the optical absorption spectrum. The Eg value can be
growth but minor effect on the thickness of nanoplates. Truncated calculated by conventional Tuac equation [75]:

112
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 12. SEM images for (a) FeS2 nanoplates and (b) FeS2 nanosheets [48,71]. Reprinted with permission from [48,71]. Copyright (2012, 2008) American Chemical
Society.

Fig. 13. SEM images of the FeS2 microspherolites by microwave irradiation (a) low magnification and (b) high magnification [72]. Reprinted with permission from
[72]. Copyright (2011) Royal Society of Chemistry.

Table 4
Reported optical properties of FeS2 materials with different structures.
Structure Preparation method α (cm−1) Eg (eV) Refs.

Nanoparticles CVD 0.5 × 105 0.95 [79]


Nanoparticles Spray pyrolysis 0.5 × 105 0.72 [80]
Nanoparticles Spray pyrolysis 0.5 × 105 0.73 [81]
Nanoparticles Magnetron sputtering 1.2 × 105 0.8–1.6 [82]
Nanoparticles Magnetron sputtering 0.5 × 105 0.8–1.3 [83]
Nanoparticles Sol-gel 3 × 105 0.9–0.99 [84]
Cubes octahedra Hydrothermal 0.26 × 104 1.25 [33]
Nanocrystals Hot injection — 0.88–0.91 [41]
Nanocrystals Hot injection 2 × 105 0.93 [42]
Nanowires Direct sulfidation — 0.89 [59]
Nanocubes Chemical bath deposition ∼106 ∼1.1 [49]
Nanorods Sulfidation of precursors 1.2 × 105 0.89 [62]
Nanotubes 1.8 × 105 0.77 [64]

showed an absorption coefficient of 5 × 104 cm−1 when the photon


Fig. 14. Schematic depicting the possible role of PVP in the synthesis of FeS2 energy was larger than 2 eV. They found that the absorption coefficient
microspheres [73]. Reprinted with permission from [73]. Copyright (2016) value changed with the change of the ratio of chemical ingredients.
Elsevier. Meng et al. [76] prepared various FeS2 films by sulfidation of magne-
tron sputtered Fe films under different sulfur pressure and tested their
αhv = A (hv − Eg )n (9) optical absorption performance. The absorption coefficient distributed
in the range of (4–8) × 104 cm−1. Recently, they fabricated one-di-
where hv represents the energy of the incident photon, and A represents mensional FeS2 nanoarray films by templated directed synthesis in the
a constant. n = 2 for an indirect bandgap semiconductor and n = 1/2 combination of subsequent sulfidation. The FeS2 nanoarray films ex-
for a direct bandgap semiconductor. The Eg value is obtained by fitting hibited excellent light absorption with absorption coefficient
the absorption coefficient to the general formula. ≥105 cm−1 and energy bandgap of 0.89 eV. Sanchez et al. [77] sys-
A large number of researches discover a common phenomenon that tematically studied the relationship between absorption edge of FeS2
pyrite FeS2 shows extremely optical absorption coefficient (105–106 films and the testing temperature. They tested the optical absorption
cm−1). Energy bandgap has a close relationship with the preparation spectra of FeS2 films at the temperature range from 10 to 300 K. When
methods, morphology and structure of materials. Takabashi et al. [15] the photon energy was lower than 2 eV, the absorption coefficient in-
prepared FeS2 films by CVD method under normal pressure. The films creased exponentially with the increase of photon energy; when the

113
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Table 5
Reported electrical properties of FeS2 materials prepared by different technologies.
Preparation method Shape Conductivity (Ω·cm)−1 Hall mobility (cm2/Vs) Carriers concentration Conducting type Refs.

18
CVD Nanoparticles 0.1 20 6.5 × 10 p [93]
Spray pyrolysis Nanoparticles 0.3 1.5 2.3 × 1017 p [94]
CVD Nanoparticles 0.97 280 5.5 × 1017 p [95]
MOCVD Nanoparticles 0.1 2.4 ≥1020 n [89]
Magnetron sputtering Nanoparticles 1.3–3.3 2 1020–1021 p [76]
Electrodeposition Nanoparticles – 213 2.9 × 1014 p [12]
Sol-gel Nanoparticles 0.01 32.9–429 1.5 × 1019 n [84]
2.7 × 1019
Direct sulfidation Nanowires 3.7–11.1 0.03 4 × 1021 p [59]
Sulfidation of Fe film Film – 7.1 2.9 × 1018 [96]

Fig. 15. SEM, TEM and electrochemical characterizations of FeS2 nanotube and nanoparticle films [64]. (a, b) SEM and TEM images for FeS2 nanotubes; (c) SEM for
FeS2 nanoparticles; (d) Tafel polarization curves; (e) EIS curves. Reprinted with permission from [64]. Copyright (2016) Elsevier.

Table 6
Electrochemical polarization parameters for FeS2 nanotube and FeS2 nanoparticle films per unit area (1 cm × 1 cm), respectively [64].
Structures Ecorr (mV/SCE) βc (mV·decade-1) βa (mV·decade-1) jcorr (mA·cm−2)

Nanotube 331 207 169 0.0096


Nanoparticle 219 148 200 0.0014

Table 7
Fitted parameters extracted from EIS of the two structured FeS2 films per unit area (1 cm × 1 cm) [64].
Structures Rs (Ω·cm2) Q (Ω−1·cm−2) Rct (Ω·cm2) W (Ω·cm2) X 2

sn n

Nanotube 20.2 8.50 × 10−4 0.94 3148 2.24 × 10−3 8.01 × 10−4
Nanoparticle 17.7 1.45 × 10−4 0.83 11,240 3.20 × 10−4 2.74 × 10−3

114
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 16. SEM, TEM and photoelectrochemical characterizations of FeS2-sensitized TiO2 nanotubes [105]. (a, b) SEM images for TiO2 nanotubes before and after
sensitization; (c, d) top-view and cross-sectional view TEM images; (e) optical absorption spectra; (f) photocurrent-potential curves. Reprinted with permission from
[105]. Copyright (2010) American Chemical Society.

photon energy was higher than 2 eV, the absorption coefficient would of temperature and grain boundary on the conducting property. Based
no longer increase and kept stable at (1.25–1.3) × 105 cm−1. The ap- on the model, Baccarnai et al. [86] further modified it and proposed an
pearance of absorption edge mainly is ascribed to the electron transi- improved version. The modified model suits to describe polycrystal
tion from t2g state to eg state. Experimental results indicated that the semiconductor materials. There are a great number of grain boundaries
energy bandgap of FeS2 films varied from 0.99 eV to 1.05 eV; it almost and defects (broken bonds, vacancies and dislocations) existed in the
kept unchanged at the temperature of 10–100 K and it decreased with inner of polycrystal materials, leading to the production of defect level
the increase of temperature when the temperature was at 100–300 K. in the region close to the bottom of the conduction band. Carriers are
Impurity was considered to be detrimental to the electron and photonic prone to be captured when crossing grain boundaries, and the barrier
performance of photoelectrodes. In contrast to the traditional view that would produce at a grain boundary, which would affect the electrical
marcasite was a deleterious phase in FeS2, a recent report suggested properties of films. The resistivity and temperature satisfy the following
that marcasite was beneficial for the charge separation across the equation:
pyrite-marcasite phase junction [78].
A′ qΦ
Results indicate that the optical properties of FeS2 films are closely ρ= exp( B )
Tn kT (10)
related to the microstructure, synthesis approach and defects. Table 4
summarizes the reported optical properties of FeS2 materials prepared where ρ is resistivity, A’ is a constant, T is thermodynamic temperature,
by different methods and with different structures. q is unit charge, n is index, k is Boltzmann constant and ΦB is the barrier
height of grain boundary. The equation is suitable for describing the
conducting behavior at high temperature.
4.2. Electrical properties Efros et al. [87] made some corrections and proposed a suitable
mechanism which follows the relationship expression:
The conventional conducting mechanism, namely, electron transi-
TE 1
tion between the conduction band and valence band, is not suitable to ρ = BT exp( )2
T (11)
explain the conductivity change of FeS2 films. Ares et al. [85] proposed
a model of grain boundary electron barrier which considered the effects where TE and B are constant.

115
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 17. SEM, TEM and photoelectrochemical analyses of FeS2-sensitized ZnO/ZnS nanorod arrays [107]. (a) SEM image; (b, c) TEM images; (d, e) photocurrent
response during on–off cycles of illumination before and after a 30 min interval; (f) schematic illustration of energy level structure before and after distribution and
photogenerated carrier separation and transport process. Reprinted with permission from [107]. Copyright (2015) WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

Table 8 For single-crystalline semiconductor materials, grain boundary


Comparison of DSSCs performances using Pt and various structure FeS2 as barrier model based on boundary theory is obviously not suitable for
counter electrodes. this system. Ares et al. [85] proposed point defect model, which
Counter electrode VOC (mV) JSC (mA/ FF% η% Refs. thought the electrical behavior of FeS2 was dominated by the point
cm2) defects occurred in crystals. The relationship between resistivity and
temperature can be expressed by the equation:
Pt 710 15.37 0.69 7.52 [111]
FeS2 nanocrystal ink 710 15.14 0.68 7.31
Em 1 σ2 1
Pt 780 14.77 0.66 7.54 [21] ρ = ρ0 exp[ ( ) − E2 ( )2]
k T 2k T (12)
FeS2 nanoparticles 790 15.20 0.65 7.97

Pt 685 13.36 68.2 6.23 [112] where Em and σE represent the average value and standard deviation
FeS2 nanoparticles 658 12.56 57.8 4.78 value of defect level.
FeS2 nanowires 653 13.68 65.7 5.88
Taking advantage of above three conducting mechanism can well
Pt 730 15.7 0.66 7.56 [113] explain the variation rule of FeS2 resistivity at different temperatures.
FeS2 nanochains 720 15.3 0.65 7.16
Besides, researchers also made detailed studies about the Hall test of
Melt FeS2 nanoparticles 700 10.4 0.58 4.22
FeS2 films. Altermatt et al. [88] applied numerical simulation to in-
Pt 758 10.10 0.273 2.09 [114] vestigate the electrical properties of FeS2 films, such as Hall coefficient,
Solid FeS2 micro-sphere 746 13.00 0.361 3.50
Mesoporous FeS2 micro-sphere 743 13.58 0.387 3.90
carrier mobility and lifetime of majority carriers. They also summarized
the parameters of FeS2 films prepared by different methods. Oertel et al.
[89] found the carrier concentration of n-type FeS2 varied from
1018–1021 cm−3, while the concentration for p-type FeS2 was lower
than 1018 cm−3. When the FeS2 was alloyed with Ni by ion injection,

116
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

(a)

(c)

Fig. 18. (a) SEM and (b) TEM image of the FeS2@carbon fiber, (c) discharge energy density vs cycling number [122]. Reprinted with permission from [122].
Copyright (2016) American Chemical Society.

the carrier type changed from holes to electrons once the Ni con- by cyclic/liner voltammograms, Tafel polarization curves and electro-
centration exceeded 20% [90]. Ferrer et al. [91] discovered that FeS2 chemical impedance spectroscopy (EIS) [101].
films obtained by sulfidation of iron films usually exhibited n-type Liu et al. [102] investigated the influence of electrolyte on the
conducting and Hall coefficient was only ∼ 0.57 cm−3. Although the electrochemical behavior of pyrite electrode by changing the con-
conducting mechanism of FeS2 films can be expressed by above the- centrations of Fe3+ in the sulfuric acid solution. They obtained a similar
ories, the specific electrical parameters tested during experiments have conclusion that the pyrite was oxidized by two steps: the first one was
a close relationship with preparation techniques. Table 5 summarizes the oxidation of pyrite to elemental sulfur and the second one is the
the reported electrical properties of FeS2 materials prepared by dif- further oxidation of sulfur to sulfate. Recently, Meng et al. [64] fabri-
ferent methods and with different structures. Due to the high con- cated FeS2 nanotubes via layer-by-layer assembly and studied the effect
ductivity and suitable energy level for hole extraction, FeS2 is con- of one-dimensional structure on the electrochemical properties of FeS2.
sidered as a promising candidate for the hole transport materials in Comparisons of the electrochemical properties of FeS2 nanotubes and
perovskite solar cells [92]. FeS2 nanoparticles were conducted under identical conditions. As the
active surface area was closely related to some electrochemical para-
4.3. Electrochemical properties meter value, they made a brief calculation of the surface area of two
structured FeS2 films. Results indicate that the surface area of FeS2
As early as 1978, Biegler et al. [97] have studied the anodic beha- nanotubes per unit area (1 cm × 1 cm) is approximately as fourteen
vior of pyrite FeS2 in acid solutions. They found the anodic oxidation of times as FeS2 nanoparticles. Fig. 15(d) showed the Tafel polarization
FeS2 in acid solution was a complicated process containing two path- curves for the two different structured FeS2 films and corresponding
ways. The anode reaction equations are described as following [97]: parameter values after considering active surface area were presented
in Table 6. Fig. 15(e) showed the EIS plots for two structured FeS2 films
FeS2 + 8H2 O→ Fe3 + + 2SO24− + 16H + + 15e− (13) and corresponding parameter values after considering active surface
area were listed in Table 7. The electrochemical results indicated that
FeS2 → Fe3 + + 2 S+ 3e− (14)
nanotube structure possessed high corrosion resistance and chemical
Before this research, there were also some relevant reports about the stability.
electrochemical research of pyrite FeS2. Peters and Majima [98]
thought that electrochemical reactions were not responsible for sulfur 5. Applications
formation during acid pressure leaching. Bailey and Peters [99] made
some modification and could obtain electrochemical data at high 5.1. DSSCs/QDSSCs
temperature (110 °C) and pressures (1.21 MPa). Nagai and Kiuchi [100]
studied the anodic oxidation of pyrite at a higher temperature (175 °C) A common application of narrow bandgap semiconductors is used as
and obtained corresponding electrochemical data such as elemental inorganic sensitizers (or called quantum dots) to sensitize with wide
sulfur yields and charge requirements. Afterwards, systematic re- bandgap semiconductors, such as TiO2 and ZnO. It can broaden the
searches about the anodic oxidation of pyrite FeS2 have been performed absorption range of composite materials to visible region and as a result

117
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

pyrite FeS2 was regarded as an effective sensitizer for TiO2 to assist the
photodegradation. Wang et al. [107] sensitized ZnO/ZnS nanorod ar-
rays with pyrite FeS2 quantum dots and the configuration of ZnO/ZnS/
FeS2 ternary composite nanorod arrays was described as following
(Fig. 17(a–c)): ZnO phase located in the core part of a nanorod, ZnS
phase adjoined ZnO core, and FeS2 quantum dots randomly scattered
around the surface of ZnS shell. The FeS2-sensitized ZnO/ZnS nanorod
arrays showed enhanced optical absorption property with the extension
of absorption edge into the visible region. Also, the ZnO/ZnS/FeS2
nanorod arrays show good stability under irradiation as shown in
Fig. 17(d) and (e). The enhanced photo performances of FeS2-sensitized
ZnO/ZnS nanorod arrays mainly benefits from the excellent optical
absorption performance of pyrite FeS2 and the suitable gradient energy
level structure which fosters the separation and transport of photo-
generated carriers. Fig. 17(f) shows the schematic illustrations of en-
ergy level structure before and after distribution and photogenerated
carrier separation and transport process in ZnO/ZnS/FeS2 ternary
composite nanorods. Bedja et al. [108] fabricated FeS2-quantum-dot
sensitized metal oxide photoelectrodes and tested their photoelec-
trochemistry and photoinduced absorption spectroscopy to study their
optical absorption and DSSCs performances.
Recently, another newly developed application of FeS2 materials is
used as counter electrodes in DSSCs and QDSSCs. DSSCs/QDSSCs has
long been considered as a promising alternatives for conventional si-
licon-based solar cells due to their low cost, easy fabrication and great
application prospect [109]. The counter electrode is an important
component of DSSCs/QDSSCs, in which electrons are injected into the
electrolyte to catalyze the reduction of oxidized dye molecules [110]. In
past decades, many researchers designed and synthesized abundant
FeS2 nanostructures, and applied them as counter electrodes in DSSCs/
QDSSCs. Comparative analysis of the cell performances using FeS2 and
Pt as counter electrodes were made. Table 8 summarizes reported cell
performances using different structured FeS2 as counter electrodes in
DSSCs/QDSSCs. Results indicate DSSCs/QDSSCs based on FeS2 counter
electrodes exhibited excellent electrochemical catalysis and stability.
Besides, the cell performances of DSSCs using FeS2 as the counter
electrode can be comparable to that of DSSCs using Pt as the counter
electrode.

5.2. Lithium-ion batteries


Fig. 19. (a) Energy level and device structure of FeS2/CdS photodetector; (b)
cross-sectional SEM image of FeS2/CdS film; (c) current–voltage characteristic Due to cheap and environmentally friendly features, pyrite FeS2 was
of FeS2/CdS photodetector; on/off switching of (d, f) FeS2 quantum dots/CdS applied in Li/FeS2 batteries. Unfortunately, the poor cyclability of FeS2
and (e, g) FeS2 cubes/CdS devices [131]. Reprinted with permission from seriously limits its application in rechargeable batteries [3]. Intercala-
[131]. Copyright (2013) WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. tion and conversion are two common electrochemical reactions in re-
chargeable Li/FeS2 batteries, corresponding to Eqs. (15) and (16), re-
strengthen the optical absorption performance of materials. Among the spectively. Both the two reactions co-exist at the first cycle [115]. In the
reported sensitizers, CdS and CdSe are two most popular sensitizers and first discharge cycle, FeS2 reacts with Li to form Li2FeS2 with metallic
have been widely used to sensitize TiO2 materials, which can accelerate conductivity [116], and then forms Fe metal and Li2S. The latter re-
the separation of photogenerated carriers and significantly enhance the action brings severe volume change to the electrode materials and
photocurrent in visible region [103]. Although FeS2 has suitable sluggish kinetics for the reconstruction of FeS2 in the charging process,
bandgap (Eg = 0.95 eV) and extremely high absorption coefficient resulting in inferior cyclic performance [117].
(α > 105 cm−1). However, the loss of open circuit voltage prevents its
FeS2 + x Li+ + x e− → Lix FeS2 (0 < x < 2, 1.5 − 1.7V) (15)
sole application in photovoltaics. A recent study through ultrafast laser
spectroscopy revealed that fast carrier localization of photoexcited Lix FeS2 + (4 − x )Li+ + (4 − x ) e− → Fe + 2Li2 S (∼ 1.5V) (16)
carriers to indirect band edge and shallow trap states were responsible
for major carrier loss [104]. Therefore, FeS2 is suitable to use as a FeS2 can be utilized as both anode and cathode in lithium-ion bat-
promising semiconductor photosensitizer to avoid the disadvantage and teries. Recently, various related researches are reported about the ap-
exhibit its advantage. Song et al. [105] fabricated TiO2 nanotube arrays plication of pyrite FeS2 or its composites acting as anode or cathode in
and sensitized it with pyrite FeS2 (Fig. 16(a-d)). The visible light re- lithium-ion batteries. Liu et al. [39] prepared FeS2 nanocubes with sizes
sponse was remarkably enhanced after sensitization (Fig. 16(e) and (f)). of around 80–120 nm using a solvothermal method and utilized it as
Recent work reported that the absorption edge was further extended to anodes in lithium-ion batteries. A reversible discharge capacity of 540
infrared light region [106]. The FeS2/TiO2 photoelectrodes had a mAh g−1 after 150 cycles at a current density of 1 A g−1 was achieved.
photocurrent enhancement by more than 3 orders of magnitude than Tan et al. [118] synthesized core–shell nano-FeS2@N-doped gra-
the pristine TiO2 under illumination of near-infrared light. Besides, phene composite and used it as a cathode for lithium-ion batteries. The
FeS2@N-graphene exhibits a remarkable specific energy (950 W h kg−1

118
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 20. (a) Schematic of the nanogap device; (b, d) SEM image of the nanogap before and after the nanocrystals removed; (c) cross-sectional SEM image of the FeS2
nanocrystal layer; (e) I-V characteristics of the nanogap with and without FeS2 in dark and under illumination; (f) photoresponse of the nanogap photodetector [30].
Reprinted with permission from [30]. Copyright (2014) Springer Science.

at 0.15 kW g−1) and higher specific power (543 W h kg−1 at The reasons for the improvement of lithium-ion battery perfor-
2.79 kW g−1) than commercial lithium-ion battery cathodes, as well as mance by the introduction of carbon materials in FeS2-based cathode
stable cycling performance (∼600 W h kg−1 at 0.75 kW g−1 after 400 are that the carbon materials provide a good conductive network, mi-
cycles). Carbon coating is an effective way to improve electrochemical tigate the volume change of pyrite FeS2, inhibit the growth of FeS2
performance of FeS2 and is widely adopted in lithium-ion batteries nanocrystals to get smaller particles, and reduce the loss of active
[119,120]. Zhang et al. [121] used FeS2/C composite as a cathode for materials during the lithiation-delithiation process.
lithium-ion battery. They found the presence of carbon not only sur-
rounded the FeS2 surface but also penetrated into the entire FeS2 par- 5.3. Sodium-ion batteries
ticle, forming continuously conductive networks throughout the FeS2
particle. As a result, the rate capability of Li/FeS2 cells was improved Besides lithium-ion batteries, FeS2 was applied as a promising
without producing much effect on the specific capacity and capacity electrode materials for sodium-ion batteries with a high theoretical
retention of the FeS2 cathode. Zhu et al. [122] obtained FeS2@carbon capacity of 894 mAh g−1 [124]. Electrochemical reactions in re-
fiber by electrospun and used it as a cathode for lithium-ion batteries chargeable Na/FeS2 batteries are as following [125]:
(Fig. 18). The FeS2@carbon fiber electrode shows stable cycling per-
FeS2 + x Na+ + x e− → Nax FeS2 (x < 2) (17)
formance in both the conventional carbonate electrolyte and the sol-
vent-in-salt-type Li − S battery electrolyte. Another porous FeS2@C Nax FeS2 + (4 − x )Na+ + (4 − x )e− → Fe + 2Na2S (below 0.8 V) (18)
nanowire was synthesized through an organic–inorganic hybrid method
Large volume expansion upon sodiation is a major barrier to the
and used as cathodes by Zhang et al. [123]. The porous FeS2@C na-
cyclability of FeS2. A higher voltage cut-off (0.8 V) could avoid that
nowire showed a specific capacity of 889 mAh g−1 at 0.1 A g−1 and
problem and achieved an unprecedented long-term cyclability (∼90%
521 mAh g−1 at 10 A g−1. Moreover, the discharge energy density re-
capacity retention for 20 000 cycles, 210 mAh g−1) at sacrificing some
mained as high as 637 mAh g−1 after 1000 cycles at 2 A g−1.
specific capacity [125]. Combining the reduced graphene oxide aerogel

119
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

method to form FeS2/CdS bulk heterojunctions (BHJs) structure and


fabricate FeS2/CdS BHJ photodetector. Fig. 19(a) and (b) showed the
schematic illustration of the structure as well as energy level and SEM
image of the synthesized FeS2/CdS film. The device showed broad
spectral photosensitivity covering almost the whole spectral range. The
photoresponse time is about 10 ms under AM 1.5 light source with an
intensity of 100 mW cm−2. When a low bias voltage of 1 V was applied,
the responsivity and external quantum efficiency under 1100 nm light
illumination were 174.9 A W−1 and 1.98 × 104%, respectively, as
shown in Fig. 19(d–g). Magnetic CdFeS phase was formed in the pre-
paration of FeS2/CdS BHJ structure, which provided magnetic field
effect for 72.6% tunability of the photocurrent under illumination.
Liu et al. [30] utilized high-quality FeS2 nanocrystals obtained by a
hydrothermal method to fabricate a crystalline nanogap pyrite photo-
detector. Fig. 20(a) displayed the schematic illustration of the nanogap
photodetector based on pyrite FeS2 nanocrystals and corresponding
SEM images (Fig. 20(b–d)). The fabrication processes were described as
follow: (i) p-type silicon substrate covered with a thick layer of thermal
oxide was obtained by standard photolithography; (ii) a bridge-like Ni
electrode with thickness of 300 nm was deposited using e-beam eva-
poration to provide ohmic contact for FeS2; (iii) a narrow gap with
Fig. 21. Schematic illustrations of the photocapacitor device and its operating
width of 200 nm and depth of 300 nm was obtained by breaking the Ni
principle [133]. (a) Schematic device structure and SEM image of the FeS2-
wire using focused ion beam; (iv) the Ni electrode nanogap was filled
based device; (b) operating principle. Reprinted with permission from [133].
Copyright (2013) Royal Society of Chemistry. by FeS2 nanocrystals and further sulfurized at 500 °C for 2 h. The na-
nogap photodetector performed a very high photocurrent in the range
of 10-2 to 1 μA for approximately 1 μm2 gap area and spectral response
with FeS2 microspheres improved the conductivity and obtained a low in the UV–vis range, as shown in Fig. 20(e) and (f).
decay rate of 0.051% per cycle over 800 cycles at 1C [126]. Tuning the
morphology and structure of FeS2 at nanoscale may enhance the reac- 5.5. Photocapacitor
tion kinetics and mitigate the large volume expansion. Paik et al. [124]
adopted an etching method coupled with sulfidation-in-nanobox Photocapacitor is a type of capacitor that can be charged by solar
strategy to synthesize FeS2@C yolk-shell nanoboxes and achieved ex- light and allows for direct storage of solar energy. Most photocapacitors
cellent sodium storage performance. The sodium-ion batteries delivered are a hybrid structure consisting of a dye-sensitized cell using standard
a high specific capacity of 511 mAh g−1 at 100 mA g−1 after 100 cycles TiO2 decorated with a dye and an electrolyte [132]. Gong et al. [133]
and 330 mAh g−1 at 2 A g−1 after 800 cycles. Replacing the carbon simplified this setup by using one material-electrolyte system which
shell with FeSe2 shell achieved a high capacity of 350 mAh g−1 at 1 A acted as both the active material and blocking electrode (Fig. 21). The
g−1 after 2700 cycles and 301.5 mAh g−1 at 5 A g−1 after 3850 cycles synthesized FeS2 nanocrystals showed promising light harvesting cap-
with over 97% coulombic efficiency [127]. abilities across the visible to near-infrared spectrum. The assembled
Besides the application in lithium-ion and sodium-ion batteries, device exhibited a high performance near-infrared photodetectivity of
FeS2 was also used in lithium-sulfur and sodium-sulfur batteries. 1.6×1011 Jones, a specific capacitance of 37.5 mAh g−1 and an energy
Okuyama et al. [128] firstly investigated the effect of FeS2 in the density of 1.13×104 J g−1. The ionic liquid electrolyte in photo-
cathode on sodium-sulfur battery performance. The capacity of the capacitors played a key role in the formation of the electric double layer
batteries was heavily decreased by doping with FeS2. A FeS2 3.0 wt% and passivating surface states of FeS2 nanocrystals to accelerate charge
doped cell exhibited a 65% resistance rise and a 65% capacity loss transport.
during cycling. Direct adding FeS2 in a fresh lithium-sulfur cell also
resulted in a significant shuttling effect, preventing further cell cycling 5.6. Photocatalysis
[129]. Pre-cycling the sodium-FeS2 cell or using LiNO3 electrolyte ad-
ditive helped to prevent the severe shuttling effect. However, FeS2 Although much attention has been delivered to transition metal
could be used as an efficient adsorbent of lithium polysulfide to im- sulfides as photocatalysis reports about the adsorption and photo-
prove lithium-sulfur batteries [130]. It was found that FeS2 had ex- catalysis of FeS2 towards organic dyes are still few. Liu et al. [134]
cellent ability for chemically adsorbing Li2Sx to sequestrate the dis- prepared FeS2 particles by a solvothermal method and studied its ad-
solved Li2Sn from diffusing out of the cathode in lithium-sulfur sorption and photocatalysis performance of organic dyes. Comparison
batteries. Ultrafine FeS2 nanoparticles showed exceptional advantages of the adsorption and photocatalytical degradation abilities of the or-
in sodium-sulfur and lithium-sulfur reversible conversion reactions ganic dyes were made. The adsorption and photocatalysis experiments
[45]. Reversible capacities over 500 and 600 mAh g−1 for sodium and were performed according to the following instructions. The first step,
lithium storage were observed with improved cycling and rate capacity. FeS2 particle catalysts were dispersed into a mixture solution containing
That was due to the ultrafine size comparable to the diffusion length of above five organic dyes. The obtained suspension was sonicated for
Fe during cation exchange. 30 min and stirred in the dark for 20 min. During this period, UV-vis
absorption spectrophotometer was utilized to measure the optical
5.4. Photodetector property changes of organic dyes at different adsorption time. An ad-
sorption and desorption equilibrium was established after sonicating for
Recently, various semiconductor quantum dots have been ex- 30 min and stirring for 20 min in the dark. The second step, the mixed
tensively used to assemble photodetectors due to the advantages of system was irradiated by a UV light for given time. Fig. 22 showed the
simple fabrication and large-area solution processable features by spin adsorption and photocatalytic degradation performance of FeS2 parti-
coating or roll-to-roll printing. Gong et al. [131] added CdS into FeS2 cles for five different organic dyes, respectively. Some key parameter
nanocrystals (quantum dots and cubes) by chemical bath deposition values extracted from Fig. 22 were presented in Table 9. Results

120
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 22. SEM image of FeS2 particles and degradation ratio of five organic dyes [134]. (a) SEM image of FeS2 particles; (b) Methylene blue; (c) Safranine T; (d) Methyl
orange; (e) Rhodamine B; (f) Pyronine B. Reprinted with permission from [134]. Copyright (2013) Elsevier.

Table 9
Degradation ratio of FeS2 particles on five different organic dyes [134].
Dyes Adsorption ratio (after 50 min) Photocatalytic degradation ratio (after 90 min) Degradation ratio (after 140 min) Degradation ratio (after 24 h)

Methylene blue 80.2% 7.7% 87.9% 99.2%


Safranine T 88.8% 6.0% 94.8% 99.5%
Methyl orange 58.6% 11.5% 70.1% 87.9%
Rhodamine B 43.0% 14.8% 57.8% 97.0%
Pyronine B 47.7% 8.6% 56.3% 97.9%

indicate that the FeS2 particles possess higher adsorption ability than sulfidation of Fe2O3 nanotube films obtained via anodization of iron
photocatalytic degradation capacity. foils. They also prepared FeS2 particle films for comparison.
Tian et al. [135] indirectly prepared FeS2 nanotube array films by Fig. 23(a–c) showed the SEM images of the as-synthesized Fe2O3

121
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

Fig. 23. SEM images of (a) Fe2O3 nanotube, (b) FeS2 nanotubes and (c) FeS2 microparticles and (d) their photodegradation performances of phenol [135]. Reprinted
with permission from [135]. Copyright (2015) Royal Society of Chemistry.

nanotube, FeS2 nanotubes and FeS2 microparticles. The photocatalysis improve the HER performance of the FeS2. The corresponding cathodic
performance of Fe2O3 nanotubes, FeS2 nanotubes and FeS2 micro- overpotential of the (Fe0.48Co0.52)S2 was 143 mV cm−2 and the Tafel
particles on methylene blue and phenol were studied using a UV slope was 47.5 mV/decade. In fact, the content of Co in the
spectrophotometer. Fig. 23(d) showed the photodegradation perfor- (Fe0.48Co0.52)S2 may be too high. Dai et al. [139] reported that the
mances of the three catalysts on phenol. The results suggested that the cathodic overpotential for the Fe0.9Co0.1S2 to drive the HER at
oxidation of pyrite affected the photocatalytic activity of pyrite nano- 20 mA cm−2 was 120 mV and the Tafel slope was 46 mV/decade.
tubes slightly. The visible light driven catalytic activity of the pyrite Density functional theory calculation revealed that the high catalytic
nanotube coating was derived from increased production of hydroxyl activity stemmed from a large reduction of the kinetic energy barrier of
radicals. The topography of the nanotube array contributed to the H atom adsorption on FeS2 surface upon Co doping. Besides the stoi-
generation of the photoexcited hole+ and separation of the e-/hole+ chiometry, the morphology of the FeS2 also play important effect on the
pairs, eventually enhanced photocatalysis performance. HER performance. Leonard et al. [140] found the synthesized 2D FeS2
Besides the photocatalytic degradation of organic dyes, FeS2 was discs displayed excellent electrochemical activity similar to Pt, with
also employed to photocatalytic production of hydrogen. Lee et al. high exchange current density and an onset potential for hydrogen
[136] compared the photocatalytic hydrogen production by pure TiO2, evolution near the thermodynamic potential. The 2D FeS2 discs are
pure FeS2 and TiO2/FeS2 composites. They showed that only 2.0 mmol remarkably stable, demonstrating the ability to generate hydrogen for
H2 was collected from pure TiO2 and pure FeS2 over 10 h, whereas over 125 h. The electrocatalytic performance of the FeS2 was further
9.8 mmol H2 was collected from TiO2/FeS2 composites. That was be- examined by using as catalyst in a proton exchange membrane elec-
cause that the electrons-hole recombination in TiO2 was suppressed and trolysis single cell [141]. The FeS2 allowed achievement of a current
the photocatalytic activity increases with the introduction of FeS2. The density of 2 A cm−2 at a voltage of 2.3 V. The choice of alloying ele-
optimized content of FeS2 doped in TiO2 was about 10.0 wt%. The FeS2 ments and the control of nanostructure to optimize the electrochemical
with the smallest particle size was proved to have the highest current performance of FeS2 towards HER still need further studies.
density for hydrogen production [137].

6. Conclusions and outlooks


5.7. Electrocatalysis
A comprehensive review of the developed researches of various
The possibility of using earth abundant and inexpensive FeS2 to structured pyrite FeS2 is provided in this paper. We firstly introduce the
replace precious Pt as hydrogen evolution reaction (HER) has attracted synthesis of FeS2 nanostructures from four categories, 0D, 1D, 2D and
intense search recently. Jin et al. [138] proved that the FeS2 prepared 3D, respectively. It is focused on various synthetic methods and the
directly on graphite substrate through thermal sulfidation of metal related growth mechanisms. The success of the preparation of FeS2
precursor film was highly efficient HER electrocatalyst. The cathodic nanostructures with desired dimensions and morphologies is critical for
overpotential required for the FeS2 to drive the HER at 1 mA cm−2 was achieving unique properties. A summary of synthesis techniques for
217 mV and the Tafel slope was 56.4 mV/decade. Co alloying could obtaining diverse types of FeS2 nanostructured with various

122
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

morphologies, including nanoparticles, nanowires, nanorods, nano- (2000) 1803–1806.


tubes, nanosheets, nanoplates and microspherolites. The influences of [15] N. Takahashi, T. Yatomi, T. Nakamura, Solid State Sci. 6 (2004) 1269–1272.
[16] N. Berry, M. Cheng, C.L. Perkins, M. Limpinsel, J.C. Hemminger, M. Law, Adv.
the synthesis parameters in producing the nanoscale features and the Energy Mater. 2 (2012) 1124–1135.
corresponding formation mechanisms are also analyzed. Also, photo- [17] L. Samad, M. Caban-Acevedo, M.J. Shearer, K. Park, R.J. Hamers, S. Jin, Chem.
electrochemical properties and potential applications are also pre- Mater. 27 (2015) 3108–3114.
[18] I.G. Orletskii, P.D. Mar'yanchuk, E.V. Maistruk, M.N. Solovan, V.V. Brus, Phys.
sented. Especially the electrochemical properties (catalytic activity, Solid State 58 (2016) 37–41.
corrosion resistance and electrochemical stability) in the applications of [19] B. Ouertani, J. Ouerfelli, M. Saadoun, B. Bessais, H. Ezzaouia, J.C. Bernede, Mater.
photoelectrochemical cells with different electrolyte systems are ana- Charact. 54 (2005) 431–437.
[20] A. Yamamoto, M. Nakamura, A. Seki, E.L. Li, A. Hashimoto, S. Nakamura, Sol.
lyzed. It is important for FeS2 in the future applications like DSSCs/ Energy Mater. Sol. Cells 75 (2003) 451–456.
QDSSCs, Lithium-ion batteries, sodium-ion batteries, photocapacitors, [21] S. Shukla, L. Nguyen Huu, P.P. Boix, T.M. Koh, R.R. Prabhakar, H.K. Mulmudi,
photocatalysis and electrocatalysis. J. Zhang, S. Chen, C.F. Ng, C.H.A. Huan, N. Mathews, T. Sritharan, Q. Xiong, ACS
Nano 8 (2014) 10597–10605.
Although encouraging progresses in FeS2 nanostructures have been
[22] G. Willeke, R. Dasbach, B. Sailer, E. Bucher, Thin Solid Films 213 (1992) 271–276.
achieved, numerous problems still exist and wait for addressing. [23] L.Y. Huang, L. Meng, Mater. Chem. Phys. 124 (2010) 413–416.
[24] H. Zhang, Y.S. Liu, B.Y. Wang, L. Wei, R.X. Kui, H.J. Qian, Chin. Phys. B 18 (2009)
1. Challenges still remain in the synthesis of single crystalline, high 2734–2738.
[25] H.F. Liu, D.Z. Chi, J. Vac. Sci. Technol. A 30 (2012) 04D102.
purity and satisfying stoichiometric ratio FeS2 nanostructures. The [26] S. Hunger, L.G. Benning, Geochem. Trans. 8 (2007) 1–20.
presence of Fe1-xS impurities is commonly observed in the FeS2 [27] H. Xian, J. Zhu, X. Liang, H. He, RSC Adv. 6 (2016) (1999) 31988–31993.
system. Further work should be performed by improving currently [28] A.M. Golsheikh, N.M. Huang, H.N. Lim, C.H. Chia, I. Harrison, M.R. Muhamad,
Chem. Eng. J. 218 (2013) 276–284.
existed preparation methods or exploring new technologies to ob- [29] C. Wadia, Y. Wu, S. Gul, S.K. Volkman, J. Guo, A.P. Alivisatos, Chem. Mater. 21
tain high-quality FeS2 materials. (2009) 2568–2570.
2. The current photoelectric conversion efficiency of FeS2-based de- [30] S. Liu, J. Wu, P. Yu, Q. Ding, Z. Zhou, H. Li, C.-C. Lai, Y.-L. Chueh, Z.M. Wang,
Nanoscale Res. Lett. 9 (2014).
vices is still low. A major reason is due to the defects (S vacancies [31] Z. Shi, A.H. Jayatissa, F.C. Peiris, J. Mater. Sci. Mater. El. 27 (2016) 535–542.
and interstitial atoms) produced during the formation process of [32] S. Middya, A. Layek, A. Dey, P.P. Ray, J. Mater. Sci. Technol. 30 (2014) 770–775.
FeS2. Researches on the defects are also required to improve the [33] D. Wang, Q. Wang, T. Wang, CrystEngComm 12 (2010) 3797.
[34] L. Zhu, B. Richardson, J. Tanumihardja, Q. Yu, CrystEngComm 14 (2012) 4188.
photoelectric properties and hence enhance the photoelectric con- [35] J. Xia, J. Jiao, B. Dai, W. Qiu, S. He, W. Qiu, P. Shen, L. Chen, RSC Adv. 3 (2013)
version efficiency, which would make the practical application 6132.
possible. [36] T.S. Yoder, J.E. Cloud, G.J. Leong, D.F. Molk, M. Tussing, J. Miorelli, C. Ngo,
S. Kodambaka, M.E. Eberhart, R.M. Richards, Y. Yang, Chem. Mater. 26 (2014)
3. Materials properties can be enhanced by structural designing, which
6743–6751.
has been well confirmed in the 1D system. Therefore, it is necessary [37] S.-C. Hsiao, C.-M. Hsu, S.-Y. Chen, Y.-H. Perng, Y.-L. Chueh, L.-J. Chen, L.-H. Chou,
to design and create more abundant-morphology FeS2 structures, Mater. Lett. 75 (2012) 152–154.
which might bring unexpected and surprising properties. [38] N. E'Jazi, M. Aghaziarati, Adv. Powder Technol. 23 (2012) 352–357.
[39] W.L. Liu, X.H. Rui, H.T. Tan, C. Xu, Q.Y. Yan, H.H. Hng, RSC Adv. 4 (2014)
4. The ultimate goal for all studied materials is commercialization and 48770–48776.
serving for people and the society. Research on FeS2-based photo- [40] B. Yuan, W. Luan, S.-T. Tu, Dalton Trans. 41 (2012) 772–776.
electrochemical devices is still deficient, and the current photo- [41] J. Puthussery, S. Seefeld, N. Berry, M. Gibbs, M. Law, J. Am. Chem. Soc. 133
(2011) 716–719.
electric conversion efficiency is far from enough for actual appli- [42] Y. Bi, Y. Yuan, C.L. Exstrom, S.A. Darveau, J. Huang, Nano Lett. 11 (2011)
cation. Great effort still needs to be put into the application 4953–4957.
research. [43] W. Li, M. Döblinger, A. Vaneski, A.L. Rogach, F. Jäckel, J. Feldmann, J. Mater.
Chem. 21 (2011) 17946.
[44] S.C. Mangham, M. Alam Khan, M. Benamara, M.O. Manasreh, Mater. Lett. 97
Given the fast progress of the nanotechniques on synthesis and (2013) 144–147.
fabrication, more and more FeS2 nanostructures with encouraging [45] A. Douglas, R. Carter, L. Oakes, K. Share, A.P. Cohn, C.L. Pint, ACS Nano 9 (2015)
11156–11165.
properties and promising applications will be available shortly.
[46] B.J. Richardson, L. Zhu, Q. Yu, Sol. Energy Mater. Sol. Cells 116 (2013) 252–261.
[47] M.G. Gong, A. Kirkeminde, S.Q. Ren, Sci. Rep. 3 (2013).
Acknowledgements [48] A. Kirkeminde, B.A. Ruzicka, R. Wang, S. Puna, H. Zhao, S. Ren, A.C.S. Appl.
Mater. Interfaces 4 (2012) 1174–1177.
[49] H.A. Macpherson, C.R. Stoldt, ACS Nano 6 (2012) 8940–8949.
This work was supported by the Zhejiang Provincial Natural Science [50] D.Y. Wang, Y.T. Jiang, C.C. Lin, S.S. Li, Y.T. Wang, C.C. Chen, C.W. Chen, Adv.
Foundation of China (Nos. LY18B060005 and LR14B060002). Mater. 24 (2012) 3415–3420.
[51] J.M. Lucas, C.-C. Tuan, S.D. Lounis, D.K. Britt, R. Qiao, W. Yang, A. Lanzara,
A.P. Alivisatos, Chem. Mater. 25 (2013) 1615–1620.
References [52] M.A. Khan, J.C. Sarker, S. Lee, S.C. Mangham, M.O. Manasreh, Mater. Chem. Phys.
148 (2014) 1022–1028.
[1] I. Gur, N.A. Fromer, M.L. Geier, A.P. Alivisatos, Science 310 (2005) 462–465. [53] T.K. Trinh, V.T.H. Pham, N.T.N. Truong, C.D. Kim, C. Park, J. Cryst. Growth 461
[2] C. Wadia, A.P. Alivisatos, D.M. Kammen, Environ. Sci. Technol. 43 (2009) (2017) 53–59.
2072–2077. [54] M. Khabbaz, M.H. Entezari, J. Colloid Inter. Sci. 470 (2016) 204–210.
[3] T.A. Yersak, H.A. Macpherson, S.C. Kim, V.-D. Le, C.S. Kang, S.-B. Son, Y.-H. Kim, [55] S. Kar, S. Chaudhuri, Chem. Phys. Lett. 398 (2004) 22–26.
J.E. Trevey, K.H. Oh, C. Stoldt, S.-H. Lee, Adv. Energy Mater. 3 (2013) 120–127. [56] M. Nath, A. Choudhury, A. Kundu, C.N.R. Rao, Adv. Mater. 15 (2003) 2098-+.
[4] A. Ennaoui, S. Fiechter, C. Pettenkofer, N. Alonsovante, K. Buker, M. Bronold, [57] Y. Bai, J. Yeom, M. Yang, S.-H. Cha, K. Sun, N.A. Kotov, J. Phys. Chem. C 117
C. Hopfner, H. Tributsch, Sol. Energy Mater. Sol. Cells 29 (1993) 289–370. (2013) 2567–2573.
[5] G. Hodes, D. Cahen, Acc. Chem. Res. 45 (2012) 705–713. [58] M. Caban-Acevedo, D. Liang, K.S. Chew, J.P. DeGrave, N.S. Kaiser, S. Jin, ACS
[6] R. Wu, Y.F. Zheng, X.G. Zhang, Y.F. Sun, J.B. Xu, J.K. Jian, J. Cryst. Growth 266 Nano 7 (2013) 1731–1739.
(2004) 523–527. [59] M. Caban-Acevedo, M.S. Faber, Y.Z. Tan, R.J. Hamers, S. Jin, Nano Lett. 12 (2012)
[7] S.A. Sorenson, J.G. Patrow, J.M. Dawlaty, J. Phys. Chem. C 120 (2016) 1977–1982.
7736–7747. [60] D.Y. Wan, Y.T. Wang, Z.P. Zhou, G.Q. Yang, B.Y. Wang, L. Wei, Mater. Sci. Eng. B
[8] L.Y. Huang, F. Wang, Z.J. Luan, L.A. Meng, Mater. Lett. 64 (2010) 2612–2615. 122 (2005) 156–159.
[9] S. Seefeld, M. Limpinsel, Y. Liu, N. Farhi, A. Weber, Y.N. Zhang, N. Berry, [61] Y. Li, Z. Han, L. Jiang, Z. Su, F. Liu, Y. Lai, Y. Liu, J. Sol-Gel. Sci. Technol. 72
Y.J. Kwon, C.L. Perkins, J.C. Hemminger, R.Q. Wu, M. Law, J. Am. Chem. Soc. 135 (2014) 100–105.
(2013) 4412–4424. [62] M. Wang, C. Xing, K. Cao, L. Zhang, J. Liu, L. Meng, J. Mater. Chem. A 2 (2014)
[10] S. Nakamura, A. Yamamoto, Sol. Energ. Mat. Sol. C 65 (2001) 79–85. 9496–9505.
[11] A. Gomes, J.R. Ares, I.J. Ferrer, M.I.D. Pereira, C. Sanchez, Mater. Res. Bull. 38 [63] J. Liu, Y. Li, H. Fan, Z. Zhu, J. Jiang, R. Ding, Y. Hu, X. Huang, Chem. Mater. 22
(2003) 1123–1133. (2010) 212–217.
[12] Y.Z. Dong, Y. Zheng, H. Duan, Y. Sun, Y. Chen, Mater. Lett. 59 (2005) 2398–2402. [64] M. Wang, D. Xue, H. Qin, L. Zhang, G. Ling, J. Liu, Y. Fang, L. Meng, Mater. Sci.
[13] L. Hou, Y. Liu, L. Meng, J. Funct. Mater. 36 (2005) 1251–1256. Eng. B 204 (2016) 38–44.
[14] L. Reijnen, B. Meester, A. Goossens, J. Schoonman, J. Electrochem. Soc. 147 [65] R. Morrish, R. Silverstein, C.A. Wolden, J. Am. Chem. Soc. 134 (2012)

123
H. Qin et al. Materials Science & Engineering B 236–237 (2018) 104–124

17854–17857. 4431–4440.
[66] Q.-H. Huang, T. Ling, S.-Z. Qiao, X.-W. Du, J. Mater. Chem. A 1 (2013) (1833) [105] X.M. Song, J.M. Wu, L. Meng, M. Yan, J. Am. Ceram. Soc. 93 (2010) 2068–2073.
11828–11831. [106] Y. Xin, Z. Li, W. Wu, B. Fu, Z. Zhang, A.C.S. Sust, Chem. Eng. 4 (2016) 6659–6667.
[67] T.N. Murakami, M. Graetzel, Inorg. Chim. Acta 361 (2008) 572–580. [107] M. Wang, C. Chen, H. Qin, L. Zhang, Y. Fang, J. Liu, L. Meng, Adv. Mater. Inter. 2
[68] L. Li, M. Caban-Acevedo, S.N. Girard, S. Jin, Nanoscale 6 (2014) 2112–2118. (2015).
[69] J. Wu, L. Liu, S. Liu, P. Yu, Z. Zheng, M. Shafa, Z. Zhou, H. Li, H. Ji, Z.M. Wang, [108] I. Bedja, A. Hagfeldt, Adv. OptoElectronics 2011 (2011) 1–6.
Nano Lett. 14 (2014) 6002–6009. [109] Y. Lin, Y. Lin, Y. Meng, Y. Tu, X. Zhang, Opt Commun. 346 (2015) 64–68.
[70] X. Shi, A. Tian, X. Xue, H. Yang, Q. Xu, Mater. Lett. 141 (2015) 104–106. [110] G. Boschloo, A. Hagfeldt, Acc. Chem. Res. 42 (2009) 1819–1826.
[71] Y. Hu, Z. Zheng, H. Jia, Y. Tang, L. Zhang, J. Phys. Chem. C 112 (2008) [111] Y.C. Wang, D.Y. Wang, Y.T. Jiang, H.A. Chen, C.C. Chen, K.C. Ho, H.L. Chou,
13037–13042. C.W. Chen, Angew. Chem. Int. Ed. 52 (2013) 6694–6698.
[72] M.-L. Li, Q.-Z. Yao, G.-T. Zhou, X.-F. Qu, C.-F. Mu, S.-Q. Fu, CrystEngComm 13 [112] Q.-H. Huang, T. Ling, S.-Z. Qiao, X.-W. Du, J. Mater. Chem. A 1 (2013) 11828.
(2011) 5936. [113] Z. Wei, Y. Qiu, H. Chen, K. Yan, Z. Zhu, Q. Kuang, S. Yang, J. Mater. Chem. A 2
[73] Y. Tao, K. Rui, Z. Wen, Q. Wang, J. Jin, T. Zhang, T. Wu, Solid State Ionics 290 (2014) 5508.
(2016) 47–52. [114] J. Xu, H. Xue, X. Yang, H. Wei, W. Li, Z. Li, W. Zhang, C.S. Lee, Small 10 (2014)
[74] X.-H. Wang, T. Huang, F. Huang, S.-N. Zhang, Y. Wang, C. Gao, Spectrosc. Spect. 4754–4759.
Anal. 31 (2011) 2508–2511. [115] J. Liu, Y. Wen, Y. Wang, P.A. van Aken, J. Maier, Y. Yu, Adv. Mater. 26 (2014)
[75] J. Xia, X. Lu, W. Gao, J. Jiao, H. Feng, L. Chen, Electrochim. Acta 56 (2011) 6025.
6932–6939. [116] Y. Yamaguchi, T. Takeuchi, H. Sakaebe, H. Kageyama, H. Senoh, T. Sakai,
[76] L. Meng, Y.H. Liu, L. Tian, J. Cryst. Growth 253 (2003) 530–538. K. Tatsumi, J. Electrochem. Soc. 157 (2010) A630–A635.
[77] C. Delasheras, I.J. Ferrer, C. Sanchez, J. Phys.-Condens. Mat. 6 (1994) [117] D. Zhang, Y.J. Mai, J.Y. Xiang, X.H. Xia, Y.Q. Qiao, J.P. Tu, J. Power Sources 217
10177–10183. (2012) 229–235.
[78] L. Wu, N.Y. Dzade, L. Gao, D.O. Scanlon, Z. Öztürk, N. Hollingsworth, [118] R. Tan, J. Yang, J. Hu, K. Wang, Y. Zhao, F. Pan, Chem. Commun. 52 (2016)
B.M. Weckhuysen, E.J. Hensen, N.H. De Leeuw, J.P. Hofmann, Adv. Mater. 28 986–989.
(2016) 9602–9607. [119] L. Xu, Y. Hu, H. Zhang, H. Jiang, C. Li, ACS Sust. Chem. Eng. 4 (2016) 4251–4255.
[79] D.Y. Wan, B.Y. Wang, Y.T. Wang, H. Sun, R.G. Zhang, L. Wei, J. Cryst. Growth 257 [120] G.X. Pan, F. Cao, X.H. Xia, Y.J. Zhang, J. Power Sour. 332 (2016) 383–388.
(2003) 286–292. [121] D.T. Tran, H. Dong, S.D. Walck, S.S. Zhang, RSC Adv. 5 (2015) 87847–87854.
[80] B. Ouertani, J. Ouerfelli, M. Saadoun, H. Ezzaouia, B. Bessais, Thin Solid Films 516 [122] Y. Zhu, X. Fan, L. Suo, C. Luo, T. Gao, C. Wang, ACS Nano 10 (2016) 1529–1538.
(2008) 8584–8586. [123] F. Zhang, C. Wang, G. Huang, D. Yin, L. Wang, J. Power Sources 328 (2016)
[81] S. Huang, X. Liu, Q. Li, J. Chen, J. Alloys. Compd. 472 (2009) L9–L12. 56–64.
[82] D.Y. Wan, Y.T. Wang, B.Y. Wang, C.X. Ma, H. Sun, L. Wei, J. Cryst. Growth 253 [124] Z. Liu, T. Lu, T. Song, X.-Y. Yu, X.W. Lou, U. Paik, Energy Environ. Sci. 10 (2017)
(2003) 230–238. 1576–1580.
[83] Y.-H. Liu, Y. Wang, L. Meng, J. Inorg. Mater. 22 (2007) 143–147. [125] Z. Hu, Z. Zhu, F. Cheng, K. Zhang, J. Wang, C. Chen, J. Chen, Energy Environ. Sci.
[84] H. Duan, Y.F. Zheng, Y.Z. Dong, X.G. Zhang, Y.F. Sun, Mater. Res. Bull. 39 (2004) 8 (2015) 1309–1316.
1861–1868. [126] W. Chen, S. Qi, L. Guan, C. Liu, S. Cui, C. Shen, L. Mi, J. Mater. Chem. A 5 (2017)
[85] J.R. Ares, A. Pascual, I.J. Ferrer, C.R. Sanchez, Thin Solid Films 451 (2004) 5332–5341.
233–236. [127] W. Zhao, C. Guo, C.M. Li, J. Mater. Chem. A 5 (2017) 19195–19202.
[86] G. Baccarani, B. Ricco, G. Spadini, J. Appl. Phys. 49 (1978) 5565–5570. [128] R. Okuyama, H. Nakashima, T. Sana, E. Nomura, J. Power Sources 93 (2001)
[87] A.L. Efros, B.I. Shklovskii, J. Phys. C 8 (1975) L49–L51. 50–54.
[88] P.P. Altermatt, T. Kiesewetter, K. Ellmer, H. Tributsch, Sol. Energy Mater. Sol. [129] K. Sun, C.A. Cama, R.A. DeMayo, D.C. Bock, X. Tong, D. Su, A.C. Marschilok,
Cells 71 (2002) 181–195. K.J. Takeuchi, E.S. Takeuchi, H. Gan, J. Electrochem. Soc. 164 (2017)
[89] J. Oertel, K. Ellmer, W. Bohne, J. Rohrich, H. Tributsch, J. Cryst. Growth 198 A6039–A6046.
(1999) 1205–1210. [130] S.S. Zhang, D.T. Tran, J. Mater. Chem. A 4 (2016) 4371–4374.
[90] E. Bastola, K.P. Bhandari, R.J. Ellingson, J. Mater. Chem. C 5 (2017) 4996–5004. [131] M. Gong, A. Kirkeminde, Y. Xie, R. Lu, J. Liu, J.Z. Wu, S. Ren, Adv. Optical. Mater.
[91] I.J. Ferrer, J.R. Ares, C.R. Sanchez, Sol. Energy Mater. Sol. Cells 76 (2003) 1 (2013) 78–83.
183–188. [132] Z. Yang, L. Li, Y. Luo, R. He, L. Qiu, H. Lin, H. Peng, J. Mater. Chem. A 1 (2013)
[92] B. Koo, H. Jung, M. Park, J.-Y. Kim, H.J. Son, J. Cho, M.J. Ko, Adv. Funct. Mater. 954–958.
26 (2016) 5400–5407. [133] M. Gong, A. Kirkeminde, N. Kumar, H. Zhao, S. Ren, Chem. Commun. 49 (2013)
[93] N. Takahashi, T. Sawada, T. Nakamura, T. Nakamura, J. Mater. Chem. 10 (2000) 9260–9262.
2346–2348. [134] S. Liu, M. Li, S. Li, H. Li, L. Yan, Appl. Surf. Sci. 268 (2013) 213–217.
[94] A.K. Raturi, S. Waita, B. Aduda, T. Nyangonda, Renew. Energy 20 (2000) 37–43. [135] A. Tian, Q. Xu, X. Shi, H. Yang, X. Xue, J. You, X. Wang, C. Dong, X. Yan, H. Zhou,
[95] N. Takahashi, Y. Nakatani, T. Yatomi, T. Nakamura, Chem. Mater. 15 (2003) RSC Adv. 5 (2015) 62724–62731.
1763–1765. [136] G. Lee, M. Kang, Curr. Appl. Phys. 12 (2013) 1482–1489.
[96] X. Liu, J.Y.L. Ho, M. Wong, H.S. Kwok, Z.J. Liu, RSC Adv. 6 (2016) 8299–8305. [137] C. Guo, X. Tong, X.-Y. Guo, Mater. Lett. 161 (2015) 220–223.
[97] T. Biegler, D.A. Swift, Electrochim. Acta 24 (1979) 415–420. [138] M.S. Faber, M.A. Lukowski, Q. Ding, N.S. Kaiser, S. Jin, J. Phys. Chem. C 118
[98] E. Peters, H. Majima, Can. Metall. Quart. 7 (1968) 111–117. (2014) 21347–21356.
[99] L.K. Bailey, E. Peters, Can. Metall. Quart. 15 (1976) 333–344. [139] D.-Y. Wang, M. Gong, H.-L. Chou, C.-J. Pan, H.-A. Chen, Y. Wu, M.-C. Lin,
[100] T. Nagai, H. Kiuchi, J. Inst. Min. Met. Japan 91 (1975) 547. M. Guan, J. Yang, C.-W. Chen, Y.-L. Wang, B.-J. Hwang, C.-C. Chen, H. Dai, J. Am.
[101] B.F. Giannetti, S.H. Bonilla, C.F. Zinola, T. Raboczkay, Hydrometall. 60 (2001) Chem. Soc. 137 (2015) 1587–1592.
41–53. [140] D. Jasion, J.M. Barforoush, Q. Qiao, Y. Zhu, S. Ren, K.C. Leonard, ACS Catal. 5
[102] Y. Liu, Z. Dang, P. Wu, J. Lu, X. Shu, L. Zheng, Ionics 17 (2010) 169–176. (2015) 6653–6657.
[103] J.C. Kim, J. Choi, Y.B. Lee, J.H. Hong, J.I. Lee, J.W. Yang, W.I. Lee, N.H. Hur, [141] C. Di Giovanni, A. Reyes-Carrnona, A. Coursier, S. Nowak, J.M. Greneche,
Chem. Commun. (2006) 5024–5026. H. Lecoq, L. Mouton, J. Roziere, D. Jones, J. Peron, M. Giraud, C. Tard, ACS Catal.
[104] S. Shukla, G. Xing, H. Ge, R.R. Prabhakar, S. Mathew, Z. Su, V. Nalla, 6 (2016) 2626–2631.
T. Venkatesan, N. Mathews, T. Sritharan, T.C. Sum, Q. Xiong, ACS Nano 10 (2016)

124

You might also like