Open Access
3 February 2020 Bridging the p-type transparent conductive materials gap: synthesis approaches for disperse valence band materials
Angela N. Fioretti, Monica Morales-Masis
Author Affiliations +
Abstract

Transparent conductive materials (TCMs) with high p-type conductivity and broadband transparency have remained elusive for years. Despite decades of research, no p-type material has yet been found to match the performance of n-type TCMs. If developed, the high-performance p-type TCMs would lead to significant advances in a wide range of technologies, including thin-film transistors, transparent electronics, flat screen displays, and photovoltaics. Recent insights from high-throughput computational screening have defined design principles for identifying candidate materials with low hole effective mass, also known as disperse valence band materials. Particularly, materials with mixed-anion chemistry and nonoxide materials have received attention as being promising next-generation p-type TCMs. However, experimental demonstrations of these compounds are scarce compared to the computational output. One reason for this gap is the experimental difficulty of safely and controllably sourcing elements, such as sulfur, phosphorous, and iodine for depositing these materials in thin-film form. Another important obstacle to experimental realization is air stability or stability with respect to formation of the competing oxide phases. We summarize experimental demonstrations of disperse valence band materials, including synthesis strategies and common experimental challenges. We end by outlining recommendations for synthesizing p-type TCMs still absent from the literature and highlight remaining experimental barriers to be overcome.

1.

Introduction

A longstanding challenge in optoelectronic materials is the experimental demonstration of high-performance p-type transparent conductive materials (p-TCMs), exhibiting properties on par with their n-type counterparts. Development of a p-TCM with broadband transparency and high conductivity that can be reliably applied in thin-film form would mark a significant breakthrough in the fields of transparent electronics, solar cells, and LEDs by enabling devices with reduced process complexity and minimum parasitic losses at the contacts.13 Figure 1 shows example architectures that would benefit from the development of high-quality p-TCMs, including a silicon heterojunction solar cell,4 a semitransparent halide perovskite solar cell or LED,5,6 and a transparent complementary circuit built of p-type and n-type thin-film transistors (TFTs).7,8

Fig. 1

Schematic diagrams of optoelectronic devices that will benefit from the development of p-type TCMs. (a) Silicon solar cells and (b) semitransparent halide perovskites solar cells or LEDs: a p-type TCM will allow to reduce parasitic absorption at the (front) contacts of the cell and could simplify processes such as the replacement of the hole selective contact and the TCO with one single layer; the p-type TCM providing hole selectivity and lateral transport. (c) Complementary electronics build of p- and n-type TFTs: p-type TCMs will allow the realization of full transparent electronics.

JPE_10_4_042002_f001.png

The difficulty in experimentally realizing excellent p-TCMs lies in a contradiction of the fundamental requirements: p-type conductivity requires high hole mobility and low ionization potential to allow for p-type doping, but transparency requires a wide bandgap, which is most often found in metal oxides. The valence band maximum (VBM) of metal oxides is typically composed of highly localized oxygen 2p orbitals, which results in large hole effective masses and very low hole mobility. Furthermore, metal oxides are usually n-type with high ionization potential that hampers efforts on p-type doping. Cu-based delafossites and Cr-based oxides are some of the few oxides with experimentally reported p-type conductivity.9,10 The first p-TCM, delafossite CuAlO2 discovered in 1997, had a free hole carrier density of Nh=1.3×1017  cm3 and a hole mobility of 10  cm2V1s1 resulting in a conductivity <1  Scm1.11 Recently, Mg-doped CuCrO2 was reported with a conductivity of 220  Scm1, representing the highest p-type conductivity reported to date.12 However, this high p-type conductivity has thus far not been reproduced by other groups.13 Moreover, even with 220  Scm1, the performance still falls short of standard n-type TCMs that routinely display conductivity >300  Scm1.3,14

With the advent of high-throughput computational screening and design approaches, numerous research articles proposing material compositions and structures with p-type conductivity and wide band gaps have been published. These predictions are based on coupling dispersion (delocalization) of the valence band (VB) with favorable defect formation chemistry in order to select materials with inherently high hole mobility and enhanced p-type dopability.1518 However, experimental realization of these materials currently lags behind theoretical predictions, creating an evident gap between computationally predicted disperse valence band materials (DVMs) and their experimental investigation. In this review, we present an overview of the experimental techniques and synthesis approaches reported on DVMs with the aim to bridge the gap between computational and experimental works on p-TCMs. The authors note that while oxide materials with promising p-type conductivity and transparency properties have received attention in the recent computational literature,14,19 we focus herein on materials beyond oxides, as oxides are considered well-known chemistries with several synthesis approaches already at advanced stages of development.

2.

Computational Screening for Disperse Valence Band Materials

VB dispersion as a design principle for p-TCMs has existed since the late 1990s.11 Termed “chemical modification of the valence band,” this strategy relies on orbital-mixing between Cu s-orbitals and O p-orbitals at the VBM to delocalize the VB and produce a shallow ionization potential thus facilitating p-type doping without self-compensation.20 A growing number of computational works have extended this strategy by screening databases of inorganic materials for low hole effective mass (mh*) and p-type dopability, and using the resulting material set to deduce design principles for identifying candidate high-performance p-TCMs.14,16,17

Figure 2 gives a visual depiction of VB dispersion, including the relationship among VB dispersion (d2E/dk2), hole effective mass (mh*), and hole mobility (μh). A sharply curved VB has a large value of d2E/dk2 and hence a small mh* and large μh. For equal free hole carrier densities (Nh), the highest conductivity (σh) is achieved at highest (μh). Importantly, in Eq. (2) in Fig. 2, it is clear that μh depends on two variables: mh* and the scattering time (τ) (e is the fundamental electron charge). While the design principles here discussed are focused on lowering mh* to enhance μh, it is important to note that the crystal lattice and film microstructure will also play an important role in μh through τ. τ depends on the collisions of the conduction holes (or conduction electrons in n-type materials) with lattice phonons (τph), impurity atoms (τi), and grain boundaries or any other film imperfection (τGB). As the rates of these collisions are—to a good approximation—independent, the net relaxation time is given by: 1/τ=1/τph+1/τi+1/τGB. More details regarding the scattering mechanisms are found in Ref. 21.

Fig. 2

(a) Schematic diagram depicting differing degrees of VB dispersion and an arbitrary conduction band. E represents the energy axis and k is the wave vector. VB1 represents the least disperse (most localized) VB with the highest hole effective mass (mh1), whereas VB3 represents the most disperse (least localized) VB with the lowest hole effective mass (mh3). This relationship is shown in writing in panel (b), where the hole effective mass of each is included in an inequality, along with expressions of hole effective mass (mh*), hole mobility (μh), and p-type conductivity (σh). The inverse proportionality between band dispersion (d2E/dk2) and hole effective mass as well as hole effective mass and hole mobility are represented in Eqs. (1) and (2), respectively.

JPE_10_4_042002_f002.png

In particular, VB delocalization driven by anion-site chemistry was highlighted as leading to increased curvature (i.e., more disperse VBs) and thus low hole effective mass. One design principle identifies materials with mixed-anion chemistry, for example, oxy-chalcogenides or oxy-pnictides. In this case, the oxygen 2p-orbitals at the VBM are mixed with more disperse 3p-orbitals from the chalcogenide or pnictide anions, leading to better hole conductivity while maintaining transparency imparted by the oxide.16 Another material class identified is binary nonoxide compounds.15,18 These materials possess a VBM with significant contributions from anion p-states, which for nonoxide anions are more delocalized than the 2p orbitals of oxygen.18 Materials such as sulfides, iodides, and especially phosphides13,15,16,19,22 were found via high-throughput screening and first-principles calculations to commonly have mh*<5  me, with phosphides including many examples <1  me, owing to the metal-p and phosphorous-p orbital mixing at the VBM.15

Notable p-TCM candidates in these material classes are boron phosphide (BP),15 copper iodide (CuI),2224 CuAlS2,25,26 (Zr,Hf)OS,16,27 and Mg:LaCuOSe.28 The band gap of BP can be up to 4 eV (direct)15 with an indirect gap measured experimentally to be 2 eV.29,30 CuI has a bandgap of 3.1 eV, a calculated hole effective mass of 2  me, and native copper vacancy point defects that lead to degenerate p-type self-doping.22,31,32 CuAlS2 exhibits p-type conductivity >500  Scm1 and a bandgap of 3.2 eV,33 and Mg:LaCuOSe has a reported bandgap of 2.8 eV and hole effective mass of 1.6  me.28 Despite these desirable properties, including some that are experimentally demonstrated, many challenges to experimental realization and/or widespread adoption of these materials exist, including difficulties in safely controlling precursors, such as sulfur and phosphorous, or avoiding secondary phases. The following sections are divided broadly into two material classes, nonoxides and oxychalcogenides, and subdivided further by specific compound. Experimental demonstrations of DVMs for p-TCM applications will be summarized and challenges particular to each chemistry will be discussed. Suggestions for overcoming these experimental roadblocks will be proposed, along with recommendations for future synthesis strategies and ultimate device integration.

3.

Experimental Demonstrations

3.1.

Nonoxide Materials

Two of the most widely studied nonoxide p-TCMs are BP and CuI. Experimental reports for BP extend back to 195734 and for CuI extend even further back to 1901.35 P-type conduction was measured in CuI as early as 1908, before the conception of positive-charge carriers was defined36 whereas BP was not identified as p-type until 1960.37 The first use of CuI as a p-type transparent layer in an optoelectronic device was published only in 2002,38 and the first mention of it as a transparent conductor occurred in 2003.39 For BP, the first mention of its potential as a p-type transparent conductor only occurred in 2017,15 although previous works do report p-type character alone. As this review intends to focus on the experimental demonstrations and challenges of DVMs as p-type transparent conductors, the remaining discussion will focus solely on the more recent reports in which synthesis and characterization for p-TCM application was the main research goal.

3.1.1.

Copper iodide

Application of CuI as a transparent hole-transport layer began in the dye-sensitized solar cell field,3840 and later continued in the fields of organic photovoltaics4148 and hybrid perovskite PV.49 For a full review of the CuI literature, see Ref. 22. The first demonstration of p-type, transparent, and conducting CuI in thin-film form was vapor iodination of a Cu2S precursor film.50 Vapor iodination is one of the most common strategies for CuI thin-film synthesis, in addition to thermal evaporation.5153 The next most popular method is some variation of solution-coated synthesis, such as chemical bath deposition (CBD),54 spin-coating,55,56 or other similar strategies.57 Vacuum deposition methods have been used, including pulsed-laser deposition39,40,5862 and reactive sputtering in iodine vapor,63,64 but these techniques have received considerably less attention. In many cases, researchers simply purchase CuI powder and evaporate it onto glass substrates,46,6567 which highlights the ease of depositing CuI films. Such simplicity in synthesis has arguably led to the proliferation of CuI applications in optoelectronic devices over the years. This observation underlines an important point about experimental development of DVMs in general, namely that facile synthesis strategies are key to a given material’s experimental success.

3.1.2.

Boron phosphide

Contrary to the case of CuI, BP has only been applied as a p-type layer in an optoelectronic device in a few limited cases. The first mention of p-type BP in thin-film form was grown by chemical vapor deposition (CVD) using diborane and phosphine on a silicon substrate,68 in which Schottky barrier diodes were prepared from p-BP/Al. Soon after, p-type BP applied in a photoelectrochemical cell for hydrogen evolution reaction.69 Already at this stage, it was reported that BP could be doped either n-type or p-type, depending on the B:P ratio in the final material.6872 After these initial investigations into the electrical properties and p-type doping of BP, no reports of the p-type material occur until much later.7375

Despite this absence of electrical characterization, many studies exist demonstrating BP synthesis by a variety of methods that include structural and optical characterization. One of the most common ways to make BP is through bulk synthesis techniques, such as flux growth or solid-state reaction.30,7679 Thin-film deposition of BP is almost invariably carried out by CVD of some kind, either metal-organic chemical vapor deposition (MOCVD),73,80,81 or standard CVD,68,7072,74,75,8285 or even plasma-enhanced CVD.86 Other, less common deposition methods include vapor–liquid–solid growth,87,88 thermal evaporation from powders in vacuum,89 or sputtering in phosphine/Ar atmosphere.90

One observation from these experimental demonstrations is that the value of the direct bandgap of BP is still somewhat controversial. While some experimental reports indicate an indirect gap at 2 eV and a direct gap at 2.5 eV direct,30 other work shows a direct gap as high as 6 eV,37 and theoretical predictions indicate the direct gap could be 4 eV.15 Experimental confirmation of the value of the direct bandgap in BP is one area where more research is needed going forward. In addition, while the level of p-type doping achievable in the literature (1017 to 1019  holes/cm3)68,73 supports the potential applicability of BP as a p-TCM, the overall dearth of electrical characterization for this material is another significant gap in the literature and represents an excellent opportunity for new experimental investigations.

3.1.3.

Copper aluminum sulfide

Chalcopyrite copper aluminum sulfide (CuAlS2) is another wide bandgap (3.2  eV)33,91 p-type material exhibiting a disperse VB that has been shown experimentally to have desirable electrical properties for TCM applications. Experimental demonstrations of this material have reported facile self-doping up to 1021  cm3 coupled with a low hole effective mass estimated at 0.33  me, due to the hybridization of Cu 3d and S 3p orbitals at the VBM.25 These features lead to high p-type conductivity in CuAlS2 in the range of 250 to 540  S/cm.25,33 Thin films of CuAlS2 have been applied as a transparent p-type contacts in CIGS and organic solar cells,33,92 but the material has otherwise received relatively limited attention in the literature. Despite the smaller number of studies, CuAlS2 has been synthesized by a wide variety of deposition techniques, including solution-based techniques, such as CBD,33,93 CVD methods, such as atomic layer deposition (ALD) and MOCVD,91,92,94 solid-state synthesis routes,25 thermal evaporation or sulfurization,9598 and even spray pyrolysis.99,100 One important finding from these investigations is Cu-rich composition leading to higher conductivity,26,33 due in part to the VB shifting upward with increasing copper content.26 This decrease in ionization potential has been shown via first principles calculations to result in the CuAl acceptor transition level becoming more shallow, thus yielding lower ΔHf and better doping efficiency.26

3.1.4.

Copper zinc sulfide

Cu-alloyed zinc sulfide (CuxZn1xS), the final nonoxide material covered in this review, represents a special case, as it can be synthesized either as a heterostructural alloy101 or as a nanocomposite of Cu2S and ZnS, in which high conductivity stems from the Cu2S phase and transparency is imparted by the wide bandgap of ZnS.102 In the case of the alloy, p-type conductivity up to 40  Scm1 and a bandgap of 3.1 eV have been reported.103,104 For the nanocomposite, conductivity up 1000  Scm1 has been shown in Refs. 102, 105, and 106 but at the cost of transparency with optical absorption onset lowering to 2.3 to 2.5 eV.107,108 CuxZn1xS can be deposited using vacuum methods such as pulsed layer deposition (PLD) and sputtering,102,104,106 which provide the nonequilibrium growth conditions necessary to obtain the metastable heterostructural alloy. The nanocomposite can be also synthesized at ambient temperature via CBD102,105,109,110 or by electrodeposition.107,108,111 Of the nonoxide materials discussed, CuxZn1xS has received the least experimental attention in the literature, perhaps due to challenges in controlling phase separation to achieve the desired properties as either a nanocomposite or alloy.

3.2.

Mixed-Anion Materials

These material systems simultaneously contain an oxide (O2) and a chalcogenide (S2, Se2, and Te2) anion. Systems range from quinary compounds (e.g., doped, layered oxysulfides, such as Sr-doped LaCuOS) to simpler ternary compounds (e.g., ZrOS). Among experimentally realized p-TCMs, layered oxysulfides and oxyselenides are the most reported in the literature to date.

3.2.1.

Multinary oxy-chalcogenides

The p-type conductivity of LaCuOS prepared by solid-state reaction was first demonstrated in 1991.112 Earlier, the material was also prepared through oxidation of LaCuS2 but only the crystal structure was reported.113 It was not until the year 2000 that the combination of transparency (70% transmittance) and p-type conductivity (0.26  Scm1) was reported in the layered Sr-doped LaCuOS films prepared by radio-frequency sputtering.114 The sputtering target was prepared as a composite of stoichiometrically mixed La2S3, Cu2S, and La2O3 via solid-state reaction. Deposition was performed at 400°C, but crystallization was not achieved until a postannealing step at 800°C was applied.

Mg-doped LaCuOSe thin films were epitaxially grown by PLD.28,115 Films were deposited on MgO (001) substrates from a La0.8Mg0.2CuOSe ceramic target. A postannealing step at 1000°C was applied to crystallize the film. Very high hole carrier density was reported for this material, in the range of 1021cm3; however, hole mobility was on the order of 3.4  cm2/Vs and an effective mass of 1.6±0.2  me (me is the rest mass of the electron) was estimated by the analysis of the free carrier absorption using the Drude model. A band gap of 2.8 eV was reported; however, the films presented high absorbance in the visible range of the spectra (estimated 50% transmission at 600 nm wavelength). After this report, hybrid density functional theory was used to suggest alternative p-type dopants for LaCuOSe.116 Sr2+ or Ca2+ was identified as optimal acceptor dopants instead of the previously reported Mg2+. Experimental validation of these acceptors has not yet been reported following this computational work.

Finally, the layered oxysulfide [Cu2S][Sr3Sc2O5] was synthesized (only in bulk form) and reported to have a hole mobility of 150  cm2/Vs. However, p-type carrier density was low (order of 1017  cm3) resulting in a conductivity of only 2.8  S/cm.117 While the mobility was very promising, no report of thin-film synthesis of this material has been reported to date.

3.2.2.

Ternary oxy-chalcogenides

P-type Y-doped ZrOS nanocrystals were synthesized by high-temperature solid-state reaction.27 The design principle consisted of synthesizing tetragonal ZrOS (t-ZrOS), as calculations indicated that the S 3px/y orbitals form a shallower and sharper VBM (both requirements for p-type dopability and high hole mobility) compared to cubic ZrOS (c-ZrOS), although the bandgap of the cubic phase was predicted to be significantly wider. Hole effective masses of 0.24  me and 0.37  me were calculated for t-ZrOS and c-ZrOS, respectively. Conductivity of 102  S/cm was determined by Seebeck coefficient measurement; however, no Hall effect measurements were reported as reliable measurements require the fabrication of thin films. It is therefore the fabrication of thin films and their optical and transport property measurements that will confirm the potential of this material.

Thin films of HfOS have not been reported in literature to our knowledge, but interestingly, it has been demonstrated that controlling the oxygen vacancies in HfO2x results in a formation of a semimetal with p-type conductivity.118 Hole carrier density on the order of 1021  cm3 was reported for films fabricated with molecular beam epitaxy, yet the band gap was too low for TCM applications (1  eV).118 Moreover, in 2013, first principles calculations were used to propose tetragonal Hf2O3 and Zr2O3 as semimetals with high hole carrier density as well.119 These reports and others focused strictly on depositing the related oxides could serve as inspiration for demonstrating thin-film HfOS and ZrOS using similar approaches.

Tin oxyselenide (SnOSe) thin films with p-type conductivity have recently been reported in the experimental literature.120 Thin films were prepared by reactive co-sputtering from Sn and SnSe targets under oxygen atmosphere (substrate temperature not reported). After deposition, the films were annealed at 300°C under vacuum. Annealed films with the composition SnSe0.56O0.44 exhibited hole mobility of 15  cm2/Vs and hole carrier density of 1.2×1017cm3. The band gap (1.93 eV) is however too low to become a state of the art p-TCM. Experimental efforts would need to focus on developing strategies to increase doping and widen the band gap in order to further assess the usefulness of this material system for p-TCM applications.

Table 1 summarizes selected properties of the DVMs covered herein, separated by whether or not the materials have been experimentally demonstrated in the literature. We caution the reader that conductivity values must always be weighed against transparency, as not every material with high conductivity in Table 1 exhibits high transparency (e.g., BP). A good indication for transparency in the visible range is the band gap (i.e., Eg>3  eV). However, several of the reported Eg in the table are from computational studies, and therefore not conclusive in terms of “real” electrode performance. Successful synthesis methods for each material are also listed, which is a key aspect of this review. One aim of the discussion in this work is to assess whether particular synthesis methods, chemistries, or other features tend to facilitate experimental demonstration of DVMs. Looking at Table 1, it is clear that no one particular synthesis method wins. Instead, chemical complexity appears to be the more important factor in deciding whether a given material has received extensive experimental attention in the literature or not.

Table 1

Summary of properties and synthesis methods for DVMs.

MaterialHole effective massBandgap (eV)Demonstrated? (Y/N)Mobility (cm2/Vs)Conductivity (S/cm)Deposition Method(s)
CuI0.3 (lh) 2.14 to 2.4 (hh)31,323.122Y6, 43.9121,1220.3 to 94.336,121124Iodination of thin film, PLD, solid-state synthesis, solution-processing, sputtering, thermal evaporation
BP0.35152.0 (ind), 2.5 to 4.0 (dir), 6.0 (dir)15,30,37Y1.77, 7073,1250.67 to 70737,73,125,126CVD, MOCVD, PECVD, solid-state synthesis, sputtering, thermal evaporation, vapor–liquid–solid
Cu-Zn-S2.3 to 3.8102,107109,111Y1.0 to 1.4, 0.5 to 1.6102,10642, 54, 752102,104,106CBD, PLD, sputtering
CuAlS20.33252.88 to 3.425,33,91Y1.52, 21.225,33250, 54625,33ALD, CBD, MOCVD, solid-state synthesis, spray pyrolysis, sulfurization of metal, thermal evaporation
LaCuOS11273Y0.012114Solid-state synthesis, sputtering
La0.95Sr0.05CuOSY0.26114Solid-state synthesis, sputtering
Mg:LaCuOSe1.6 ± 0.2282.828Y3.42895028PLD with postannealing
[Cu2S][Sr3Sc2O5]3.1117Y1501172.8117Solid-state synthesis
ZrOS (tetragonal)0.24272.5 (dir. forbidden)27Y0.0127Solid-state synthesis
SnOSe0.6 to 1.991201.93120Y151200.29120Sputtering
Computationally identified only
ZrOS (cubic)0.37274.25 (ind)16N
HfOS1.25164.5N
La2SeO20.921283.49N

4.

Experimental Challenges

4.1.

Nonoxide Materials

4.1.1.

Copper iodide

The experimental challenges faced by CuI are not in the synthesis itself, which, as shown above, is straightforward and can be carried out by a number of methods. On the contrary, the main challenges are in the properties of the layer after deposition. Air and moisture stability of is a well-documented issue,51,129 due at least in part to the low migration energy barrier for Cu, especially at temperatures above 200°C.130132 Another problem relevant to its application as a transparent conductor is high surface roughness when synthesized using the common method of vapor iodination of Cu films or by PLD,52,58,121 which reduces the transparency significantly. However, thermal evaporation133,134 and sputtering techniques63,64 do successfully address the surface roughness problem, in addition to vapor iodination using a compound precursor film, such as Cu3N or Cu2S.52,135 Despite these challenges, CuI is one of the most widely studied DVMs in the literature today, owing to a few factors, one important factor being the ease with which the material can be synthesized in phase-pure form. Indeed, the absence of competing secondary phases and the propensity to self-dope are two features of CuI that continue to facilitate its investigation and application in the literature.

4.1.2.

Boron phosphide

In the case of BP, the biggest experimental challenge is the highly inert nature of elemental boron,78 which has led researchers to rely on deposition techniques that utilize more reactive forms of boron—specifically, diborane (B2H2) in CVD-based processes. The drawback to this approach is the high toxicity of both diborane and phosphine,86 which require special safety measures for use. Toxicity of phosphine is also an obstacle for attempts at sputtering BP,90 which increases the reactivity of boron via plasma-based deposition but thus far has utilized a gas-phase phosphorous source. One possible alternative could be a solid phosphorous source for evaporation in vacuum in conjunction with a boron sputtering target, although careful management of side reactions with phosphorous and water or air would still need to be undertaken.

Keeping the above concerns in mind, it is still the case that successful thin-film deposition of BP has been demonstrated in the literature for years. What is lacking in the literature are studies focused on electrical characterization of p-type BP. Given that several reports of p-type conductivity already exist,68,71,7375 and that this material has been computationally identified as having a disperse VB suitable for high mobility of holes,15 all that remains is to couple known deposition recipes with standard electrical characterization techniques to experimentally confirm the computational predictions.

4.1.3.

Sulfides

As with any sulfide, experimental challenges to synthesizing CuAlS2 or Cu-alloyed ZnS thin films arise mostly related to sulfur management, in particular the need to avoid toxicity from hydrogen sulfide (H2S). This is the case in any vacuum-based method, CVD-based deposition, or thermal evaporation, as precursors used to deliver sulfur to the growing film typically lead to unintentional formation (or incomplete utilization) of H2S.91,92 For this reason, solution-based methods are more attractive since sulfur-containing reagents can be more readily controlled. However, CBD can lead to inhomogeneity in the resulting film,33 and often such films require postdeposition annealing to obtain good crystallinity and the desired TCM properties, which leads again to the issues of sulfur management. However, the challenges presented by solution synthesis may be easier to overcome by a wider range of researchers than those presented by CVD methods, given that less infrastructure is required to properly trap or ventilate sulfur-containing by-products of an anneal than in the case of direct H2S delivery to a deposition chamber.

4.2.

Mixed-Anion Materials

Some of the challenges of growing mixed-anion materials lie in the different formation energies for oxides versus chalcogenides and in the metastability of some phases, resulting in phase segregation. Therefore, synthesis approaches promoting chemical stabilization should be preferred. In literature, solid-state synthesis has been the most common technique used for making oxychalcogenide bulk crystals.136 But this technique requires high thermal budgets, is a lengthy process, and is not well-suited to optoelectronic device integration. However, reported recipes can be applied for the fabrication of sputtering or PLD targets, through powder milling and pressing. Once a target is created, a single-source deposition should be possible with these techniques.

For the formation of mixed-anion thin films, PVD (evaporation, sputtering, and PLD), and CVD (ALD and MOCVD) offer several advantages, including their compatibility with device fabrication. Differing volatility or sputtering rates of the various species in mixed-anion materials can lead to challenges in single-source deposition. To overcome such challenges, multisource deposition can be applied, such as thermal coevaporation or cosputtering. Multistep processes can also be an effective synthesis strategy in making multinary oxychalcogenides, such as LaCuOS,114 in which initial deposition is followed by postdeposition annealing in reactive or nonreactive atmospheres and allows control of the phases formed at each processing step.137,138 In the case of oxychalcogenides, controlled oxidation could be a third synthesis step. It is important to note that the order sulfurization (selenization) + oxidation or oxidation + sulfurization (selenization) should be determined based on the formation energy and binding energy of the competing binary phases. For example, to fabricate ZrOS films using a multistep process, formation of ZrSx must be undertaken first (in an oxygen-free atmosphere), as the formation energy of ZrO2 is much lower than that of ZrS2 (by 1.8eV).139 Once a sulfide phase is formed, oxidation can then be performed by, for example, annealing in controlled oxygen atmosphere as well as heating at a controlled rate to avoid the segregation of ZrOx and ZrSx phases. The same recommendations apply for HfOS. Cosputtering from compound targets, such as ZrS2 and ZrO2 for ZrOS or HfS2 and HfO2 sources for HfOS, is another promising method for synthesizing the ternary phases without segregation of binary compounds.

ALD and PLD are also promising techniques for these oxychalcogenide compounds. ALD allows the alternating introduction of precursors, such as H2S, water, and metal–organic precursor gases, and gives atomic-level control to build these compounds element by element.140 PLD on the other hand has the advantage that the use of a high energy laser allows nonequilibrium ablation of a target (containing all the required elements of the multicomponent film) and therefore stoichiometric transfer is possible independent of the volatility of the components.141 It is important to note that PLD has commonly been used for epitaxial growth of films at high temperature. However, PLD has a great potential for nonepitaxial low-temperature fabrication of thin films, due to the possibility of enabling single-source deposition of multicomponent materials.

Up to now, we have only discussed vacuum-based deposition techniques for mixed-anion materials. Beyond vacuum-based processes, solution processing, such as single-step hydrothermal synthesis and two-step CBD, has proven successful to access selenide and oxy-selenide phases at low temperatures.142 The accessibility of these approaches makes them worth exploring as well, in order to evaluate the feasibility of using low-cost solution synthesis for integration of oxy-chalcogenide p-TCM materials in optoelectronic devices.

5.

Summary and Outlook

One general similarity among the DVMs discussed herein that have been experimentally demonstrated as a thin film is that each could be deposited by a variety of techniques, ranging from vacuum-based deposition to solution-coating. In each case, the nonoxide anion (sulfide, selenide, phosphide, or iodide) presents challenges to safe delivery during film growth that are not similarly present when dealing with oxides. These challenges must be handled differently depending on the deposition technique chosen, but no particular method of thin-film synthesis is consistently superior, so long as sufficient resources are dedicated to developing a finely controlled process. Researchers interested in investigating one of the nonoxides or oxychalcogenides discussed in this review can apparently take their pick of deposition methods for which their lab is best equipped, given the sheer number of different techniques represented in the literature thus far. Overall, the important point is to recognize that thin-film synthesis is critically enabling for electrical characterization and ultimate device integration for candidate p-TCMs and should be the focus of experimental efforts going forward.

Looking more closely at publication trends among the nonoxide and oxychalcogenide materials, a few more observations can be made. The total number of publications for each material decreases going from CuI to BP to CuAlS2 and finally to Cu-Zn-S and the oxychalcogenides. One observation is that the two materials with the most publications are both binary compounds, whereas the ternary and multinary materials have received significantly less research attention. A possible explanation for this is the relatively simpler phase space of binary compounds compared to multinary materials, which facilitates synthesis and characterization. Another observation is that materials, such as Cu-Zn-S and [Cu2S][Sr3Sc2O5], are known to exhibit Cu2S phase segregation, which renders them more difficult to accurately characterize, despite the fact that the nanocomposite of Cu-Zn-S in particular has been shown to exhibit desirable TCM properties. In general, these observations suggest that focusing research attention on binary and ternary materials with fewer (or less favorable to form) secondary phases, as opposed to the more-complicated quaternary and quinary compounds, is a prudent investigative strategy to bridge the p-TCMs gap.

Moreover, we propose that computational screening and first principles predictions of p-TCMs attempt to translate chemical potential diagrams into practical experimental considerations such as partial pressures or material fluxes in a reaction chamber. An example of such translation can be seen for La2SeO2 in Refs. 116 and 128. For the experimentalist, it is important to build a good understanding between reported chemical potentials at which phases are stable, and partial pressures during the growth of the phases.

One final point is the absence of experimental confirmations of predicted properties for many of the materials discussed herein. For materials such as BP, reports of electrical properties are scarce, and the absence of thin film examples for some multinary oxychalcogenides has thus far precluded typical characterization campaigns for transparency and conductivity. Another important characterization that is within reach for experimentalists today is visualizing VB dispersion using techniques such as angle-resolved photoemission spectroscopy. Applying this measurement to predicted DVMs to assess the curvature of the VB as composition changes for a given material (especially those with clear binary endpoints, such as CuAlS2 and ZrOS) would be a fascinating and informative research thrust. A summary of the recommendations distilled from this review is depicted graphically in Fig. 3. Ultimately, it will be shared knowledge between computational materials science and thin film synthesis supported by state-of-the-art characterization techniques that will allow us to correlate experiment with predictions and finally close the p-TCMs gap and integrate the best material into optoelectronic devices for maximum impact.

Fig. 3

Summary of the four recommendations for bridging the experimental literature gap in p-TCMs. Materials with reduced chemical complexity deposited in thin-film form should be the focus of experimental research efforts going forward. Investigating new candidates for p-type dopants in previously synthesized materials and working to document conductivity, transparency, and VB dispersion for computationally predicted materials are all fruitful research paths for the p-TCM community.

JPE_10_4_042002_f003.png

Acknowledgments

A.N.F. acknowledges funding from the European Union’s Horizon 2020 Marie Skłodowska-Curie Actions under Grant Agreement No. 792720 (CLAReTE). M.M.M. acknowledges financial support from the NWO Start Up Grant 2019, Grant No. STU.019.026 (BRIDGE). The authors declare no conflicts of interest.

References

1. 

D. S. Ginley, J. D. Perkins, “Transparent conductors,” Handbook of Transparent Conductors, 1 –25 Springer, Boston, Massachusetts (2011). Google Scholar

2. 

H. Hosono, “Exploring electro-active functionality of transparent oxide materials,” Jpn. J. Appl. Phys., 52 (9), 090001 (2013). https://doi.org/10.7567/JJAP.52.090001 Google Scholar

3. 

M. Morales-Masis et al., “Transparent electrodes for efficient optoelectronics,” Adv. Electron. Mater., 3 (5), 1600529 (2017). https://doi.org/10.1002/aelm.201600529 Google Scholar

4. 

C. Battaglia, A. Cuevas and S. De Wolf, “High-efficiency crystalline silicon solar cells: status and perspectives,” Energy Environ. Sci., 9 (5), 1552 –1576 (2016). https://doi.org/10.1039/C5EE03380B EESNBY 1754-5692 Google Scholar

5. 

B. R. Sutherland and E. H. Sargent, “Perovskite photonic sources,” Nat. Photonics, 10 (5), 295 –302 (2016). https://doi.org/10.1038/nphoton.2016.62 NPAHBY 1749-4885 Google Scholar

6. 

J.-P. Correa-Baena et al., “Promises and challenges of perovskite solar cells,” Science, 358 (6364), 739 –744 (2017). https://doi.org/10.1126/science.aam6323 SCIEAS 0036-8075 Google Scholar

7. 

A. Daus et al., “Flexible CMOS electronics based on p-type Ge2Sb2Te5 and n-type InGaZnO4 semiconductors,” in IEEE Int. Electron Devices Meeting, 8.1.1 –8.1.4 (2017). https://doi.org/10.1109/IEDM.2017.8268349 Google Scholar

8. 

K. Nomura, T. Kamiya and H. Hosono, “Ambipolar oxide thin-film transistor,” Adv. Mater., 23 (30), 3431 –3434 (2011). https://doi.org/10.1002/adma.201101410 ADVMEW 0935-9648 Google Scholar

9. 

H. Hosono, “Recent progress in transparent oxide semiconductors: materials and device application,” Thin Solid Films, 515 (15), 6000 –6014 (2007). https://doi.org/10.1016/j.tsf.2006.12.125 THSFAP 0040-6090 Google Scholar

10. 

H. Kawazoe et al., “Transparent p-type conducting oxides: design and fabrication of p-n heterojunctions,” MRS Bull., 25 (8), 28 –36 (2000). https://doi.org/10.1557/mrs2000.148 MRSBEA 0883-7694 Google Scholar

11. 

H. Kawazoe et al., “P-type electrical conduction in transparent thin films of CuAlO2,” Nature, 389 (6654), 939 –942 (1997). https://doi.org/10.1038/40087 Google Scholar

12. 

R. Nagarajan et al., “p-type conductivity in CuCr1xMgxO2 films and powders,” J. Appl. Phys., 89 (12), 8022 –8025 (2001). https://doi.org/10.1063/1.1372636 JAPIAU 0021-8979 Google Scholar

13. 

L. Farrell et al., “Synthesis of nanocrystalline Cu deficient CuCrO2: a high figure of merit p-type transparent semiconductor,” J. Mater. Chem. C, 4 (1), 126 –134 (2016). https://doi.org/10.1039/C5TC03161C Google Scholar

14. 

G. Brunin et al., “Transparent conducting materials discovery using high-throughput computing,” npj Comput. Mater., 5 (1), 1 –13 (2019). https://doi.org/10.1038/s41524-019-0200-5 Google Scholar

15. 

J. B. Varley et al., “High-throughput design of non-oxide p-type transparent conducting materials: data mining, search strategy, and identification of boron phosphide,” Chem. Mater., 29 (6), 2568 –2573 (2017). https://doi.org/10.1021/acs.chemmater.6b04663 CMATEX 0897-4756 Google Scholar

16. 

G. Hautier et al., “Identification and design principles of low hole effective mass p-type transparent conducting oxides,” Nat. Commun., 4 2292 (2013). https://doi.org/10.1038/ncomms3292 NCAOBW 2041-1723 Google Scholar

17. 

B. A. D. Williamson et al., “Engineering valence band dispersion for high mobility p-type semiconductors,” Chem. Mater., 29 (6), 2402 –2413 (2017). https://doi.org/10.1021/acs.chemmater.6b03306 CMATEX 0897-4756 Google Scholar

18. 

R. K. M. Raghupathy et al., “Rational design of transparent p-type conducting non-oxide materials from high-throughput calculations,” J. Mater. Chem. C, 6 (3), 541 –549 (2018). https://doi.org/10.1039/C7TC05311H Google Scholar

19. 

M. Graužinytė, S. Goedecker and J. A. Flores-Livas, “Towards bipolar tin monoxide: revealing unexplored dopants,” Phys. Rev. Mater., 2 (10), 104604 (2018). https://doi.org/10.1103/PhysRevMaterials.2.104604 PRBMDO 0163-1829 Google Scholar

20. 

A. Zunger, “Practical doping principles,” Appl. Phys. Lett., 83 (1), 57 –59 (2003). https://doi.org/10.1063/1.1584074 APPLAB 0003-6951 Google Scholar

21. 

K. Ellmer, “Past achievements and future challenges in the development of optically transparent electrodes,” Nat. Photonics, 6 (12), 809 –817 (2012). https://doi.org/10.1038/nphoton.2012.282 NPAHBY 1749-4885 Google Scholar

22. 

M. Grundmann et al., “Cuprous iodide: a p-type transparent semiconductor, history, and novel applications,” Phys. Status Solidi A, 210 1671 –1703 (2013). https://doi.org/10.1002/pssa.201329349 PHSSAK 0031-8957 Google Scholar

23. 

H. Hernández-Cocoletzi et al., “Density functional study of the structural properties of copper iodide: LDA vs. GGA calculations,” J. Nano Res., 5 (1), 25 –30 (2009). https://doi.org/10.4028/www.scientific.net/JNanoR.5 1662-5250 Google Scholar

24. 

T. Jun et al., “Material design of p-type transparent amorphous semiconductor, Cu-Sn-I,” Adv. Mater., 30 (12), 1706573 (2018). https://doi.org/10.1002/adma.v30.12 ADVMEW 0935-9648 Google Scholar

25. 

F. Q. Huang, M. L. Liu and C. Yang, “Highly enhanced p-type electrical conduction in wide band gap Cu1xAl1xS2 polycrystals,” Sol. Energy Mater. Sol. Cells, 95 (10), 2924 –2927 (2011). https://doi.org/10.1016/j.solmat.2011.05.031 SEMCEQ 0927-0248 Google Scholar

26. 

D. Huang et al., “Understanding the high p-type conductivity in Cu-excess CuAlS2: a first-principles study,” Appl. Phys. Express, 9 (3), 031202 (2016). https://doi.org/10.7567/APEX.9.031202 Google Scholar

27. 

T. Arai et al., “Chemical design and example of transparent bipolar semiconductors,” J. Am. Chem. Soc., 139 (47), 17175 –17180 (2017). https://doi.org/10.1021/jacs.7b09806 JACSAT 0002-7863 Google Scholar

28. 

H. Hiramatsu et al., “Heavy hole doping of epitaxial thin films of a wide gap p-type semiconductor, LaCuOSe, and analysis of the effective mass,” Appl. Phys. Lett., 91 (1), 012104 (2007). https://doi.org/10.1063/1.2753546 APPLAB 0003-6951 Google Scholar

29. 

A. Agui, S. Shin and Y. Kumashiro, “Electronic structure of BP studied by resonant soft x-ray emission spectroscopy,” J. Phys. Soc. Jpn., 68 (1), 166 –169 (1999). https://doi.org/10.1143/JPSJ.68.166 JUPSAU 0031-9015 Google Scholar

30. 

Y. A. Nikolaev et al., “Photosensitive structures based on boron phosphide single crystals,” Semiconductors, 37 (8), 923 –926 (2003). https://doi.org/10.1134/1.1601659 SMICES 1063-7826 Google Scholar

31. 

J. Wang, J. Li and S.-S. Li, “Native p-type transparent conductive CuI via intrinsic defects,” J. Appl. Phys., 110 (5), 054907 (2011). https://doi.org/10.1063/1.3633220 JAPIAU 0021-8979 Google Scholar

32. 

B. Hönerlage, C. Klingshirn and J. B. Grun, “Spontaneous emission due to exciton—electron scattering in semiconductors,” Phys. Status Solidi B, 78 (2), 599 –608 (1976). https://doi.org/10.1002/pssb.2220780219 PHSSAK 0031-8957 Google Scholar

33. 

X. Dai et al., “A simple synthesis of transparent and highly conducting p-type CuxAl1xSy nanocomposite thin films as the hole transporting layer for organic solar cells,” RSC Adv., 8 16887 –16896 (2018). https://doi.org/10.1039/C8RA01299G Google Scholar

34. 

P. Popper and T. A. Ingles, “Boron phosphide, a III-V compound of zinc-blende structure,” Nature, 179 (4569), 1075 (1957). https://doi.org/10.1038/1791075a0 Google Scholar

35. 

W. Spring, “Über das spezifische Gewicht des Kupferjodürs,” Z. Anorg. Chem., 27 (1), 308 –309 (1901). https://doi.org/10.1002/zaac.19010270126 Google Scholar

36. 

K. Bädeker, “Über die elektrische Leitfähigkeit und die thermoelektrische Kraft einiger Schwermetallverbindungen,” Ann. Phys., 327 (4), 749 –766 (1907). https://doi.org/10.1002/(ISSN)1521-3889 Google Scholar

37. 

B. Stone and D. Hill, “Semiconducting properties of cubic boron phosphide,” Phys. Rev. Lett., 4 (6), 282 –284 (1960). https://doi.org/10.1103/PhysRevLett.4.282 PRLTAO 0031-9007 Google Scholar

38. 

A. Konno et al., “Effect of imidazolium salts on the performance of solid-state dyesensitized photovoltaic cell using copper iodide as a hole collector,” Electrochemistry, 70 (6), 432 –434 (2002). https://doi.org/10.5796/electrochemistry.70.432 ECHMBU 0305-9979 Google Scholar

39. 

M. Rusop et al., “Properties of pulsed-laser-deposited cui and characteristics of constructed dye-sensitized TiO2|Dye|CuI solid-state photovoltaic solar cells,” Jpn. J. Appl. Phys., 42 (8), 4966 –4972 (2003). https://doi.org/10.1143/JJAP.42.4966 Google Scholar

40. 

M. Rusop et al., “Copper iodide thin films as a p-type electrical conductivity in dye-sensitized p-CuI|Dye|n-TiO2 heterojunction solid state solar cells,” Surf. Rev. Lett., 11 (6), 577 –583 (2004). https://doi.org/10.1142/S0218625X04006554 SRLEFH 0218-625X Google Scholar

41. 

Y. Zhou et al., “Glancing angle deposition of copper iodide nanocrystals for efficient organic photovoltaics,” Nano Lett., 12 (8), 4146 –4152 (2012). https://doi.org/10.1021/nl301709x NALEFD 1530-6984 Google Scholar

42. 

J. W. Kim et al., “High performance organic planar heterojunction solar cells by controlling the molecular orientation,” Curr. Appl. Phys., 13 (1), 7 –11 (2013). https://doi.org/10.1016/j.cap.2012.06.003 1567-1739 Google Scholar

43. 

J. C. Bernede et al., “MoO3/CuI hybrid buffer layer for the optimization of organic solar cells based on a donor-acceptor triphenylamine,” Sol. Energy Mater. Sol. Cells, 110 107 –114 (2013). https://doi.org/10.1016/j.solmat.2012.12.003 SEMCEQ 0927-0248 Google Scholar

44. 

J. A. Christians, R. C. M. Fung and P. V. Kamat, “An inorganic hole conductor for organo-lead halide perovskite solar cells. Improved hole conductivity with copper iodide,” J. Am. Chem. Soc., 136 (2), 758 –764 (2014). https://doi.org/10.1021/ja411014k JACSAT 0002-7863 Google Scholar

45. 

W. Sun et al., “Solution-processed copper iodide as an inexpensive and effective anode buffer layer for polymer solar cells,” J. Phys. Chem. C, 118 (30), 16806 –16812 (2014). https://doi.org/10.1021/jp412784q JPCCCK 1932-7447 Google Scholar

46. 

V. V. Travkin et al., “A hybrid CuI/fullerene heterojunction in transparent flexible photovoltaic cells,” Fullerenes Nanotubes Carbon Nanostruct., 23 (8), 721 –724 (2015). https://doi.org/10.1080/1536383X.2014.986799 Google Scholar

47. 

K. Zhao et al., “Highly efficient organic solar cells based on a robust room-temperature solution-processed copper iodide hole transporter,” Nano Energy, 16 458 –469 (2015). https://doi.org/10.1016/j.nanoen.2015.07.018 Google Scholar

48. 

S. A. Mohamed et al., “CuI as versatile hole-selective contact for organic solar cell based on anthracene-containing PPE-PPV,” Sol. Energy Mater. Sol. Cells, 143 369 –374 (2015). https://doi.org/10.1016/j.solmat.2015.07.003 SEMCEQ 0927-0248 Google Scholar

49. 

B. Gil et al., “Recent progress in inorganic hole transport materials for efficient and stable perovskite solar cells,” Electron. Mater. Lett., 15 (5), 505 –524 (2019). https://doi.org/10.1007/s13391-019-00163-6 Google Scholar

50. 

T. K. Chaudhuri et al., “A chemical method for preparing copper iodide thin films,” Jpn. J. Appl. Phys., 29 (2), L352 –L354 (1990). https://doi.org/10.1143/JJAP.29.L352 Google Scholar

51. 

V. Raj et al., “Introduction of TiO2 in CuI for its improved performance as a p-type transparent conductor,” ACS Appl. Mater. Interfaces, 11 (27), 24254 –24263 (2019). https://doi.org/10.1021/acsami.9b05566 AAMICK 1944-8244 Google Scholar

52. 

N. Yamada et al., “High-mobility transparent p-type CuI semiconducting layers fabricated on flexible plastic sheets: toward flexible transparent electronics,” Adv. Electron. Mater., 3 (12), 1700298 (2017). https://doi.org/10.1002/aelm.201700298 Google Scholar

53. 

K. Tennakone et al., “Preparation of thin polycrystalline films of cuprous iodide and photoelectrochemical dye-sensitization,” Thin Solid Films, 217 (1–2), 129 –132 (1992). https://doi.org/10.1016/0040-6090(92)90618-L THSFAP 0040-6090 Google Scholar

54. 

R. N. Bulakhe et al., “Deposition of copper iodide thin films by chemical bath deposition (CBD) and successive ionic layer adsorption and reaction (SILAR) methods,” Curr. Appl. Phys., 13 (8), 1661 –1667 (2013). https://doi.org/10.1016/j.cap.2013.05.014 1567-1739 Google Scholar

55. 

M. Rahman et al., “Unraveling the electrical properties of solution-processed copper iodide thin films for CuI/n-Si solar cells,” Mater. Res. Bull., 118 110518 (2019). https://doi.org/10.1016/j.materresbull.2019.110518 MRBUAC 0025-5408 Google Scholar

56. 

A. Liu et al., “Room-temperature solution-synthesized p-type copper(I) iodide semiconductors for transparent thin-film transistors and complementary electronics,” Adv. Mater., 30 (34), 1802379 (2018). https://doi.org/10.1002/adma.v30.34 ADVMEW 0935-9648 Google Scholar

57. 

K. Tennakone et al., “Deposition of thin conducting films of CuI on glass,” Sol. Energy Mater. Sol. Cells, 55 (3), 283 –289 (1998). https://doi.org/10.1016/S0927-0248(98)00117-2 SEMCEQ 0927-0248 Google Scholar

58. 

P. M. Sirimanne et al., “Characterization of transparent conducting CuI thin films prepared by pulsed laser deposition technique,” Chem. Phys. Lett., 366 (5–6), 485 –489 (2002). https://doi.org/10.1016/S0009-2614(02)01590-7 CHPLBC 0009-2614 Google Scholar

59. 

M. Rusop et al., “Annealing temperature effects on synthesis of n-TiO2/dye/p-CuI solid-state solar cells,” Jpn. J. Appl. Phys., 44 (4B), 2560 –2567 (2005). https://doi.org/10.1143/JJAP.44.2560 Google Scholar

60. 

M. Rusop et al., “Optical band gap excitation and photoelectron generation in titanium dioxide-based solid state solar cells,” Surf. Rev. Lett., 12 (5–6), 681 –689 (2005). https://doi.org/10.1142/S0218625X05007694 SRLEFH 0218-625X Google Scholar

61. 

M. Rusop et al., “Study on the properties and charge generation in dye-sensitized n-TiO2/dye/p-CuI solid state photovoltaic solar cells,” Appl. Surf. Sci., 252 (20), 7389 –7396 (2006). https://doi.org/10.1016/j.apsusc.2005.08.088 ASUSEE 0169-4332 Google Scholar

62. 

B. L. Zhu and X. Z. Zhao, “Transparent conductive CuI thin films prepared by pulsed laser deposition,” Phys. Status Solidi A, 208 (1), 91 –96 (2011). https://doi.org/10.1002/pssa.201026239 PHSSAK 0031-8957 Google Scholar

63. 

T. Tanaka, K. Kawabata and M. Hirose, “Transparent, conductive CuI films prepared by rf-dc coupled magnetron sputtering,” Thin Solid Films, 281–282 179 –181 (1996). https://doi.org/10.1016/0040-6090(96)08607-5 THSFAP 0040-6090 Google Scholar

64. 

C. Yang et al., “Room-temperature synthesized copper iodide thin film as degenerate p-type transparent conductor with a boosted figure of merit,” Proc. Natl. Acad. Sci. U. S. A., 113 (46), 12929 –12933 (2016). https://doi.org/10.1073/pnas.1613643113 PNASA6 0027-8424 Google Scholar

65. 

P. Y. Stakhira and V. V. Cherpak, “The properties of heterojunction based on CuI/pentacene/Al,” Vacuum, 83 (8), 1129 –1131 (2009). https://doi.org/10.1016/j.vacuum.2009.02.004 VACUAV 0042-207X Google Scholar

66. 

P. Stakhira et al., “Characteristics of organic light emitting diodes with copper iodide as injection layer,” Thin Solid Films, 518 (23), 7016 –7018 (2010). https://doi.org/10.1016/j.tsf.2010.06.051 THSFAP 0040-6090 Google Scholar

67. 

D. K. Kaushik et al., “Thermal evaporated copper iodide (CuI) thin films: a note on the disorder evaluated through the temperature dependent electrical properties,” Sol. Energy Mater. Sol. Cells, 165 52 –58 (2017). https://doi.org/10.1016/j.solmat.2017.02.030 SEMCEQ 0927-0248 Google Scholar

68. 

Y. Kumashiro and Y. Okada, “Schottky barrier diodes using thick, well-characterized boron phosphide wafers,” Appl. Phys. Lett., 47 (1), 64 –66 (1985). https://doi.org/10.1063/1.96406 APPLAB 0003-6951 Google Scholar

69. 

J. Lee et al., “Photoelectrochemical behavior of p-type boron phosphide photoelectrode in acidic solution,” Bull. Chem. Soc. Jpn., 58 (9), 2634 –2637 (1985). https://doi.org/10.1246/bcsj.58.2634 BCSJA8 0009-2673 Google Scholar

70. 

A. Goossens, E. M. Kelder and J. Schoonman, “Polycrystalline boron phosphide semiconductor electrodes,” Ber. Bunsengesellschaft Phys. Chem., 93 (10), 1109 –1114 (1989). https://doi.org/10.1002/bbpc.v93:10 Google Scholar

71. 

Y. Kumashiro, M. Hirabayashi and S. Takagi, “Boron phosphide as a refractory semiconductor,” MRS Proc., 162 585 –594 (1989). https://doi.org/10.1557/PROC-162-585 Google Scholar

72. 

A. Goossens et al., “Structural, optical, and electronic properties of silicon/boron phosphide heterojunction photoelectrodes,” Ber. Bunsengesellschaft Phys. Chem., 95 (4), 503 –510 (1991). https://doi.org/10.1002/bbpc.v95:4 Google Scholar

73. 

T. Udagawa and G. Shimaoka, “Heteroepitaxial growth of boronphosphide III–V semiconductor on silicon by organometallic chemical vapor deposition,” J. Ceram. Process. Res., 4 (2), 80 –83 (2003). Google Scholar

74. 

B. Padavala et al., “Cubic boron phosphide epitaxy on zirconium diboride,” J. Cryst. Growth, 483 115 –120 (2018). https://doi.org/10.1016/j.jcrysgro.2017.11.014 JCRGAE 0022-0248 Google Scholar

75. 

N. Ding et al., “Controllable carrier type in boron phosphide nanowires toward homostructural optoelectronic devices,” ACS Appl. Mater. Interfaces, 10 (12), 10296 –10303 (2018). https://doi.org/10.1021/acsami.7b17204 AAMICK 1944-8244 Google Scholar

76. 

X. Feng et al., “Low temperature synthesis of boron phosphide nanocrystals,” Mater. Lett., 59 (8–9), 865 –867 (2005). https://doi.org/10.1016/j.matlet.2004.10.067 MLETDJ 0167-577X Google Scholar

77. 

V. A. Mukhanov et al., “Self-propagating high-temperature synthesis of boron phosphide,” J. Superhard Mater., 35 (6), 415 –417 (2013). https://doi.org/10.3103/S1063457613060105 1063-4576 Google Scholar

78. 

K. Woo, K. Lee and K. Kovnir, “BP: synthesis and properties of boron phosphide,” Mater. Res. Express, 3 (7), 074003 (2016). https://doi.org/10.1088/2053-1591/3/7/074003 Google Scholar

79. 

L. Shi et al., “n-type boron phosphide as a highly stable, metal-free, visible-light active photocatalyst for hydrogen evolution,” Nano Energy, 28 158 –163 (2016). https://doi.org/10.1016/j.nanoen.2016.08.041 Google Scholar

80. 

M. Odawara et al., “Suppression of indium vaporization from GaN/GaInN superlattice by BP capping layer,” J. Cryst. Growth, 263 (1–4), 645 –647 (2004). https://doi.org/10.1016/j.jcrysgro.2003.11.044 JCRGAE 0022-0248 Google Scholar

81. 

M. Odawara, T. Udagawa and G. Shimaoka, “Organometallic chemical vapor deposition growth of heterostructure of wide band gap and transparent boron phosphide on silicon,” Jpn. J. Appl. Phys., 44 (1B), 681 –683 (2005). https://doi.org/10.1143/JJAP.44.681 Google Scholar

82. 

A. Goossens and J. Schoonman, “An impedance study of boron phosphide semiconductor electrodes,” J. Electrochem. Soc., 139 (3), 893 –900 (1992). https://doi.org/10.1149/1.2069321 JESOAN 0013-4651 Google Scholar

83. 

B. Padavala et al., “Epitaxy of boron phosphide on aluminum nitride(0001)/sapphire substrate,” Cryst. Growth Des., 16 (2), 981 –987 (2016). https://doi.org/10.1021/acs.cgd.5b01525 CGDEFU 1528-7483 Google Scholar

84. 

B. Padavala et al., “CVD growth and properties of boron phosphide on 3C-SiC,” J. Cryst. Growth, 449 15 –21 (2016). https://doi.org/10.1016/j.jcrysgro.2016.05.031 JCRGAE 0022-0248 Google Scholar

85. 

S. P. Huber et al., “Exploiting the P L2,3 absorption edge for optics: spectroscopic and structural characterization of cubic boron phosphide thin films,” Opt. Mater. Express, 6 (12), 3946 (2016). https://doi.org/10.1364/OME.6.003946 Google Scholar

86. 

W. Liu et al., “Structural, mechanical properties and composition analysis of boron phosphide coatings,” J. Alloys Compd., 538 169 –172 (2012). https://doi.org/10.1016/j.jallcom.2012.05.043 JALCEU 0925-8388 Google Scholar

87. 

E. Schroten, A. Goossens and J. Schoonman, “Synthesis of nanometer-scale boron phosphide whiskers by vapor-liquid-solid chemical vapor deposition,” J. Appl. Phys., 79 (8), 4465 –4467 (1996). https://doi.org/10.1063/1.361759 JAPIAU 0021-8979 Google Scholar

88. 

E. Schroten, A. Goossens and J. Schoonman, “Large-surface-area boron phosphide liquid junction solar cells,” J. Electrochem. Soc., 146 (6), 2045 –2048 (1999). https://doi.org/10.1149/1.1391889 JESOAN 0013-4651 Google Scholar

89. 

S. Dalui et al., “Boron phosphide films prepared by co-evaporation technique: synthesis and characterization,” Thin Solid Films, 516 (15), 4958 –4965 (2008). https://doi.org/10.1016/j.tsf.2007.09.047 THSFAP 0040-6090 Google Scholar

90. 

Z. C. Jia et al., “Effect of gas flow ratio on the microstructure and mechanical properties of boron phosphide films prepared by reactive magnetron sputtering,” Appl. Surf. Sci., 258 (1), 356 –360 (2011). https://doi.org/10.1016/j.apsusc.2011.08.067 ASUSEE 0169-4332 Google Scholar

91. 

J. Damisa et al., “Morphological and optical study of thin films of CuAlS2 deposited by metal organic chemical vapour deposition technique,” Mater. Res. Express, 4 (8), 086412 (2017). https://doi.org/10.1088/2053-1591/aa851d Google Scholar

92. 

N. Schneider et al., “Transparent ohmic contact for CIGS solar cells based on p-type aluminum copper sulfide material synthesized by atomic layer deposition,” ACS Appl. Energy Mater., 1 (12), 7220 –7229 (2018). https://doi.org/10.1021/acsaem.8b01647 Google Scholar

93. 

D. N. Okoli, A. J. Ekpunobi and C. E. Okeke, “Optical properties of chemical bath deposited CuAlS2 thin films,” Pac. J. Sci. Technol., 7 (1), (2006). Google Scholar

94. 

L. Duclaux et al., “Simulation and growing study of Cu-Al-S thin films deposited by atomic layer deposition,” Thin Solid Films, 594 232 –237 (2015). https://doi.org/10.1016/j.tsf.2015.06.014 THSFAP 0040-6090 Google Scholar

95. 

A. U. Moreh, M. Momoh and B. Hamza, “The effect of sulfurisation temperature on structural properties of CuAlS2 thin films,” IOSR J. Appl. Phys., 3 (1), 12 –17 (2013). https://doi.org/10.9790/4861-0311217 Google Scholar

96. 

J. Olejníček et al., “CuIn1xAlxS2 thin films prepared by sulfurization of metallic precursors,” J. Alloys Compd., 509 (41), 10020 –10024 (2011). https://doi.org/10.1016/j.jallcom.2011.08.016 JALCEU 0925-8388 Google Scholar

97. 

R. Brini et al., “Study of the growth of CuAlS2 thin films on oriented silicon (111),” Thin Solid Films, 517 (7), 2191 –2194 (2009). https://doi.org/10.1016/j.tsf.2008.10.086 THSFAP 0040-6090 Google Scholar

98. 

M. Abaab, A. S. Bouazzi and B. Rezig, “Competitive CuAlS2 oxygen gas sensor,” Microelectron. Eng., 51 343 –348 (2000). https://doi.org/10.1016/S0167-9317(99)00495-5 MIENEF 0167-9317 Google Scholar

99. 

S. M. Ahmad, “Study of structural and optical properties of quaternary CuxAg1xAlS2 thin films,” Optik, 127 (20), 10004 –10013 (2016). https://doi.org/10.1016/j.ijleo.2016.07.034 OTIKAJ 0030-4026 Google Scholar

100. 

M. Caglar, S. Ilican and Y. Caglar, “Structural, morphological and optical properties of CuAlS2 films deposited by spray pyrolysis method,” Opt. Commun., 281 (6), 1615 –1624 (2008). https://doi.org/10.1016/j.optcom.2007.11.031 OPCOB8 0030-4018 Google Scholar

101. 

R. Woods-Robinson et al., “Combinatorial tuning of structural and optoelectronic properties in CuxZn1xS,” Matter, 1 (4), 862 –880 (2019). https://doi.org/10.1016/j.matt.2019.06.019 Google Scholar

102. 

R. Woods-Robinson et al., “P-type transparent Cu-alloyed ZnS deposited at room temperature,” Adv. Electron. Mater., 2 (6), 1500396 (2016). https://doi.org/10.1002/aelm.201500396 Google Scholar

103. 

R. Woods-Robinson et al., “Wide band gap chalcogenide semiconductors,” (2019). Google Scholar

104. 

A. M. Diamond et al., “Copper-alloyed ZnS as a p-type transparent conducting material,” Phys. Status Solidi A, 209 (11), 2101 –2107 (2012). https://doi.org/10.1002/pssa.v209.11 Google Scholar

105. 

L. Wei et al., “Machine learning optimization of p-type transparent conducting films,” Chem. Mater., 31 (18), 7340 –7350 (2019). https://doi.org/10.1021/acs.chemmater.9b01953 CMATEX 0897-4756 Google Scholar

106. 

S. K. Maurya et al., “High figure-of-merit p-type transparent conductor, Cu alloyed ZnS via radio frequency magnetron sputtering,” J. Phys. D Appl. Phys., 50 (50), 505107 (2017). https://doi.org/10.1088/1361-6463/aa95b3 Google Scholar

107. 

T. T. Thanh et al., “Synthesis and photocatalytic application of ternary Cu-Zn-S nanoparticle-sensitized TiO2 nanotube arrays,” Chem. Eng. J., 210 425 –431 (2012). https://doi.org/10.1016/j.cej.2012.09.004 Google Scholar

108. 

F. Di Benedetto et al., “Electrodeposited semiconductors at room temperature: an x-ray absorption spectroscopy study of Cu-, Zn-, S-bearing thin films,” Electrochim. Acta, 179 495 –503 (2015). https://doi.org/10.1016/j.electacta.2015.05.168 ELCAAV 0013-4686 Google Scholar

109. 

D. E. Ortíz-Ramos, L. A. González and R. Ramirez-Bon, “P-type transparent Cu doped ZnS thin films by the chemical bath deposition method,” Mater. Lett., 124 267 –270 (2014). https://doi.org/10.1016/j.matlet.2014.03.082 MLETDJ 0167-577X Google Scholar

110. 

M. Dula, K. Yang and M. Ichimura, “Photochemical deposition of a p-type transparent alloy semiconductor CuxZnyS,” Semicond. Sci. Technol., 27 (12), 125007 (2012). https://doi.org/10.1088/0268-1242/27/12/125007 SSTEET 0268-1242 Google Scholar

111. 

K. Yang and M. Ichimura, “Fabrication of transparent p-type CuxZnyS thin films by the electrochemical deposition method,” Jpn. J. Appl. Phys., 50 (4), 040202 (2011). https://doi.org/10.1143/JJAP.50.040202 Google Scholar

112. 

K. Ishikawa et al., “Preparation and electrical properties of (LaO)AgS and (LnO)CuS (Ln = La, Pr, or Nd),” J. Electrochem., 138 (4), 1166 –1170 (1991). https://doi.org/10.1149/1.2085735 Google Scholar

113. 

M. Palazzi et al., “Crystal structure and properties of (LaO) CuS and (LaO) AgS,” The Rare Earths in Modern Science and Technology, 347 –350 Springer, Boston, Massachusetts (1982). Google Scholar

114. 

K. Ueda et al., “Transparent p-type semiconductor: LaCuOS layered oxysulfide,” Appl. Phys. Lett., 77 (17), 2701 –2703 (2000). https://doi.org/10.1063/1.1319507 APPLAB 0003-6951 Google Scholar

115. 

H. Hiramatsu et al., “Fabrication of heteroepitaxial thin films of layered oxychalcogenides LnCuOCh (Ln = La–Nd; Ch = S–Te) by reactive solid-phase epitaxy,” J. Mater. Res., 19 (7), 2137 –2143 (2004). https://doi.org/10.1557/JMR.2004.0273 JMREEE 0884-2914 Google Scholar

116. 

D. O. Scanlon et al., “Understanding doping anomalies in degenerate p-type semiconductor LaCuOSe,” J. Mater. Chem. C, 2 (17), 3429 –3438 (2014). https://doi.org/10.1039/C4TC00096J Google Scholar

117. 

M.-L. Liu et al., “A promising p-type transparent conducting material: layered oxysulfide [Cu2S2][Sr2Sc2O5],” J. Appl. Phys., 102 (11), 116108 (2007). https://doi.org/10.1063/1.2817643 JAPIAU 0021-8979 Google Scholar

118. 

E. Hildebrandt et al., “Controlled oxygen vacancy induced p-type conductivity in HfO2x thin films,” Appl. Phys. Lett., 99 (11), 112902 (2011). https://doi.org/10.1063/1.3637603 APPLAB 0003-6951 Google Scholar

119. 

K.-H. Xue et al., “Prediction of semimetallic tetragonal Hf2O3 and Zr2O3 from first principles,” Phys. Rev. Lett., 110 (6), 065502 (2013). https://doi.org/10.1103/PhysRevLett.110.065502 PRLTAO 0031-9007 Google Scholar

120. 

T. Kim et al., “Material design of new p-type tin oxyselenide semiconductor through valence band engineering and its device application,” ACS Appl. Mater. Interfaces, 11 (43), 40214 –40221 (2019). https://doi.org/10.1021/acsami.9b12186 Google Scholar

121. 

F.-L. Schein, H. von Wenckstern and M. Grundmann, “Transparent p-CuI/n-ZnO heterojunction diodes,” Appl. Phys. Lett., 102 (9), 092109 (2013). https://doi.org/10.1063/1.4794532 APPLAB 0003-6951 Google Scholar

122. 

D. Chen et al., “Growth strategy and physical properties of the high mobility p-type cui crystal,” Cryst. Growth Des., 10 (5), 2057 –2060 (2010). https://doi.org/10.1021/cg100270d CGDEFU 1528-7483 Google Scholar

123. 

C. S. Herrick and A. D. Tevebaugh, “Oxygen-controlled conduction in thin films of cuprous iodide: a mixed valency anion semiconductor,” J. Electrochem. Soc., 110 (2), 119 –121 (1963). https://doi.org/10.1149/1.2425687 JESOAN 0013-4651 Google Scholar

124. 

K. Bädeker, “Über eine eigentümliche Form elektrischen Leitvermögens bei festen Körpern,” Ann. Phys., 334 (8), 566 –584 (1909). https://doi.org/10.1002/(ISSN)1521-3889 Google Scholar

125. 

Y. Kumashiro, “Refractory semiconductor of boron phosphide,” J. Mater. Res., 5 (12), 2933 –2947 (1990). https://doi.org/10.1557/JMR.1990.2933 JMREEE 0884-2914 Google Scholar

126. 

M. Iwami, T. Tohda and K. Kawabe, “Crystal growth of boron mono-phosphide and its electrical and optical properties,” IEEJ Trans. Fundam. Mater., 95 (5), 216 –222 (1975). https://doi.org/10.1541/ieejfms1972.95.216 Google Scholar

127. 

Y. Youn et al., “Large-scale computational identification of p-type oxide semiconductors by hierarchical screening,” Chem. Mater., 31 (15), 5475 –5483 (2019). https://doi.org/10.1021/acs.chemmater.9b00816 CMATEX 0897-4756 Google Scholar

128. 

N. Sarmadian et al., “Easily doped p-type, low hole effective mass, transparent oxides,” Sci. Rep., 6 1 –9 (2016). https://doi.org/10.1038/srep20446 SRCEC3 2045-2322 Google Scholar

129. 

P. M. Sirimanne and V. P. S. Perera, “Progress in dye-sensitized solid state solar cells,” Phys. Status Solidi B, 245 (9), 1828 –1833 (2008). https://doi.org/10.1002/pssb.200779540 Google Scholar

130. 

S. Villain et al., “Electrical properties of CuI and the phase boundary Cu|CuI,” Solid State Ionics, 76 (3–4), 229 –235 (1995). https://doi.org/10.1016/0167-2738(94)00285-Z SSIOD3 0167-2738 Google Scholar

131. 

O. Ohtaka et al., “Ionic conductivities of CuI phases at high pressures and temperatures,” J. Phys. Soc. Jpn., 79 (Suppl. A), 51 –53 (2010). https://doi.org/10.1143/JPSJS.79SA.51 Google Scholar

132. 

A. N. Gruzintsev and V. N. Zagorodnev, “Change in optical properties of CuI crystals upon annealing in vacuum,” Phys. Solid State, 54 (1), 117 –122 (2012). https://doi.org/10.1134/S1063783412010167 PSOSED 1063-7834 Google Scholar

133. 

D. Kim et al., “Thermal-strain-induced splitting of heavy- and light-hole exciton energies in CuI thin films grown by vacuum evaporation,” Phys. Rev. B, 60 (19), 13879 –13884 (1999). https://doi.org/10.1103/PhysRevB.60.13879 Google Scholar

134. 

Y. Peng et al., “Efficient organic solar cells using copper(I) iodide (CuI) hole transport layers,” Appl. Phys. Lett., 106 (24), 243302 (2015). https://doi.org/10.1063/1.4922758 APPLAB 0003-6951 Google Scholar

135. 

R. Heasley et al., “Vapor deposition of transparent, p-type cuprous iodide via a two-step conversion process,” ACS Appl. Energy Mater., 1 (12), 6953 –6963 (2018). https://doi.org/10.1021/acsaem.8b01363 Google Scholar

136. 

S. J. Clarke et al., “Structures, physical properties, and chemistry of layered oxychalcogenides and oxypnictides,” Inorg. Chem., 47 (19), 8473 –8486 (2008). https://doi.org/10.1021/ic8009964 INOCAJ 0020-1669 Google Scholar

137. 

M. Morales-Masis et al., “Conductance switching in Ag2S devices fabricated by in situ sulfurization,” Nanotechnology, 20 (9), 095710 (2009). https://doi.org/10.1088/0957-4484/20/9/095710 NNOTER 0957-4484 Google Scholar

138. 

H. Lin et al., “Growth of environmentally stable transition metal selenide films,” Nat. Mater., 18 (6), 602 –607 (2019). https://doi.org/10.1038/s41563-019-0321-8 NMAACR 1476-1122 Google Scholar

139. 

A. Jain et al., “The materials project: a materials genome approach to accelerating materials innovation,” APL mater., 1 (1), 011002 (2013). https://doi.org/10.1063/1.4812323 Google Scholar

140. 

T. S. Tripathi and M. Karppinen, “Atomic layer deposition of p-type semiconducting thin films: a review,” Adv. Mater. Interfaces, 4 (24), 1700300 (2017). https://doi.org/10.1002/admi.201700300 Google Scholar

141. 

R. Eason, Pulsed Laser Deposition of Thin Films. Applications-Led Growth of Functional Materials, John Wiley & Sons, Inc., Hoboken (2007). Google Scholar

142. 

C. Stock and E. E. McCabe, “The magnetic and electronic properties of oxyselenides: influence of transition metal ions and lanthanides,” J. Phys. Condens. Matter, 28 (45), 453001 (2016). https://doi.org/10.1088/0953-8984/28/45/453001 JCOMEL 0953-8984 Google Scholar

Biography

Angela N. Fioretti received her PhD in materials science from Colorado School of Mines in 2017. Her graduate work focused on combinatorial development of ternary nitride semiconductors for photovoltaics. Currently, she is a Marie Curie postdoctoral fellow at the Ecole Polytechnique Fédérale de Lausanne in Switzerland, working on development of novel carrier selective contact materials for solar cells.

Monica Morales-Masis received her PhD from Leiden University, Netherlands in 2012. From 2012 to 2017 she was at the Photovoltaics Laboratory (PVLab) of the Ecole Polytechnique Federale de Lausanne (EPFL), Switzerland as team leader for its transparent conducting oxides activities. Since 2018, she is an assistant professor at the University of Twente, Netherlands, focusing on synthesis and study of optoelectronic materials, such as transparent conducting oxides, sulfides and halide perovskites for photovoltaics and applications beyond.

© 2020 Society of Photo-Optical Instrumentation Engineers (SPIE)
Angela N. Fioretti and Monica Morales-Masis "Bridging the p-type transparent conductive materials gap: synthesis approaches for disperse valence band materials," Journal of Photonics for Energy 10(4), 042002 (3 February 2020). https://doi.org/10.1117/1.JPE.10.042002
Received: 1 November 2019; Accepted: 14 January 2020; Published: 3 February 2020
Lens.org Logo
CITATIONS
Cited by 20 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Thin films

Copper

Sputter deposition

Oxides

Transparency

Solid state physics

Chemical vapor deposition

RELATED CONTENT

Multichromism in vanadium oxides: the influence of dopants
Proceedings of SPIE (January 01 1900)
Oxide thin film transistors on novel flexible substrates
Proceedings of SPIE (February 15 2010)
Fabrication of p-ZnO thin films by ZrN codoping
Proceedings of SPIE (October 15 2012)
Infrared transparent conductive oxides
Proceedings of SPIE (September 07 2001)

Back to Top