mo sep 2012

Upload: omar-dosky

Post on 03-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Mo sep 2012

    1/7

    Layered semiconductor molybdenum disulfide nanomembrane based

    Schottky-barrier solar cellsMariyappan Shanmugam, Chris A. Durcan and Bin Yu*

    Received 21st August 2012, Accepted 27th September 2012

    DOI: 10.1039/c2nr32394j

    We demonstrate Schottky-barrier solar cells employing a stack of layer-structured semiconductor

    molybdenum disulfide (MoS2) nanomembranes, synthesized by the chemical-vapor-deposition method,

    as the critical photoactive layer. An MoS2nanomembrane forms a Schottky-barrier with a metal

    contact by the layer-transfer process onto an indium tin oxide (ITO) coated glass substrate. Two

    vibrational modes in MoS2nanomembranes, E12g(in-plane) and A1g(perpendicular-to-plane), were

    verified by Raman spectroscopy. With a simple stacked structure of ITOMoS2Au, the fabricated

    solar cell demonstrates a photo-conversion efficiency of 0.7% for110 nm MoS2 and 1.8% for 220 nm

    MoS2. The improvement is attributed to a substantial increase in photonic absorption. The MoS2nanomembrane exhibits efficient photo-absorption in the spectral region of 350950 nm, as confirmed

    by the external quantum efficiency. A sizable increase in MoS2 thickness results in only minor change in

    MottSchottky behavior, indicating that defect density is insensitive to nanomembrane thickness

    attributed to the dangling-bond-free layered structure.

    Introduction

    Schottky-barrier solar cells have the appealing advantage of low-

    cost manufacture attributable to their simple and versatile

    fabrication process and thin-film-like structure (reduced active

    material consumption as compared with pn or pin junctionbulk semiconductor solar cells). Schottky-barrier solar cells rely

    mostly on a single semiconductor material (either p-or n-type

    doped) in association with a metal, leading to energy barrier

    formation and carrier dissociation at the metalsemiconductor

    (MS) interface. Various MS material systems have been used

    in Schottky-barrier solar cells, including Alamorphous carbon

    nitride,1 Au, Al, Mg and CaPbSe quantum dots,2 AlSnS,3 and

    AuGaAs.4 Recently, carbon derivatives have attracted attention

    due to their promising optical/electrical properties. Schottky

    junctions made by carbon nanotubeCdSe,5 grapheneCdS

    nanowire,6 graphene, carbon nanotubeSi,7,8 and grapheneSi

    nanowire911 have been reported. In general, MS interfacial

    behavior and carrier mobility/photon absorption in the semi-conductor material are key factors that impact the overall

    performance of Schottky-barrier solar cells. While a group of

    semiconductors such as boron doped diamond,12 GaAs,13,14

    InP,15 SiGe,16 GaP,17 GaN,18 and SiC19 can be used, the design of

    an MS system with appropriate material work function differ-

    ence is an essential factor to determine the Schottky-barrier

    height and efficiency of carrier separation.

    The development of Schottky-barrier solar cells has been

    decelerated over the past decade due to several fundamental

    material challenges, including interfacial oxide layer formation

    and Fermi-level pinning caused by interfacial states (which are

    inevitable at MS interfaces).2024

    While poor interface properties play a substantial role in theimpeded improvement of energy-conversion efficiency, searching

    for a new class of semiconductor materials free of interfacial

    states (hence without Fermi level pinning) would be of strategic

    significance to the ultimate realization of highly efficient

    Schottky-barrier solar cells.

    Transition-metal dichalcogenide semiconductors exhibit two

    dimensional (2D) layered lattice structures with strong in-plane

    covalent bonds and weak interlayer van der Waals interaction.

    Molybdenum disulphide (MoS2), a layered semiconductor, has

    recently attracted attention in the fields of nanoelectronics,2530

    energy storage3133 and photovoltaics.34 Nanosheets of monolayer

    or multilayer MoS2can be obtained via mechanical or chemical

    exfoliation,35,36similar to theway graphene is produced.Recently,direct assembly of MoS2 nanosheets on SiO2 by the chemical-

    vapor-deposition (CVD) method has been demonstrated.37 MoS2exhibits direct energy bandgap (1.85 eV) in monolayer and

    indirect bandgap (1.3 eV) in multilayer/bulk form, respec-

    tively.38 On electronic properties, it has beenrecentlyreportedthat

    MoS2 shows carrier mobility values of 200 cm2 V1 s1 in

    monolayer and 517 cm2 V1 s1 in a few layers, suggesting

    promising conducting capacity.39,40 Most importantly, the unique

    structure of layered semiconductors provides inherent advantages

    of forming a smooth surface free of any dangling chemical bonds,

    potentially enabling the design of highly efficientSchottky-barrier

    College of Nanoscale Science and Engineering, State University of NewYork, Albany, NY-12203, USA. E-mail: [email protected]; Fax: +1 518956-7492; Tel: +1 518 956-7492

    This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 73997405 | 7399

    Dynamic Article LinksC

  • 8/12/2019 Mo sep 2012

    2/7

    solar cells. Various metal contacts, including Ag, Al, Au, Co and

    Fe, have been investigated for the formation of Schottky-barriers

    with MoS2 bulk crystals. No Fermi level pinning was observed

    even for a high density of material defect (1014 cm3).41

    In this paper we demonstrate solar cells employing a layer-

    structured semiconductor MoS2 nanomembranemetal (Au)

    Schottky barrier. The CVD-assembled MoS2 nanomembrane

    also functions as the key photo-active layer. The solar cells are

    made by a simple stack structure of ITOMoS2Au. Basicmaterial properties of MoS2, major solar cell device performance

    metrics, and the impact of MoS2 nanomembrane thickness are

    discussed. The MoS2Au Schottky-barrier interface is charac-

    terized by capacitance vs. voltage (CV) measurements.

    Experimental section

    Large-area (up to several square centimeters) semiconductor

    MoS2nanomembranes were synthesized by the CVD approach.

    Fig. 1(a) is the schematic of the CVD growth furnace system

    showing Ar as a carrier gas and a sulphur (S) source flowing

    towards a molybdenum (Mo)-coated SiSiO2 substrate (where

    vaporsolid reaction occurs to assemble the MoS2 nano-

    membrane). CVD growth of MoS2silicon coated with 100 nm of

    SiO2 was used as the substrate for the pre-deposition of Mo. A

    thin Mo metal layer was deposited with an approximate thick-

    ness of 50 nm by the electron-beam evaporation method. The

    CVD growth process started with flowing Ar of 200 sccm in thequartz-tube chamber with a pressure level of 1 Torr. The

    furnace was subsequently heated up to 750 C and stabilized over

    one hour. Pure S powder, placed upstream in the chamber, was

    heated to just above its melting temperature (115 C). The S

    vapor flowed over the Mo thin film pre-deposited on the SiO 2Si

    substrate for approximately 30 minutes. Vaporsolid reaction

    took place at an elevated temperature (750 C), leading to the

    growth of an MoS2 nanomembrane driven by increase in

    enthalpy. While the MoMo bond is relatively weak (metallic in

    Fig. 1 (a) Schematic of theCVD growthsetupshowingthe reaction of sulphur vaporwith Mo filmpre-depositedon SiSiO2 substratefor thegrowth of

    an MoS2nanomembrane. (b) Major process steps involved in the synthesis of the MoS2nanomembrane, layer transfer, and subsequent fabrication of a

    Schottky-barrier solar cell. The KOH wet etching process was used to detach the MoS 2 nanomembrane from SiO2Si substrate. (c and d) Partially

    floating MoS2nanomembrane in KOH solution. (e) Free-floating MoS2nanomembrane in KOH to be transferred onto an ITO-coated glass substrate

    for solar cell fabrication.

    7400 | Nanoscale, 2012, 4, 73997405 This journal is The Royal Society of Chemistry 2012

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J
  • 8/12/2019 Mo sep 2012

    3/7

    nature), the MoS bond is a strong ionic one and therefore more

    thermodynamically stable. As the temperature increases, Mo

    atoms gain sufficient thermal energy (exceeding the activation

    energy) and make covalent bonds with S atoms. After the

    completion of CVD growth, the sample was annealed at 1000 C

    for one hour to improve the crystalline quality and the unifor-

    mity of the resultant nanomembrane. The major steps in the

    MoS2 layer transfer sequence are depicted in Fig. 1(b) starting

    from the SiSiO2substrate, Mo metal deposition, MoS2growth,and transfer to the target substrate in which the solar cells were

    fabricated.

    The CVD-assembled MoS2 nanomembrane was transferred

    from the growth substrate (SiO2Si) to an indium tin oxide (ITO

    serving as transparent electrode) coated device substrate,

    involving a wet chemistry etching step of the SiO2 layer in the

    growth substrate to yield a freestanding MoS2nanomembrane in

    potassium hydroxide (KOH) solution. Fig. 1(c)(e) show the

    optical images of the process to separate the MoS2 nano-

    membrane from SiO2Si substrate. The MoS2 nanomembrane

    was transferred onto the transparent conducting oxide (ITO)

    substrate to fabricate the Schottky-barrier solar cell. KOH

    solution with a concentration of 15 M was prepared in which theMoS2SiSiO2substrate was immersed to etch away the SiO2. A

    slow and smooth KOH etching of SiO2 released the MoS2nanomembrane partially from the Si substrate. Within approx-

    imately 10 minutes a completely free-floating MoS2 nano-

    membrane in KOH solution was obtained. The thickness of the

    MoS2 nanomembrane is determined by that of the Mo film

    deposited by electron beam evaporation (it is hence scalable).

    The surface morphology and thickness of the as-grown MoS2nanomembrane was characterized by Atomic Force Microscopy

    (AFM, dimension 3100). Raman spectroscopy (Horiba Scien-

    tific) with a laser of 532 nm was performed on the samples to

    confirm the growth of MoS2 via identified vibrational modes.

    After transferring the MoS2 nanomembrane onto the ITO-coated glass substrate, deposition of 50 nm of Au contact was

    made by electron-beam evaporation to form a Schottky-barrier

    (with MoS2). The JV characteristics of the solar cell were

    measured under both dark and standard AM 1.5 illumination

    conditions using an Agilent B1500A semiconductor parameter

    analyzer. An Xe arc lamp was used to simulate the solar light.

    The EQE characteristics of the solar cells were measured in the

    wavelength range of 350950 nm using a Newport mono-

    chromator (equipped with the same Xe arc lamp that was used to

    measure the JV characteristics). The CV measurement was

    performed on the solar cell using a B1520A-FG Multi Frequency

    Capacitance Measurement Unit Module attached to an Agilent

    B1500A using a frequency of 1 kHz and AC perturbation of10 mV at room temperature.

    Results and discussion

    Fig. 2(a) is the micrograph showing the surface morphology of

    an MoS2 nanomembrane after being transferred onto the ITO

    substrate, as obtained from the 2D mapping measurement of an

    Atomic Force Microscope (AFM). The AFM line-scan was

    carried out, from point A to point B as shown in Fig. 2(a), to

    measure the physical thickness of the MoS2nanomembrane. The

    step height between MoS2 and the ITO substrate, as seen in

    Fig. 2(b), confirmed the thickness of the MoS2 nanomembrane to

    be about 55 nm, an average value from the measured data. We

    believe that the as-grown MoS2 nanomembrane on the SiSiO2substrate in our CVD process is highly uniform in thickness, as

    reported by Y. Zhan et al.in a similar growth experiment.37 The

    variation in AFM line-scan data (approximately 5 nm) could

    arise from deformation of the MoS2 nanomembrane as a result

    of the layer transfer process. We expect that micro-level wrinkles

    might have been formed on the MoS2nanomembrane caused by

    various types of mechanical interactions during the wet chem-

    istry etching and transferring process.Material characterization using Raman spectroscopy was

    conducted on two transferred MoS2 nanomembrane samples

    with different physical thicknesses. While the as-assembled MoS2exhibits an average thickness of55 nm, as shown in Fig. 2(b),

    stacked MoS2 nanomembranes were prepared via multiple (two

    and four, respectively, in our experiment) transfers.

    The samples were subsequently used for Raman spectroscopic

    investigation. The expected thicknesses of the two samples are

    110 nm (two transfers) and 220 nm (four transfers), respec-

    tively. Prior research results show that there is a significant

    difference in the signature modes in Raman spectra for

    Fig. 2 (a) 2D mapping micrograph of AFM surface scan obtained on an

    as-assembled MoS2 nanomembrane after transferring onto an ITO

    substrate. The line-scan measurement was performed from point A to B

    to evaluate the physical thickness of MoS2. (b) Measured step profile ofthe MoS2nanomembrane on ITO substrate.

    This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 73997405 | 7401

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J
  • 8/12/2019 Mo sep 2012

    4/7

    monolayer, bilayer and few-layer MoS2.37 In our experiment, a

    much thicker stack of MoS2 was employed to enhance the photo-

    absorption in the Schottky-barrier solar cells. Fig. 3(a) shows the

    Raman spectra of the MoS2 nanomembrane samples with

    thicknesses of110 nm and 220 nm, respectively. Two signif-

    icant Raman active modes were identified at the wavenumbers of

    383 cm1 and 408 cm1 which correspond to E12gand A1gmodes

    respectively.

    Fig. 3(b) displays schematically both E12g and A1g modes. Inthe E12g mode both Mo and S atoms vibrate along the in-plane

    direction (yet opposite to each other), whereas in the A1g mode

    the S atoms vibrate in the perpendicular-to-plane direction. The

    Raman signature peaks observed in our experiment confirm both

    in-plane and perpendicular-to-plane Raman active modes. As we

    used relatively thick MoS2 (>100 nm), there are no blue- and red-

    shifts for the E12gand A1gmodes, different from that observed in

    monolayer/bilayer MoS2 nanomembranes.37 The significant

    variation in the peak intensity of Raman spectra, as observed in

    the two examined samples, was due to the substantial difference

    in physical thickness (100 nm).

    Fig. 4(a) shows the schematic cross-sectional view of the MoS2

    nanomembraneAu Schottky-barrier solar cell. The MoS2nano-membrane was transferred onto an ITO substrate (serving as the

    window layer through which the solar cell is illuminated). Fig. 4(b)

    shows the energy band diagram of the fabricated Schottky-barrier

    solar cell structure with a stack of ITOMoS2Au.

    In general, if the metal work function is higher than that of the

    n-type semiconductor, then an ideal Schottky-barrier could be

    formed. In our experiment, Au (work function: 5.1 eV) forms abarrier on the MoS2 semiconductor (work function: 4.6 eV).

    The ITO used in this study is a heavily n-type doped transparent

    conductive oxide with a work function of4.5 eV. It forms an

    almost perfect ohmic contact with the n-type doped MoS2nanomembrane. As shown in Fig. 4(b), the Fermi level of ITO is

    above the conduction band due to its heavily doped nature.

    Although layer-dependent optical properties (e.g., change of

    bandgap with number of layers) has been demonstrated in ultra-

    thin MoS2(from a monolayer to six layers) as reported by K. F.

    Mak et al.,39 the bulk properties dominate in thicker stacks of

    MoS2. In our experiment, MoS2 nanomembrane stacks

    Fig. 3 (a) Raman spectra obtained on two samples of MoS2 nano-

    membranes with thicknesses of110 nm and 220 nm, respectively. The

    samples were prepared by multiple transfers/stacking of CVD-grown

    MoS2. (b) Schematic showing the two Raman active modes corre-

    sponding to the signature peaks observed in MoS2.

    Fig. 4 (a) Schematic cross-sectional view of the demonstrated Schottky

    barrier solar cell structure showing a stack of MoS2nanomembrane on

    an ITO substrate with Au contact. (b) Energy band diagram of the solar

    cell with formation of a Schottky-barrier between the MoS2 nano-

    membrane and Au metal contact.

    7402 | Nanoscale, 2012, 4, 73997405 This journal is The Royal Society of Chemistry 2012

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J
  • 8/12/2019 Mo sep 2012

    5/7

    (110 nm and 220 nm) were used for the fabrication of solar

    cells. Therefore layer dependency of the optical bandgap

    becomes negligible. We consider the MoS2stack with an optical

    bandgap of 1.3 eV which effectively absorbs photons in the

    wavelength region of 350950 nm. The Schottky-barrier solar

    cells are illuminated through the ITO window layer which does

    not absorb any photons in the visible light spectral region, as its

    bandgap is 3.8 eV and more than 95% optically transparent.

    Incident photons generate electronhole pairs in the MoS2nanomembrane and electrons are subsequently excited to the

    conduction band (EC) of MoS2. As the work function difference

    between Au and MoS2 results in a Schottky-barrier, the photo-

    excited electrons move from the EC of MoS2 towards the Fermi

    level of ITO (where they can be collected to the load). A larger work

    function difference between Au and n-type MoS2would generate a

    higher electric field in the depletion region of MoS2. The built-in

    electric field facilitates the dissociation of the photo-excited elec-

    tronhole pairs, transporting separated carriers towards ITO and

    Au contacts. The processes of photo-excitation in MoS2 and elec-

    tron transport from MoS2to ITO are demonstrated in Fig. 4(b).

    Fig. 5(a) and (b) show the dark and illuminated JVcharac-

    teristics of the fabricated Schottky-barrier solar cells with MoS2thicknesses of110 nm and 220 nm, respectively. The output

    power as a function of voltage in each solar cell is shown in the

    insets of Fig. 5(a) and (b), respectively. For the Schottky-barrier

    solar cell with MoS2thickness of 110 nm key metricsJSC(short-

    circuit current density), VOC (open-circuit voltage), PMAX(maximum power), FF (fill factor) and h (photo-conversion

    efficiency) are 2.52 mA cm2, 590 mV, 1.22 104 W, 0.48 and

    0.7%, respectively. For the solar cell with MoS2 thickness of

    220 nm, JSC, VOC, PMAX, FFand h are 5.37 mAcm2, 597 mV,

    2.99 104 W, 0.55 and 1.8% respectively. It is observed that an

    increase of MoS2thickness helps to significantly enhanceJSCof

    the solar cell (by 113%) due to more efficient photo-absorption in

    the thicker MoS2 stack, and hence creates a noteworthy changein overall device performance. It should be noted that there is no

    sizable change in VOC between solar cells with different thick-

    nesses of MoS2 stacks (merely 1% change). It is evident that an

    increase in thickness of the MoS2 stack would not cause a notable

    change in the band structure (hence no change in the Fermi level)

    due to the absence of layer-dependent properties.

    Fig. 5(c) shows the schematic to compare carrier recombina-

    tion in conventional and layered semiconductors. Process 1

    represents optical excitation leading to electronhole pair

    generation. In conventional semiconductors, the amount of

    interface states (recombination centers) in the bandgap is huge

    due to unsaturated chemical dangling bonds which capture

    electrons from the conduction band (Process 2) and holes fromthe valence band (Process 3), leading to recombination. In

    layered semiconductors, the amount of interface states is signif-

    icantly less due to the self-saturated surface bonds which do not

    actively participate in the recombination process. Thus, we

    attribute the enhancement in solar cell performance to the

    increase in MoS2thickness which improved photon absorption.

    We expect that as a layered material, defect density in MoS2does

    not increase drastically with increasing thickness compared to a

    conventional semiconductor.

    Fig. 6 shows External Quantum Efficiency (EQE) spectra of

    the two Schottky-barrier solar cells with different MoS2

    thicknesses in the wavelength range of 350950 nm. The

    maximum EQE values of 41% and 52% were measured for solar

    cells with 110 nm and 220 nm of MoS2stack, respectively, at

    wavelength of 600 nm. The increase in MoS2 thickness plays a

    vital role in photo-absorption. Our prior work on MoS2-based

    Fig. 5 Measured dark and illuminated JV characteristics of the

    Schottky-barrier solar cells with a stack of MoS2nanomembrane having

    thicknesses of (a) 110 nm and (b) 220 nm. Inset of each figure showsthe power output as a function of voltage for the fabricated solar cell. (c)

    Recombination process via electronhole capture by interfacial trap

    states present in conventional and layered semiconductors, showing

    significantly fewer traps in the latter. This leads to reduced probability of

    carrier recombination in a layered semiconductor.

    This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 73997405 | 7403

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J
  • 8/12/2019 Mo sep 2012

    6/7

    solar cells34 employing a combination of monolayer/multi-layerstacks of MoS2on TiO2(serving as electron acceptor) has shown

    photon absorption of MoS2 in the wavelength region of 350

    750 nm. The measured EQE spectra in the extended wavelength

    region, as compared with the previously reported values (up to

    950 nm),34 suggests that an MoS2 nanomembrane with an

    optical bandgap of 1.3 eV could effectively absorb photons in

    both visible and near infrared spectral regions due to the absence

    of the layer-dependent optical properties of MoS2 as governed by

    its physical thickness.

    Fig. 7(a) shows the MottSchottky (C2V) characteristics of

    the Schottky-barrier solar cells. The C2Vcharacteristics were

    used to estimate the Schottky-barrier heights associated with

    each of the two fabricated solar cells. Fig. 7(b) shows schemati-cally the energy-level alignment at the MoS2Au interface under

    zero, forward and reverse voltage bias conditions. Since the

    MoS2 samples are n-type doped, Fermi levels in MoS2 and Au

    shift upwards and downwards, respectively, with reference to the

    equilibrium Fermi level at zero voltage bias, yielding

    the maximum barrier height. In the case of reverse voltage bias,

    the situation becomes opposite. We found that both solar cells

    (with different MoS2 thickness) resulted in nearly the same

    Schottky-barrier height (1.0 eV), as measured from the linear

    extrapolation in the reverse-biasing region (from 1 V to 4 V). It

    can be understood that a change of the MoS2should not make a

    difference in the Schottky-barrier height. While we expect that

    MoS2thickness (down to a few layers) may change the Schottky-barrier height, the layer dependence becomes insignificant in the

    nanomembranes used in our experiment. The observed linear

    behavior of the MottSchottky plots suggests that the MoS2nanomembranes possess a relatively low density of surface

    defects.

    J. M. Lutheret al.reported a Schottky-barrier solar cell using

    PbSe quantum dots.2 The result showed significant variation in

    the voltage-dependent C2 for two PbSe samples (thickness

    difference > 300 nm).2 In our study, MoS2Au solar cells (with

    active layer thickness difference of 100 nm) exhibit no significant

    change in the C2 values with reverse bias. This could be

    attributed to the layered MoS2 nanomembrane (defect-free or

    with negligibly low density of defects) as compared with

    conventional semiconductors in which unsaturated danglingbonds make the MS interface highly defective. In layered

    semiconductors the crystal lattice planes are weakly bound by

    van der Waals interactions and hence the surface exhibits self-

    saturated electronic bonds which are no longer active for surface

    reaction with the environment. The nature of a self-saturated

    surface in layered MoS2 nanomembrane eliminates Fermi level

    pinning and interfacial oxide layer formation (both could impede

    the transport significantly). The voltage-bias-dependent C2

    values are almost identical in both device samples (with different

    thicknesses) due to negligible change in the trap states at MoS 2

    Au interface where carriers are stored and released, affecting the

    junction capacitance. This is a key appealing property of layered

    semiconductors for solar cell applications as compared withbulk/thin film semiconductors which possess a particular type of

    electronic bonding along and across the crystal planes (leading to

    defective bulk and surface upon cleaving).

    We project that the performance of the MoS2nanomembrane-

    based Schottky-barrier solar cell could be further improved.

    Layered semiconductors are typically very sensitive to growth

    conditions such as temperature and pressure. Temperature fluc-

    tuation across the growth furnace plays a significant role in

    influencing the solidvapor reaction between pre-deposited Mo

    and S vapor, and eventually the electronic quality of the MoS2nanomembrane. In addition, defects and impurities introduced

    Fig. 6 Measured EQE spectra of the Schottky-barrier solar cells with a

    stack of MoS2 nanomembranes having thicknesses of 110 nm and

    220 nm, respectively, in the wavelength region of 350950 nm. This

    confirms an optical bandgap of approximately 1.3 eV for the prepared

    MoS2 nanomembranes.

    Fig. 7 (a) The measured MottSchottky (C2V) plot of the Schottky-

    barrier solar cells with a stack of MoS2 nanomembranes having thick-

    nesses of 110 nm and 220 nm, respectively, showing a linear C2

    dependency (shown in the circled region) on the applied reverse bias, VR.

    (b) Schematic showing MoS2Au interfacial energy level change at zero

    (equilibrium), forward, and reverse bias conditions and the effect on

    Schottky barrier height.

    7404 | Nanoscale, 2012, 4, 73997405 This journal is The Royal Society of Chemistry 2012

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J
  • 8/12/2019 Mo sep 2012

    7/7

    during the CVD growth and deformation formed during layer

    transfer would all contribute to reduced carrier mobility and

    degraded carrier transport properties. We expect that further

    optimization of the growth process, use of a cleaner fabrication

    environment, and direct assembly of MoS2 nanomembranes on

    an ITO substrate would produce less defective materials, leading

    to improved solar cell performance.

    Conclusions

    Schottky-barrier solar cells have been demonstrated using layer-

    assembled, CVD-grown MoS2 semiconductor nanomembranes

    as the critical photo-active layer. While relatively thick MoS2nanomembranes are used in this experiment, the material is

    scalable via a controlled CVD growth process and large-area

    solar cell fabrication is potentially feasible. The dependency of

    solar cell performance on MoS2 thickness was evaluated. Further

    improvement in solar cell performance is expected through

    optimizing the growth process, eliminating structural deforma-

    tion (induced during fabrication), and minimizing the amount of

    surface defects and impurities. The reported results would

    promote continued efforts towards developing highly efficientSchottky-barrier solar cells taking advantage of the unique

    interfacial properties of layered semiconductor nanostructures.

    References

    1 Z. B. Zhou, R. Q. Cui, Q. J. Pang, G. M. Hadi, Z. M. Ding andW. Y. Li, Sol. Energy Mater. Sol. Cells, 2002, 70, 487493.

    2 J. M. Luther, M. Law, M. C. Beard, Q. Song, M. O. Reese,R. J. Ellingson and A. J. Nozik, Nano Lett., 2008, 8, 34883492.

    3 B. Ghosh, M. Das, P. Banerjee and S. Das,Solid State Sci., 2009,11,461466.

    4 M. Soylu and F. Yakuphanoglu, Thin Solid Films, 2011, 519, 19501954.

    5 L. Zhang, Y. Jia, S. Wang, Z. Li., C. Ji, J. Wei, H. Zhu, K. Wang,

    D. Wu, E. Shi, Y. Fang and A. Cao, Nano Lett., 2010,10, 35833589.6 Y. Ye, Y. Dai, L. Dai, Z. Shi, N. Liu, F. Wang, L. Fu, R. Peng,X. Wen, Z. Chen, Z. Liu and G. Qin, ACS Appl. Mater. Interfaces,2010, 2, 34063410.

    7 C. C. Chen, M. Aykol, C. C. Chang, A. F. J. Levi and S. B. Cronin,Nano Lett., 2011,11, 18631867.

    8 J. Wei, Y. Jia, Q. Shu, Z. Gu, K. Wang, D. Zhuang, G. Zhang,Z. Wang, J. Luo, A. Cao and D. Wu,Nano Lett., 2007,7, 23172321.

    9 C. Xie, P. Lv, B. Nie, J. Jie, X. Zhang, Z. Wang, P. Jiang, Z. Hu,L. Luo, Z. Zhu, L. Wang and C. Wu, Appl. Phys. Lett., 2011, 99,133113.

    10 G. Fan, H. Zhu, K. Wang, J. Wei, X. Li, Q. Shu, N. Guo and D. Wu,ACS Appl. Mater. Interfaces, 2011,3, 721725.

    11 C. Xie, J. Jie, B. Nie, T. Yan, Q. Li, P. Lv, F. Li, M. Wang, C. Wu,L. Wang and L. Luo, Appl. Phys. Lett., 2012, 100, 193103.

    12 G. H. Glover, Solid State Electrochem., 1973,16, 973978.13 F. Dubecky and B. Olejnkova,J. Appl. Phys., 1991, 69, 1769.14 C. Ghezzit, E. Gombiaz and R. Moscaf, Semicond. Sci. Technol.,

    1991, 6, 3133.15 A. Singh, P. Cova and R. A. Masut, J. Appl. Phys., 1994, 76, 2336.16 F. Lu, D. Gong, J. Wang, Q. Wang, H. Sun and X. Wang, Phys. Rev.

    B: Condens. Matter Mater. Phys., 1996, 53, 46234629.17 D. Vanmaekelbergh, A. Koster and F. I. Marn,Adv. Mater., 1997,9,

    575578.18 Y. Fukushima, K. Ogisu, M. Kuzuharaand K. Shiojima, Phys. Status

    Solidi C, 2009, 6, 856859.19 S. Tongay, T. Schumann and A. F. Hebard, Appl. Phys. Lett., 2009,

    95, 222103.20 F. Leonard and J. Tersoff, Phys. Rev. Lett., 2000, 84, 46934696.21 A. Dimoulas, P. Tsipas, A. Sotiropoulos and E. K. Evangelou, Appl.

    Phys. Lett., 2006, 89, 252110.22 T. Nishimura, K. Kita and A. Toriumi, Appl. Phys. Lett., 2007, 91,

    123123.23 Y. Zhou, W. Han, Y. Wang, F. Xiu, J. Zou, R. K. Kawakami and

    K. L. Wang,Appl. Phys. Lett., 2010, 96, 102103.24 S. J.Oh,D. K.Kim and C.R. Kagan, ACS Nano, 2012, 6, 43284334.25 Y. Yoon, K. Ganapathi and S. Salahuddin, Nano Lett., 2011, 11,

    37683773.26 Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang,

    X. Chen and H. Zhang, ACS Nano, 2012, 6, 7480.

    27 H. S. Lee, S. W. Min, Y. G. Chang, M. K. Park, T. Nam, H. Kim,J. H. Kim, S. Ryu and S. Im, Nano Lett., 2012,12, 36953700.

    28 D. J. Late, B. Liu, H. S. S. R. Matte, V. P. Dravid and C. N. R. Rao,ACS Nano, 2012, 6, 56355641.

    29 Y. Zhang, J. Ye, Y. Matsuhashi and Y. Iwasa, Nano Lett., 2012,12,11361140.

    30 S. Ghatak, A. N. Pal and A. Ghosh,ACS Nano, 2011,5, 77077712.31 K. Chang and W. Chen, Chem. Commun., 2011,47, 42524254.32 K. Chang, W. Chen, L. Ma, H. Li, H. Li, F. Huang, Z. Xu, Q. Zhang

    and J. Y. Lee, J. Mater. Chem., 2011, 21, 62516257.33 K. Chang and W. Chen, J. Mater. Chem., 2011, 21, 1717517184.34 M. Shanmugam, T. Bansal, C. Durcan and B. Yu, Appl. Phys. Lett.,

    2012, 100, 153901.35 K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich,

    S. V. Morozov and A. K. Geim, Proc. Natl. Acad. Sci. U. S. A. , 2005,102, 10451.

    36 J. N. Coleman, M. Lotya, A. ONeill, S. D. Bergin, P. J. King,U. K.Young, A. Gaucher, S. De, R. J. Smith and I. V. Shvets,et al., Science, 2011, 331, 568571.

    37 Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan and J. Lou, Small, 2012, 8,966971.

    38 H. Liu and P. D. Ye, IEEE Electron Device Lett., 2012, 33, 546548.

    39 K.F. Mak, C. Lee, J. Hone, J. Shanand T. F. Heinz, Phys. Rev. Lett.,2010, 105, 136805.

    40 B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis,Nat. Nanotechnol., 2011,6, 147150.

    41 J. R. Lince, D. J. Carre and P. D. Fleischauer, Phys. Rev. B: Condens.Matter Mater. Phys., 1987, 36, 16471656.

    This journal is The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 73997405 | 7405

    View Article Online

    http://dx.doi.org/10.1039/C2NR32394J