Nano Research 1 DOINano 10.1007/s12274Res -015-0961-1

Enhanced thermoelectric properties of topological crystalline insulator PbSnTe nanowires grown by a vapor transport approach

Enzhi Xu1, Zhen Li1, Jaime Avilés Acosta1, Nan Li2, Brian Swartzentruber3, ShiJian Zheng2, Nikolai Sinitsyn4,

H. Htoon2, Jian Wang5 and Shixiong Zhang1 (*) Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-015-0961-1 http://www.thenanoresearch.com on December. 1, 2015

© Tsinghua University Press 2015

Just Accepted

This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance, which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP) provides “Just Accepted” as an optional and free service which allows authors to make their results available to the research community as soon as possible after acceptance. After a manuscript has been technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Please note that technical editing may introduce minor changes to the manuscript text and/or graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event shall TUP be held responsible for errors or consequences arising from the use of any information contained in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI®), which is identical for all formats of publication.

64Nano Res.

Enhanced thermoelectric properties of topological crystalline insulator PbSnTe nanowires grown by a vapor transport approach

Enzhi Xu1, Zhen Li1, Jaime Avilés Acosta1, Nan Li2, Brian Swartzentruber3, ShiJian Zheng2 , Nikolai Sinitsyn4, H. Htoon2, Jian Wang5 and Shixiong Zhang1*

1Department of Physics, Indiana University, Bloomington, Indiana 47405, USA.

2Center for Integrated Nanotechnologies, Materials, Physics and Applications Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA

Single-crystalline PbSnTe nanowires were synthesized via a vapor 3Center for Integrated Nanotechnologies, Sandia National Laboratories, Albuquerque, New Mexico transport approach, and thermoelectric devices were fabricated to 87185, USA measure electrical conductivity, thermopower and thermal conductivity 4Theoretical Division, Los Alamos National on the same individual nanowires. The PbSnTe nanowires show Laboratory, Los Alamos, New Mexico 87545, USA enhanced thermoelectric figure of merit ZTs than bulk single crystals 5MST-8, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA (Orihashi et al. J. Phys. Chem. Solids 2000, 61, 919-923), owing to both

the improved thermopower and the suppressed thermal conductivity.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research Nano Research

DOI (automatically inserted by the publisher) Research Article

Enhanced thermoelectric properties of topological crystalline insulator PbSnTe nanowires grown by a vapor transport approach

Enzhi Xu1, Zhen Li1, Jaime Avilés Acosta1, Nan Li2, Brian Swartzentruber3, ShiJian Zheng2 , Nikolai 4 2 5 1 Sinitsyn , H. Htoon , Jian Wang and Shixiong Zhang (*)

1 Department of Physics, Indiana University, Bloomington, Indiana 47405, USA. 2 Center for Integrated Nanotechnologies, Materials, Physics and Applications Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA 3 Center for Integrated Nanotechnologies, Sandia National Laboratories, Albuquerque, New Mexico 87185, USA 4. Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA 5. MST-8, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA

Received: day month year ABSTRACT Revised: day month year Bulk telluride (PbTe) and its alloy compounds are well-known Accepted: day month year for electric power generation. Among all these alloys, (automatically inserted by PbSnTe has recently been demonstrated to host unique topological surface the publisher) states that may have improved thermoelectric properties. Here we report on a vapor transport growth and thermoelectric study of high quality single-crystalline PbTe and PbSnTe nanowires. The nanowires were grown

along <001> direction with dominant {100} facets and their chemical compositions are strongly dependent on the substrate position in the growth © Tsinghua University Press reactor. We measured the thermopower, electrical and thermal conductivities of and Springer-Verlag Berlin the same individual nanowires to determine their thermoelectric figure of merit Heidelberg 2014 ZTs. In comparison with bulk samples, the PbSnTe nanowires show both improved thermopower and suppressed thermal conductivity, leading to enhanced ZTs of ~0.018 and ~0.035 at room temperature. The enhanced

thermopower may arise from the unique topological surface states and the KEYWORDS suppression of thermal conductivity is possibly due to the increased phonon-surface scattering. Compared with the PbTe nanowires, the PbSnTe has Type keywords lower thermopower but significantly higher electrical conductivity. Our work PbSnTe, highlights nanostructuring in combination with alloying as an important thermoelectrics, approach to enhancing thermoelectric properties. topological crystalline insulator, nanowire 1 Introduction promising approaches to harvesting renewable energy [1, 2]. The energy conversion efficiency is

Address correspondence to [email protected] Converting waste heat into electric power quantified by the so-called figure of merit ZT of through the Seebeck effect is one of the most the material that is used in the thermoelectric

2Nano Res. generator. The ZT is dependent on the material study the thermoelectric properties of PbSnTe properties, i.e. thermopower S, electrical nanowires. As a prototypical PbTe-based alloy, conductivity s, and thermal conductivity κ in a Pb1-xSnxTe can form a complete solid solution relation of ZT=S2 σT/κ. The small bandgap (0≤x≤1) and its bulk bandgap can be tuned (PbTe) with a continuously from 0.3 eV to -0.18 eV by changing rock-salt crystal structure is one of the most x from 0 to 1. The Pb1-xSnxTe alloy with a Sn widely studied thermoelectric materials [3]. concentration of x>0.37 has a negative bulk While the ZT of pure PbTe bulk samples is less bandgap and has recently been demonstrated to than 1, it can be dramatically increased by possess unique topological surface states that are alloying with other compounds or by doping protected by crystalline mirror symmetry [47-49]. with impurities [4-8]. Indeed, bulk Pb1-xSnxTe Recent theoretical studies have suggested that alloys exhibit a maximum ZT of ~ 1 [9, 10] and topological surface states may have enhanced the doped PbTe1-xSex alloys have a peak ZT as thermoelectric properties than bulk states owing high as ~ 1.8 [7, 8]. The enhanced thermoelectric to their unique electronic band structures [50, 51]. properties are attributed to the fact that alloying Nanowires are anticipated to have more not only reduces thermal conductivity by advantages than bulk samples in the sense that inducing additional phonon scatterings, but also they have significantly higher influences thermal power and electrical surface-area-to-volume ratio which could conductivity favorably by modifying electronic magnify contributions from surface states. band structures [3]. Single-crystalline PbSnTe nanowires were Another effective approach to enhancing ZT grown by a vapor transport approach with Au is to fabricate nanostructured materials [11, 12]. nanoparticles as catalysts. The nanowires were Quantum confinement effect in nanostructures grown along <001> direction of the rock-salt gives rise to an increased thermopower, while crystal structure and their chemical compositions surface/boundary scattering of phonons are strongly dependent on the substrate position suppresses thermal conductivity [13, 14]. As an in the growth reactor. The thermopower, example, PbTe-based nanostructures, such as electrical and thermal conductivities were quantum dot superlattices [15] and dual-phase systematically measured on the same nanocomposites [16], have shown improved ZTs single-nanowire devices to accurately determine than their bulk counterparts. One-dimensional their figure of merit ZTs. In comparison with bulk nanowires (NWs) are another format of samples of nearly the same chemical composition, nanostructures that have enhanced PbSnTe nanowires show both enhanced thermoelectric properties as demonstrated in thermopower and reduced thermal conductivity, many material systems [17-38]. The PbTe NWs leading to a significantly improved ZT. The same have been fabricated by a variety of synthetic growth and thermoelectric studies were carried approaches, including chemical vapor transport out on the undoped PbTe nanowires for [22, 27, 34, 39], template-directed comparison. electrodeposition [40] and hydrothermal process [41, 42]. Thermoelectric studies have indeed 2 Experimental details shown enhanced thermopower and/or reduced The PbTe and PbSnTe NWs were grown by a vapor thermal conductivity of NWs than their bulk transport approach that was widely employed for counterparts [22, 27, 34, 38, 39, 41-46]. the controlled synthesis of single-crystalline In this work, we combined an alloying nanostructures of SnTe [52-57] and its related approach with the nanostructuring strategy to compounds [58]. Figure 1 (a) illustrates a schematic

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research Nano Res. 3 of the quartz tube reactor where the growth took 3.1 Nanowire growth and characterizations place. In contrast to the earlier chemical vapor As shown in Figure 1 (b), a single vapor transport growth of PbTe NWs in which the transport growth gives rise to a mixture of precursors are lead chloride and [22, 27, nanowires and micro-crystals. The dominant 34, 39], we used PbTe (Alfa Aesar, 99.999%) and facets of both PbTe and PbSnTe are {100} planes SnTe (Alfa Aesar, 99.999%) powders as precursors. of the cubic structure, as evidenced by the Argon gas with a flow rate of 15 sccm was rectangular shape of each surface [Fig. 1 (b) for introduced and the pressure inside the tube PbSnTe and Fig. S1 for PbTe]. This is consistent furnace was kept at ~ 33 Torr. Temperatures in the three zones were maintained at 820/750/600 ˚C for with recent density functional theory 30 minutes during the growth of PbTe NWs, and at calculations which suggest that the {100} 760/710/575 ˚C for 30 minutes during the growth of surfaces of PbTe have the lowest surface PbSnTe NWs. For the PbSnTe growth, the PbTe and energies among all low-index planes [59]. SnTe precursors with an atomic ratio of ~1:1 were While Au alloy particles are observed at the tip placed in the same boat and hence were at nearly of some nanowires as shown in the lower left the same temperature. Au nanoparticles of 20 nm panels of Fig. 1 (b) and Fig. S1, some other diameter (Ted Pella) were deposited on the NWs are free of particles [lower right panel of substrates and acted as metal catalysts for the Fig. 1 (b) and Fig. S1]. We note that catalyst is vapor-liquid-solid growth. To avoid a possible not always required in the vapor phase growth coating of substrate material on nanowire surfaces of lead chalcogenide nanowires even though as observed in previous chemical vapor deposition they have isotropic, cubic crystal structures (e.g. of PbTe [22], we used the relatively more stable in the dislocation-driven growth process) [60, tungsten foils as substrates. The nanowire 61]. To test if our nanowire growth is indeed morphology and chemical composition were assisted by Au, we further performed a growth studied by scanning electron microscope (SEM, at the same condition but without Au Quanta FEI) and energy-dispersive X-ray nanoparticles. As shown in Fig. S2, the spectrometer (Oxford INCA EDX and AMETEK catalyst-free growth only to high density EDAX), respectively. The nanowire crystalline of micro-crystals, which suggests that the structures were characterized using transmission nanowires were indeed grown through an electron microscope (FEI Tecnai F30). Thermoelectric platforms that consist of a heater Au-catalyzed vapor-liquid-solid process. The absence of alloy particle on a nanowire may and four electrodes were fabricated on SiNx/SiO2/Si substrates via e-beam lithography and e-beam arise from the diffusion of Au atoms into the evaporation of 10 nm Ti & ~ 90 nm Au. Individual nanowire during the growth process [62]. PbTe and PbSnTe nanowires were picked up from The chemical composition of PbSnTe the growth substrates using a nanomanipulator micro-crystals and nanowires has a strong and were then transferred onto the thermoelectric dependence on the substrate position in the platforms. For the PbTe devices, a layer of ~120 nm growth reactor. As shown in Figure 1 (a), three Ni & ~200 nm Au was deposited on top of the four ~ 3-inch long substrates were placed in three electrodes after e-beam lithography to wrap unsealed quartz tubes in the growth reactor around the nanowires (nanowires were rinsed in side by side from left to right. The substrates 1, buffered oxide etch solution for 20 s before metal 2, and 3 cover a range of x: 0 ~ 3 inches, 3.1 ~ deposition). For the PbSnTe devices, a layer of ~10 6.1 inches and 6.2 ~ 9.2 inches, respectively (the nm Ti & ~300 nm Au was deposited on top of the gap between two substrates was about 0.1 four electrodes after the nanowires were rinsed in inch). As seen in Figure 1 (c), the Pb: (Pb+Sn) 1% HCl for 20 s. The devices were then annealed at ratio increases from the left side of each 300 °C for 20 s to promote good contacts. substrate to the right side. However, it

decreases dramatically

3 Resutls and discussion

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

Figure 1 (a) a schematic of the vapor transport reactor for the growth of PbTe and PbSnTe nanowires. The substrates were placed in unsealed small quartz tubes. The position and the size of the substrates are not drawn to scale. (b) SEM images of PbSnTe nanowires and nano/micro-crystals. The lower left panel shows a typical nanowire with an Au alloy particle at its tip, while the lower right panel shows a nanowire free of particle. The scale bars are: 100 nm (left) and 200 nm (right). The chemical composition ratio (c) Pb: (Pb+Sn) and (d) Te: (Pb+Sn) as a function of position. The error bar of the position x on each substrate is ~0.2 inches. from one substrate to the next, i.e. from x ~ 3 result, the microcrystals/nanowires deposited inches to ~ 3.1 inches and from x ~ 6.1 inches to on the left side of the first substrate (i.e. near ~ 6.2 inches. The composition ratio covers a the precursors) have a Pb: (Pb+Sn) ratio < 0.5. broad range of ~ 0.15 to ~ 0.55 although the Since the vapors were carried from left to right ratio of PbTe: (PbTe+SnTe) precursors is ~ 0.5. by the inert Ar gas, after less Pb (or more Sn) In contrast, the Te composition does not show a vapor was consumed at the left side, the vapor clear position/substrate dependence [Te: concentration of Pb: (Pb+Sn) increases towards (Pb+Sn) ~ 1 in Figure 1 (d)]. The position the right side of the substrate, leading to an dependence of the Pb: (Pb+Sn) ratio can be increased ratio in the deposited qualitatively understood by comparing the microcrystals/nanowires. On the left side of the melting points of the two precursors: PbTe has second substrate, the vapor of high Pb a slightly higher melting point than SnTe (924 concentration from the first quartz tube [pink oC versus 790 oC), so SnTe sublimates more arrow in Figure 1 (a)] mixes with the vapor of rapidly than PbTe at the same temperature low Pb concentration from the outside of the (~710 oC) and hence the vapor concentration of quartz tube (purple arrow). As a result, the Pb: Pb is less than Sn near the precursors. As a (Pb+Sn) ratio of the microcrystals/nanowires

Nano Res. 5 deposited on the left side of the second boundary was observed. The selected area substrate (x ~ 3.1 inches) is lower than that on diffraction patterns in Fig. 2 (b) and (e) clearly the right side of the first substrate (x~3 inches). show that the NWs are single crystalline and The same scenario is applicable to substrates 2 their growth direction is along <001>. The high and 3. crystalline quality is further confirmed by the Microstructural characterizations using high-resolution TEM images shown in Figure 2 transmission electron microscopy (TEM) (c) and (f). The interplanar spacing between the suggest that both the PbTe and PbSnTe {200} planes were derived from the fast Fourier nanowires are single crystalline. Figure 2 (a) transform (FFT). It is ~ 6.4 Å for PbTe and ~ 6.3 and (d) are conventional bright field strain Å for PbSnTe, in good agreement with bulk contrast TEM images, in which no grain samples [63].

Figure 2 (a) A low magnification TEM image and (b) a selected area diffraction pattern of a typical PbTe nanowire; (c) a high resolution TEM image of another typical PbTe nanowire; (d) a low magnification TEM image and (e) a selected area diffraction pattern of a typical PbSnTe nanowire; (f) a high resolution TEM image of another PbSnTe nanowire. Fast Fourier Transform filtering was applied to obtain the lattice-resolved TEM images.

3.2 Thermoelectric measurement of substrate in the same growth so that their chemical individual nanowires composition ratios of Pb: (Pb+Sn) are close to ~ 0.5. Figure 3 (a) shows a typical thermoelectric device The nanowire dimensions were summarized in used for a correlated measurement of S, σ and κ on table 1. Two-probe current-voltage measurements the same nanowire. We measured three PbSnTe indicate good Ohmic contacts of the devices (Fig. NWs (denoted as PbSnTe NW-1, -2 and -3) along S3 in the Supplementary Information). In the with three undoped PbTe NWs (denoted as PbTe thermopower measurement, we applied an electric NW-1, -2, and -3) for comparison. The three PbSnTe current through the resistive heater to create Joule NWs were taken from the right side of the third heating, and then measure the thermal voltage ∆V

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research 6Nano Res. and the temperature difference ∆T between the two dramatically as the temperature decreases [Figure 3 electrodes B and C. Fig. S4 in the supplemental (b)]. information shows the ∆V - ∆T curves taken on a The electrical conductivities of PbTe NW-1 and typical nanowire device at different measurement PbSnTe NW-1 and -2 were determined by a temperatures from 25 to 300 K. The thermopower standard four-probe resistance measurement of a device is calculated as S=∆V/∆T, and the (four-probe measurement was not performed on thermopower of the nanowire was determined by PbTe NW-2, -3 and PbSnTe NW-3 due to device subtracting the contributions from the metal failure). As shown in Figure 3 (c), the electrical electrodes. conductivity of PbTe NW-1 at 300 K is ~1290 S/m, an order of magnitude lower than the bulk value Table 1 The dimensions of PbTe and PbSnTe nanowires (~25641 S/m) [64]. This difference suggests that the measured in this work. cation vacancy density is much lower in our nanowires than in the bulk samples. Indeed, the Nanowire Width (nm) Thickness (nm) electrical conductivity decreases during cooling, indicating a lightly doped semiconducting PbTe NW-1 217 281 behavior. In contrast, the PbSnTe NWs show a PbTe NW-2 222 124 metallic behavior [Figure 3 (c)], manifesting a degenerately doped nature. The electrical PbTe NW-3 232 80 conductivities at 300 K are on the same order of magnitude as the bulk Pb0.5Sn0.5Te single crystals [9], PbSnTe NW-1 459 170 confirming their comparable charge carrier densities and chemical compositions. As a result, PbSnTe NW-2 207 252 the much higher thermopower of PbSnTe NWs in PbSnTe NW-3 156 97 comparison with bulk crystals should be attributed to other factors. The thermopowers of both PbTe and PbSnTe We note that undoped SnTe NWs also show nanowires are positive [Figure 3 (b)], indicating enhanced thermopower than bulk samples [65]. that the majority charge carriers are holes for all Although the precise nature is not clear, the nanowires. We note that both positive and negative enhancement in NW samples may be related to the thermopowers were reported in undoped PbTe topological surface states possessed by both SnTe nanowires, depending on the growth methods [34, and PbSnTe. The thermopower of a material 38, 39, 41-43, 45, 46]. The p-type conductivity in our depends on its electronic band structure and can be nanowires suggest the existence of native defects written as 2 2 such as Pb/Sn vacancies that act as acceptors. The S=-[∫v τ(E-Ef)∂f0/∂ED(E)dE]/[eT∫v τ∂f0/∂ED(E)dE], thermopowers of the three PbTe nanowires at 300 K where v is the electron/hole velocity, τ is the are: 372 µV/K (PbTe NW-1), 354 µV/K (PbTe NW-2) relaxation time, Ef is the Fermi energy, f0 is the and 399 µV/K (PbTe NW-3), which are about 40% Fermi-Dirac distribution, and D(E) is the density of higher than the bulk value (~265 µV/K) [64]. An states (DOS). A large thermopower can be realized enhancement of thermopower was also observed in by creating a delta function in the DOS. The {100} the PbTe nanowires grown by hydrothermal surfaces of the SnTe-based topological crystalline process [41, 42, 46], stress-induced method [45] and insulators (TCIs) indeed have Van-Hove lithographically patterned nanowire singularities in their DOS [47, 48]. In other words, electrodeposition [43]. The room temperature the topological surface may have a significantly thermopowers of PbSnTe NWs are: ~ 33 µV/K higher thermopower than the interior of the TCI. (PbSnTe NW-1), 26 µV/K (PbSnTe NW-2) and 24 Since nanowires have a large µV/K (PbSnTe NW-3), about 2~3 times higher than surface-area-to-volume ratio than bulk samples, they could have enhanced thermopower. Future that of the bulk Pb0.5Sn0.5Te single crystals (~10.4 µV/K) [9]. In contrast to PbTe NWs, the theoretical study of the thermopower of topological thermopower of PbSnTe shows a strong surfaces would be of great importance to temperature dependence, i.e. it decreases understand the observed experiment results.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

Figure 3 (a) An SEM image of a typical thermoelectric device. Temperature dependent (b) thermopower S, (c) electrical conductivity σ, and (d) power factor S2σ of PbTe and PbSnTe NWs. The corresponding data of bulk samples reported in reference [64] (Heremans et al.) and reference [9] (Orihashi et al.) are shown for comparison.

The power factor of PbTe NW is limited by its suppressed carrier mobility is likely due to the low electrical conductivity, while that of PbSnTe enhanced scattering of charge carriers by surface NWs is magnified by its enhanced thermopower. trap states in nanowires [23]. In contrast, the power As shown in Figure 3 (d), the room temperature factors of the two PbSnTe NWs at room power factor of PbTe NW is only ~ 179 µWm-1K-2, temperature are 154 µWm-1K-2 (NW-1) and an order of magnitude lower than the bulk value 104 µWm-1K-2 (NW-2), both of which are much (~1800 µWm-1K-2) [64]. This result agrees higher than that of the bulk single crystal (~ qualitatively with the previous studies on lead 41 µWm-1K-2) [9]. The difference in power factors chalcogenide (e.g. PbTe and PbSe) nanowires [23, between the two PbSnTe NWs highlights the 41, 43, 45]. The lower electrical conductivity in importance of measuring thermopower, electrical nanowires than in bulk samples may arise from a and thermal conductivities on the same nanowire reduced charge carrier density and/or a suppressed in order to accurately determine the figure of merit carrier mobility. As discussed earlier, the lower ZT. charge carrier density may be due to a smaller We further measured the thermal conductivity concentration of Pb vacancies (p-type defects) in using a self-heating method [66], in which the our high quality single crystalline nanowires. The nanowire was heated up by passing a high electric

8Nano Res. current I. Thermal conductivity was extracted by measuring the mean temperature increase ∆TM of the nanowire as a function of the Joule heating power I2R based on the relation: κ=(I2Rl)/(12∆TMA) in the limit of ∆R<

We decomposed the total thermal conductivity into the electronic part κe and the lattice part κlatt, i.e. κtot=κe+κlatt. According to the Wiedemann-Franz law,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research Nano Res. 9 the electronic component can be written as a function of electrical conductivity: κe=LσT, where L is the Lorentz number. Given the degenerately doped nature of the PbSnTe NWs, we used L=2.44×10-8 V2K-2 [9] to calculate κe and then determined the lattice thermal conductivity κlatt=κtot-κe. As shown in Figure 4 (c), κlatt at 300 K are ~ 0.28 W m-1 K-1 for PbSnTe NW-1 and ~ 0.59 W m-1 K-1 for NW-2, lower than the bulk value of ~ 1.1 W m-1 K-1 [9]. This further suggests that the suppression of thermal conductivity is primarily due to the enhanced surface scattering of phonons in nanowire samples. We finally calculated the figure of merit ZT of the PbSnTe nanowires. The ZT increases as a function of temperature and it reaches ~ 0.035 for NW-1 and Figure 5 Temperature dependent figure of merit ZT of PbSnTe 0.018 for NW-2 at 300 K [Figure 5]. The NWs. The ZT of bulk single crystal sample calculated based on S, σ, κ at ~300 K in reference [9] (Orihashi et al.) is shown for temperature dependence of ZT is mainly attributed comparison. to the thermopower which increases dramatically with temperature [Figure 3 (b)]. The maximum ZT Comparing with the bulk single crystal samples of of PbSnTe NWs at 300 K is nearly an order of nearly the same chemical composition, PbSnTe magnitude higher than the bulk value [9], owing to nanowires show both enhanced thermopower and both the enhanced thermopower and the reduced reduced thermal conductivity, leading to a thermal conductivity. We also note that the n-type maximum improvement of ZT by a factor of nearly PbTe NWs grown by a similar vapor transport ~10. We suggest that the suppression of thermal approach (using different precursors) have a ZT of conductivity may be due to the increased ~0.0054 at 300 K [34]. Therefore, the ZTs of the phonon-surface scattering, while the enhanced PbSnTe NWs are several times higher than that of thermopower could be related to the unique the PbTe NWs [34]. This enhancement is attributed topological surface states. In comparison with the to a significantly higher electrical conductivity in PbTe nanowires grown using the same method, the the PbSnTe NWs. PbSnTe has a lower thermopower but a

significantly higher electrical conductivity. The ZTs 4 Conclusions of the PbSnTe NWs are also several times higher than that of the n-type PbTe NWs reported in the In summary, we performed a vapor transport literature [34]. Our work highlights growth and thermoelectric studies of high quality nanostructuring in combination with alloying as an single-crystalline PbTe and PbSnTe nanowires. The important approach to enhancing thermoelectric NWs were grown along the <001> direction with properties. We note that the PbSnTe bulk crystals dominant {100} facets. The chemical composition show high ZTs at high temperatures [9], so future ratio of Pb: (Pb+Sn) in the PbSnTe NWs has a measurements on nanowires at those temperatures strong dependence on the substrate position in the will be important to test if the thermoelectric growth reactor. Correlated measurements of S, σ, enhancement observed here persists well above and κ were performed on the same room temperature. single-nanowire devices to accurately determine the figure of merit ZT.

Acknowledgements

We thank Dr. Julio Martinez, John Nogan, Anthony R. James, Douglas V. Pete, Denise B. Webb and

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research 10Nano Res.

Renjie Chen for experimental assistances. S.X.Z., 473, 66-69. H.H., and N.S. acknowledge support from the [8] Zhang, Q.; Cao, F.; Liu, W. S.; Lukas, K.; Yu, B.; Chen, Laboratory Directed Research & Development S.; Opeil, C.; Broido, D.; Chen, G.; Ren, Z. F. Heavy program at Los Alamos National Laboratory. This Doping and Band Engineering by Potassium to Improve work was performed, in part, at the Center for the Thermoelectric Figure of Merit in p-Type PbTe, PbSe, and PbTe1-ySey. J. Am. Chem. Soc. 2012, 134, Integrated Nanotechnologies, an Office of Science 10031-10038. User Facility operated for the U.S. Department of Energy (DOE) Office of Science by Los Alamos [9] Orihashi, M.; Noda, Y.; Chen, L. D.; Goto, T.; Hirai, T. National Laboratory (Contract Effect of content on thermoelectric properties of p-type lead . J Phys Chem Solids 2000, 61, DE-AC52-06NA25396) and Sandia National 919-923. Laboratories (Contract DE-AC04-94AL85000). We also thank the Indiana University Nanoscale [10] Han, M.-K.; Zhou, X.; Uher, C.; Kim, S.-J.; Kanatzidis, Characterization Facility for access to the M. G. Increase in the Figure of Merit by Cd-Substitution in Sn1-xPbxTe and Effect of Pb/Sn Ratio on instrumentation. Thermoelectric Properties. Advanced Energy Materials 2012, 2, 1218-1225. Electronic Supplementary Material: [11] Minnich, A. J.; Dresselhaus, M. S.; Ren, Z. F.; Chen, G. Supplementary material (SEM images of PbTe Bulk nanostructured thermoelectric materials: current nanowires & microcrystals and PbSnTe research and future prospects. Energ Environ Sci 2009, microcrystals; two probe I-V characteristics of PbTe 2, 466-479. and PbSnTe nanowires; and detailed information [12] Vineis, C. J.; Shakouri, A.; Majumdar, A.; Kanatzidis, about thermopower measurement) is available in M. G. Nanostructured Thermoelectrics: Big Efficiency the online version of this article at Gains from Small Features. Adv Mater 2010, 22, 3970-3980. http://dx.doi.org/10.1007/s12274-***-****-*

[13] Hicks, L. D.; Dresselhaus, M. S. Effect of Quantum-Well Structures on the Thermoelectric Figure of Merit. References Physical Review B 1993, 47, 12727-12731.

[1] Bell, L. E. Cooling, heating, generating power, and [14] Hicks, L. D.; Dresselhaus, M. S. Thermoelectric figure recovering waste heat with thermoelectric systems. of merit of a one-dimensional conductor. Physical Science 2008, 321, 1457-1461. Review B 1993, 47, 16631-16634.

[2] Zebarjadi, M.; Esfarjani, K.; Dresselhaus, M. S.; Ren, Z. [15] Harman, T. C.; Taylor, P. J.; Walsh, M. P.; LaForge, B. E. F.; Chen, G. Perspectives on thermoelectrics: from Quantum dot superlattice thermoelectric materials and fundamentals to device applications. Energ Environ Sci devices. Science 2002, 297, 2229-2232. 2012, 5, 5147-5162. [16] Biswas, K.; He, J. Q.; Blum, I. D.; Wu, C. I.; Hogan, T. [3] LaLonde, A. D.; Pei, Y. Z.; Wang, H.; Snyder, G. J. Lead P.; Seidman, D. N.; Dravid, V. P.; Kanatzidis, M. G. telluride alloy thermoelectrics. Mater Today 2011, 14, High-performance bulk thermoelectrics with all-scale 526-532. hierarchical architectures. Nature 2012, 489, 414-418.

[4] Pei, Y. Z.; LaLonde, A.; Iwanaga, S.; Snyder, G. J. High [17] Weathers, A.; Shi, L. Thermal Transport Measurement thermoelectric figure of merit in heavy hole dominated Techniques for Nanowires and Nanotubes. Annual PbTe. Energ Environ Sci 2011, 4, 2085-2089. Review of Heat Transfer, edited by G. Chen 2013, 16, 101-134. [5] LaLonde, A. D.; Pei, Y. Z.; Snyder, G. J. Reevaluation of PbTe1-xIx as high performance n-type thermoelectric [18] Szczech, J. R.; Higgins, J. M.; Jin, S. Enhancement of material. Energ Environ Sci 2011, 4, 2090-2096. the thermoelectric properties in nanoscale and nanostructured materials. J Mater Chem 2011, 21, [6] Heremans, J. P.; Jovovic, V.; Toberer, E. S.; Saramat, A.; 4037-4055. Kurosaki, K.; Charoenphakdee, A.; Yamanaka, S.; Snyder, G. J. Enhancement of thermoelectric efficiency [19] Li, D. Y.; Wu, Y. Y.; Kim, P.; Shi, L.; Yang, P. D.; in PbTe by distortion of the electronic density of states. Majumdar, A. Thermal conductivity of individual silicon Science 2008, 321, 554-557. nanowires. Appl Phys Lett 2003, 83, 2934-2936.

[7] Pei, Y. Z.; Shi, X. Y.; LaLonde, A.; Wang, H.; Chen, L. [20] Hochbaum, A. I.; Chen, R. K.; Delgado, R. D.; Liang, W. D.; Snyder, G. J. Convergence of electronic bands for J.; Garnett, E. C.; Najarian, M.; Majumdar, A.; Yang, P. high performance bulk thermoelectrics. Nature 2011,

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research Nano Res. 11

D. Enhanced thermoelectric performance of rough [33] Zhang, T.; Wu, S. L.; Zheng, R. T.; Cheng, G. A. silicon nanowires. Nature 2008, 451, 163-167. Significant reduction of thermal conductivity in silicon nanowire arrays. Nanotechnology 2013, 24, 505718. [21] Boukai, A. I.; Bunimovich, Y.; Tahir-Kheli, J.; Yu, J. K.; Goddard, W. A.; Heath, J. R. Silicon nanowires as [34] Lee, S. H.; Shim, W.; Jang, S. Y.; Roh, J. W.; Kim, P.; efficient thermoelectric materials. Nature 2008, 451, Park, J.; Lee, W. Thermoelectric properties of individual 168-171. single-crystalline PbTe nanowires grown by a vapor transport method. Nanotechnology 2011, 22, 295707. [22] Fardy, M.; Hochbaum, A. I.; Goldberger, J.; Zhang, M. M.; Yang, P. D. Synthesis and thermoelectrical [35] Bui, C. T.; Xie, R. G.; Zheng, M. R.; Zhang, Q. X.; Sow, characterization of lead chalcogenide nanowires. Adv C. H.; Li, B. W.; Thong, J. T. L. Diameter-Dependent Mater 2007, 19, 3047. Thermal Transport in Individual ZnO Nanowires and its Correlation with Surface Coating and Defects. Small [23] Liang, W. J.; Rabin, O.; Hochbaum, A. I.; Fardy, M.; 2012, 8, 738-745. Zhang, M. J.; Yang, P. D. Thermoelectric Properties of p-Type PbSe Nanowires. Nano Research 2009, 2, [36] Lee, E. K.; Yin, L.; Lee, Y.; Lee, J. W.; Lee, S. J.; Lee, J.; 394-399. Cha, S. N.; Whang, D.; Hwang, G. S.; Hippalgaonkar, K.; Majumdar, A.; Yu, C.; Choi, B. L.; Kim, J. M.; Kim, [24] Kang, J.; Roh, J. W.; Shim, W.; Ham, J.; Noh, J. S.; Lee, K. Large Thermoelectric Figure-of-Merits from SiGe W. Reduction of Lattice Thermal Conductivity in Single Nanowires by Simultaneously Measuring Electrical and Bi-Te Core/Shell Nanowires with Rough Interface. Adv Thermal Transport Properties. Nano Letters 2012, 12, Mater 2011, 23, 3414–3419. 2918-2923.

[25] Soini, M.; Zardo, I.; Uccelli, E.; Funk, S.; Koblmuller, [37] Martinez, J. A.; Provencio, P. P.; Picraux, S. T.; Sullivan, G.; Morral, A. F. I.; Abstreiter, G. Thermal conductivity J. P.; Swartzentruber, B. S. Enhanced thermoelectric of GaAs nanowires studied by micro-Raman figure of merit in SiGe alloy nanowires by boundary and spectroscopy combined with laser heating. Appl Phys hole-phonon scattering. Journal of Applied Physics Lett 2010, 97, 263107. 2011, 110, 074317.

[26] Zuev, Y. M.; Lee, J. S.; Galloy, C.; Park, H.; Kim, P. [38] Finefrock, S. W.; Zhang, G. Q.; Bahk, J. H.; Fang, H. Y.; Diameter Dependence of the Transport Properties of Yang, H. R.; Shakouri, A.; Wu, Y. Structure and Antimony Telluride Nanowires. Nano Letters 2010, 10, Thermoelectric Properties of Spark Plasma Sintered 3037-3040. Ultrathin PbTe Nanowires. Nano Letters 2014, 14, 3466-3473. [27] Roh, J. W.; Jang, S. Y.; Kang, J.; Lee, S.; Noh, J. S.; Kim, W.; Park, J.; Lee, W. Size-dependent thermal [39] Jang, S. Y.; Kim, H. S.; Park, J.; Jung, M.; Kim, J.; Lee, conductivity of individual single-crystalline PbTe S. H.; Roh, J. W.; Lee, W. Transport properties of nanowires. Appl Phys Lett 2010, 96, 103101. single-crystalline n-type semiconducting PbTe nanowires. Nanotechnology 2009, 20, 415204. [28] Moore, A. L.; Pettes, M. T.; Zhou, F.; Shi, L. Thermal conductivity suppression in bismuth nanowires. Journal [40] Jung, H.; Park, D. Y.; Xiao, F.; Lee, K. H.; Choa, Y. H.; of Applied Physics 2009, 106, 034310. Yoo, B.; Myung, N. V. Electrodeposited Single Crystalline PbTe Nanowires and Their Transport [29] Martin, P. N.; Aksamija, Z.; Pop, E.; Ravaioli, U. Properties. J Phys Chem C 2011, 115, 2993-2998. Reduced Thermal Conductivity in Nanoengineered Rough Ge and GaAs Nanowires. Nano Letters 2010, 10, [41] Tai, G. A.; Guo, W. L.; Zhang, Z. H. Hydrothermal 1120-1124. synthesis and thermoelectric transport properties of uniform single-crystalline pearl-necklace-shaped PbTe [30] Li, G. D.; Liang, D.; Qiu, R. L. J.; Gao, X. P. A. Thermal nanowires. Cryst Growth Des 2008, 8, 2906-2911. conductivity measurement of individual Bi2Se3 nano-ribbon by self-heating three-omega method. Appl [42] Yan, Q. Y.; Chen, H.; Zhou, W. W.; Hng, H. H.; Boey, F. Phys Lett 2013, 102, 043104. Y. C.; Ma, J. A Simple Chemical Approach for PbTe Nanowires with Enhanced Thermoelectric Properties. [31] Davami, K.; Weathers, A.; Kheirabi, N.; Mortazavi, B.; Chem Mater 2008, 20, 6298-6300. Pettes, M. T.; Shi, L.; Lee, J. S.; Meyyappan, M. Thermal conductivity of ZnTe nanowires. Journal of [43] Yang, Y. A.; Taggart, D. K.; Cheng, M. H.; Hemminger, Applied Physics 2013, 114, 134314. J. C.; Penner, R. M. High-Throughput Measurement of the and the Electrical Conductivity [32] Guan, Z.; Gutu, T.; Yang, J. K.; Yang, Y.; Zinn, A. A.; Li, of Lithographically Patterned Polycrystalline PbTe D. Y.; Xu, T. T. Boron carbide nanowires: low Nanowires. J Phys Chem Lett 2010, 1, 3004-3011. temperature synthesis and structural and thermal conductivity characterization. J Mater Chem 2012, 22, [44] Wrasse, E. O.; Torres, A.; Baierle, R. J.; Fazzio, A.; 9853-9860. Schmidt, T. M. Size-effect induced high thermoelectric figure of merit in PbSe and PbTe nanowires. Physical

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research 12Nano Res.

Chemistry Chemical Physics 2014, 16, 8114-8118. Cha, J. J. Synthesis of SnTe Nanoplates with {100} and {111} Surfaces. Nano Letters 2014, 14, 4183-4188. [45] Dedi; Lee, P. C.; Chien, C. H.; Dong, G. P.; Huang, W. C.; Chen, C. L.; Tseng, C. M.; Harutyunyan, S. R.; Lee, [57] Zou, Y.-C.; Chen, Z.-G.; Lin, J.; Zhou, X.; Lu, W.; C. H.; Chen, Y. Y. Stress-induced growth of Drennan, J.; Zou, a. J. Controlling morphologies of SnTe single-crystalline lead telluride nanowires and their nanostructures by tuning catalyst composition. Nano thermoelectric transport properties. Appl Phys Lett Research 2015, In press, DOI 2013, 103, 023115. 10.1007/s12274-015-0806-y.

[46] Tai, G.; Zhou, B.; Guo, W. L. Structural characterization [58] Zhao, S. L.; Wang, H. A.; Zhou, Y.; Liao, L.; Jiang, Y.; and thermoelectric transport properties of uniform Yang, X.; Chen, G. C.; Lin, M.; Wang, Y.; Peng, H. L.; single-crystalline lead telluride nanowires. J Phys Chem Liu, Z. F. Controlled synthesis of single-crystal SnSe C 2008, 112, 11314-11318. nanoplates. Nano Research 2015, 8, 288-295.

[47] Hsieh, T. H.; Lin, H.; Liu, J. W.; Duan, W. H.; Bansil, A.; [59] Deringer, V. L.; Dronskowski, R. Stabilities and Fu, L. Topological crystalline insulators in the SnTe Reconstructions of PbTe Crystal Surfaces from material class. Nat. Commun. 2012, 3, 982. Density-Functional Theory. J Phys Chem C 2013, 117, 24455-24461. [48] Liu, J. W.; Duan, W. H.; Fu, L. Two types of surface states in topological crystalline insulators. Physical [60] Bierman, M. J.; Lau, Y. K. A.; Kvit, A. V.; Schmitt, A. Review B 2013, 88, 241303(R). L.; Jin, S. Dislocation-driven nanowire growth and Eshelby twist. Science 2008, 320, 1060-1063. [49] Xu, S. Y.; Liu, C.; Alidoust, N.; Neupane, M.; Qian, D.; Belopolski, I.; Denlinger, J. D.; Wang, Y. J.; Lin, H.; [61] Lau, Y. K. A.; Chernak, D. J.; Bierman, M. J.; Jin, S. Wray, L. A.; Landolt, G.; Slomski, B.; Dil, J. H.; Formation of PbS Nanowire Pine Trees Driven by Screw Marcinkova, A.; Morosan, E.; Gibson, Q.; Sankar, R.; Dislocations. J. Am. Chem. Soc. 2009, 131, Chou, F. C.; Cava, R. J.; Bansil, A.; Hasan, M. Z. 16461-16471. Observation of a topological crystalline insulator phase and topological phase transition in Pb1-xSnxTe. Nat. [62] Allen, J. E.; Hemesath, E. R.; Perea, D. E.; Lensch-Falk, Commun. 2012, 3, 1192. J. L.; Li, Z. Y.; Yin, F.; Gass, M. H.; Wang, P.; Bleloch, A. L.; Palmer, R. E.; Lauhon, L. J. High-resolution [50] Xu, Y.; Gan, Z. X.; Zhang, S. C. Enhanced detection of Au catalyst atoms in Si nanowires. Nat Thermoelectric Performance and Anomalous Seebeck Nanotechnol 2008, 3, 168-173. Effects in Topological Insulators. Physical Review Letters 2014, 112, 226801. [63] Short, N. R. A Redetermination of Lattice Parameters of Pbxsn1-Xte Alloys. J Phys D Appl Phys 1968, 1, 129. [51] Shi, H. L.; Parker, D.; Du, M. H.; Singh, D. J. Connecting Thermoelectric Performance and [64] Heremans, J. P.; Thrush, C. M.; Morelli, D. T. Topological-Insulator Behavior: Bi2Te3 and Bi2Te2Se Thermopower enhancement in PbTe with pb from First Principles. Phys Rev Appl 2015, 3, 014004. precipitates. Journal of Applied Physics 2005, 98, 063703. [52] Safdar, M.; Wang, Q. S.; Mirza, M.; Wang, Z. X.; Xu, K.; He, J. Topological Surface Transport Properties of [65] Xu, E. Z.; Li, Z.; Martinez, J. A.; Sinitsyn, N.; Htoon, H.; Single-Crystalline SnTe Nanowire. Nano Letters 2013, Li, N.; Swartzentruber, B.; Hollingsworth, J. A.; Wang, 13, 5344-5349. J.; Zhang, S. X. Diameter dependent thermoelectric properties of individual SnTe nanowires. Nanoscale [53] Li, Z.; Shao, S.; Li, N.; McCall, K.; Wang, J.; Zhang, S. 2015, 7, 2869-2876. X. Single Crystalline Nanostructures of Topological Crystalline Insulator SnTe with Distinct Facets and [66] Volklein, F.; Reith, H.; Cornelius, T. W.; Rauber, M.; Morphologies. Nano Letters 2013, 13, 5443-5448. Neumann, R. The experimental investigation of thermal conductivity and the Wiedemann-Franz law for single [54] Saghir, M.; Lees, M. R.; York, S. J.; Balakrishnan, G. metallic nanowires. Nanotechnology 2009, 20, 8. Synthesis and Characterization of Nanomaterials of the Topological Crystalline Insulator SnTe. Cryst Growth [67] Grasser, T.; Tang, T. W.; Kosina, H.; Selberherr, S. A Des 2014, 14, 2009-2013. review of hydrodynamic and energy-transport models for semiconductor device simulation. Proceeding of the [55] Safdar, M.; Wang, Q. S.; Mirza, M.; Wang, Z. X.; He, J. IEEE 2003, 91, 251-274. Crystal Shape Engineering of Topological Crystalline Insulator SnTe Microcrystals and Nanowires with Huge [68] Brucker, H.; Jantsch, W. Modulation of Fundamental Thermal Activation Energy Gap. Cryst Growth Des Absorption by Free Carrier Heating in Degenerate 2014, 14, 2502-2509. N-Pbte. Solid State Commun 1981, 37, 83-85.

[56] Shen, J.; Jung, Y.; Disa, A. S.; Walker, F. J.; Ahn, C. H.;

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

13

Silver Nanowires with Semiconducting Ligands for Low Temperature Transparent Conductors

Brion Bob,1 Ariella Machness,1 Tze-Bin Song,1 Huanping Zhou,1 Choong-Heui Chung,2 and Yang Yang1,* 1 Department of Materials Science and Engineering and California NanoSystems Institute,

University of California Los Angeles, Los Angeles, CA 90025 (USA)

2 Department of Materials Science and Engineering, Hanbat National University, Daejeon

305-719, Korea

Abstract Metal nanowire networks represent a promising candidate for the rapid fabrication of transparent electrodes with high transmission and low sheet resistance values at very low deposition temperatures. A commonly encountered obstacle in the formation of conductive nanowire electrodes is establishing high quality electronic contact between nanowires in order to facilitate long range current transport through the network. A new system of nanowire ligand removal and replacement with a semiconducting sol-gel tin oxide matrix has enabled the fabrication of high performance transparent electrodes at dramatically reduced temperatures with minimal need for post-deposition treatments of any kind.

Keywords: Silver Nanowires, Sol-Gel, Transparent Electrodes, Nanocomposites

14

1. Introduction. Silver nanowires (AgNWs) are long, thin, and possess conductivity values on the same order of magnitude as bulk silver (Ag) [1]. Networks of overlapping nanowires allow light to easily pass through the many gaps and spaces between nanowires, while transporting current through the metallic conduction pathways offered by the wires themselves. The high aspect ratios achievable for solution-grown AgNWs has allowed for the fabrication of transparent conductors with very promising sheet resistance and transmission values, often approaching or even surpassing the performance of vacuum-processed materials such as indium tin oxide (ITO) [2-6]. Significant electrical resistance within the metallic nanowire network is encountered only when current is required to pass between nanowires, often forcing it to pass through layers of stabilizing ligands and insulating materials that are typically used to assist with the synthesis and suspension of the nanowires [7, 8]. The resistance introduced by the insulating junctions between nanowires can be reduced through various physical and chemical means, including burning off ligands and partially melting the wires via thermal annealing [9, 10], depositing additional materials on top of the nanowire network [11-14], applying mechanical forces to enhance network morphology [15-17], or using various other post-treatments to improve the contact between adjacent wires [18-21]. Any attempt to remove insulating materials the network must be weighed against the risk of damaging the wires or blocking transmitted light, and so many such treatments must be reined in from their full effectiveness to avoid endangering the performance of the completed electrode. We report here a process for forming inks with dramatically enhanced electrical contact between AgNWs through the use of a semiconducting ligand system consisting of tin oxide (SnO2) nanoparticles. The polyvinylpyrrolidone (PVP) ligands introduced during AgNW synthesis in order to encourage one-dimensional growth are stripped from the wire surface using ammonium ions, and are replaced with substantially more conductive SnO2, which then fills the space between wires and enhances the contact geometry in the vicinity of wire/wire junctions. The resulting transparent electrodes are highly conductive immediately upon drying, and can be effectively processed in air at virtually any temperature below 300 °C. The capacity for producing high performance transparent electrodes at room temperature may be useful in the fabrication of devices that are damaged upon significant heating or upon the application of harsh chemical or mechanical post-treatments.

2. Results and Discussion

2.1. Ink Formulation and Characterization

Dispersed AgNWs synthesized using copper chloride seeds represent a particularly challenging material system for promoting wire/wire junction formation, and often require thermal annealing at temperatures near or above 200 °C to induce long range electrical conductivity within the deposited network [22, 23]. The difficulties that these wires present regarding junction formation is potentially due to their relatively large diameters compared to nanowires synthesized using other seeding materials, which has the capacity to enhance the thermal stability of individual wires according to the Gibbs-Thomson effect. We have chosen these wires as a demonstration of pre-deposition semiconducting ligand substitution in order to best illustrate the contrast between treated and untreated wires.

Completed nanocomposite inks are formed by mixing AgNWs with SnO2 nanoparticles in the presence of a compound capable of stripping the ligands from the AgNW surface. In this work, we have found that ammonia or ammonium salts act as effective stripping agents that are able to remove the PVP layer from the AgNW surface and allow for a new stabilizing matrix to take its place. Figure 1 shows a schematic of the process, starting from the precursors used in nanowire and nanoparticle synthesis and ending with the deposition of a completed film. The SnO2 nanoparticle solution naturally contains enough ammonium ions from its own synthesis to effectively peel the insulating ligands from the AgNWs and allow the nanoparticles to replace them as a stabilizing agent. If not enough SnO2 nanoparticles are used in the mixture, then the wires will rapidly agglomerate and settle to the bottom as large clusters. Large amounts of SnO2 in the mixture gradually begin to increase the sheet resistance of the nanowire network upon deposition, but greatly enhance the uniformity, durability, and wetting properties of the resulting films. We have found that AgNW:SnO2 weight ratios ranging between 2:1 and 1:1 produce well dispersed inks that are still highly conductive when deposited as films.

The nanowires were synthesized using a polyol method that has been adapted from the recipe described by Lee et al. [22, 23] Silver nitrate dissolved in ethylene glycol via ultrasonication was used as a precursor in the presence of copper chloride and PVP to provide seeds and produce anisotropic morphologies in the reaction products. Synthetic details can be found in the experimental section. Distinct from previous recipes, we have found that repeating the synthesis two times without cooling down the reaction mixture generally produces significantly longer nanowires than a single reaction step. The lengths of nanowires produced using this method fall over a wide range from 15 to 65 microns, with diameters between 125 and 250 nm. This range of diameters is common for wires grown using copper chloride seeds, although the double reaction produces a number of wires with roughly twice their usual diameter. The morphology of the as-deposited AgNWs as determined via SEM is shown in Figure 2(a), higher magnification images are also provided in Figures 2(c) and 2(d).

15

The SnO2 nanoparticles were synthesized using a sol-gel method typical for multivalent metal oxide gelation reactions. A large excess of deionized water was added to SnCl4·5H2O dissolved in ethylene glycol along with tetramethylammonium chloride and ammonium acetate to act as surfactants. The reaction was then allowed to progress for at least one hour at near reflux conditions, after which the resulting nanoparticle dispersion can be collected, washed, and dispersed in a polar solvent of choice. The material properties of SnO2 nanoparticles formed using a similar synthesis method have been reported previously [24], although the present recipe uses excess water to ensure that the hydrolysis reaction proceeds nearly to completion.

After mixing with SnO2 nanoparticles, films deposited from AgNW/SnO2 composite inks show a largely continuous nanoparticle layer on the substrate surface with some nanowires partially buried and some sitting more or less on top of the film. Representative scanning electron microscopy (SEM) images of nanocomposite films are shown in Figure 2(b). Regardless of their position relative to the SnO2 film, all nanowires show a distinct shell on their outer surface that gives them a soft and slightly rough appearance, as is visible in the higher magnification images shown in Figure 2(e) and 2(f). The SnO2 nanoparticles do a particularly good job coating the regions near and around junctions between wires, and frequently appear in the SEM images as bulges wrapped around the wire/wire contact points.

The precise morphology of the SnO2 shell that effectively surrounded each AgNW was analyzed in more detail using transmission electron microscopy (TEM) imaging. Figures 3(a) to 3(c) show individual nanowires in the presence of different ligand systems: as-synthesized PVP in Figure 3(a), inactive SnO2 in Figure 3(b), and SnO2 activated with trace amounts of ammonium ions in Figure 3(c). The as-synthesized nanowires show sharp edges, and few surface features. In the presence of inactive SnO2, which is formed by repeatedly washing the SnO2 nanoparticles in ethanol until all traces of ammonium ions are removed, the nanowires coexist with somewhat randomly distributed nanoparticles that deposit all over the surface of the TEM grid. When AgNWs are mixed with activated SnO2, a thick and continuous SnO2 shell is formed along the nanowire surface. In when sufficiently dilute SnO2 solutions are used to form the nanocomposite ink, nearly all of the nanoparticles are consumed during shell formation and effectively no nanoparticles are left to randomly populate the rest of the image.

As the AgNWs acquire their metal oxide coatings in solution, the properties of the mixture change dramatically. Freshly synthesized AgNWs coated with residual PVP ligands slowly settle to the bottom of their vial or flask over a time period of several hours to one day, forming a dense layer at the bottom. The AgNWs with SnO2 shells do not settle to the bottom, but remain partially suspended even after many weeks at concentrations that are dependent on the amount of SnO2 present in the solution.

A comparison of the settling behavior of various AgNW and SnO2 mixtures after 24 hours is shown in Figures 3(d) and 3(e). The ratios 8:4, 8:16, and 8:8 indicate the concentrations of AgNWs and SnO2 (in mg/mL) present in each solution. The 8:8 uncoupled solution, in which the PVP is not removed from the AgNW surface with ammonia, produces a situation in which the nanowires and nanoparticles do not interact with one another, and instead the nanowires settle as in the isolated nanowire solution while the nanoparticles remain well-dispersed as in the solution of pure SnO2. The mixtures of nanowires and nanoparticles in which trace amounts of ammonia are present do not settle to the bottom, but instead concentrate themselves until repulsion between the semiconducting SnO2 clusters is able to prevent further settling.

Our current explanation for the settling behavior of the wire/particle mixtures is that the PVP coating on the surface of the as-synthesized wires is sufficient to prevent interaction with the nanoparticle solution. The addition of ammonia into the solution quickly strips off the PVP surface coating and allowing the nanoparticles to coordinate directly with the nanowire surface. This explanation is in agreement with the effects of ammonia has on a solution of pure AgNWs, which rapidly begin to agglomerate into clusters and sink to the bottom as soon as any significant quantity of ammonia is added to the ink. We attribute the stripping ability of ammonia in these mixtures to the strong dative interactions that occur via the lone pair on the nitrogen atom interacting with the partially filled d-orbitals of the Ag atoms on the nanowire surface. These interactions are evidently strong enough to displace the existing coordination of the five-membered rings and carbonyl groups contained in the original PVP ligands and allow the ammonia to attach directly to the nanowire surface. Since ammonia is one of the original surfactants used to stabilize the surface of the SnO2 nanoparticles, we consider it reasonable that ammonia coordination on the nanowire surface would provide an appropriate environment for the nanoparticles to adhere to the AgNWs.

16

Scanning Energy Dispersive X-ray (EDX) Spectroscopy was also conducted on nanoparticle-coated AgNWs in order to image the presence of Sn and Ag in the nanowire and shell layer. The line scan results are shown in Figure 3(f), having been normalized to better compare the widths of the two signals. The visible broadening of the Sn lineshape compared to that of Ag is indicative of a Sn layer along the outside of the wire. The increasing strength of the Sn signal toward the center of the AgNW is likely due to the enhanced interaction between the TEM’s electron beam and the dense AgNW, which then improves the signal originating from the SnO2 shell as well. It is also possible that there is some intermixing between the Ag and Sn x-ray signals, but we consider this to be less likely as the distance between their characteristic peaks should be larger than the detection system’s energy resolution.

2.2. Network Deposition and Device Applications

For the deposition of transparent conducting films, a weight ratio of 2:1 of AgNWs to SnO2 nanoparticles was chosen in order to obtain a balance between the dispersibility of the nanowires, the uniformity of coated films, and the sheet resistance of the resulting conductive networks. Nanocomposite films were deposited on glass by blade coating from an ethanolic solution using a scotch tape spacer, with deposited networks then being allowed to dry naturally in air over several minutes.

The as-dried nanocomposite films are highly conductive, and require only minimal thermal treatment to dry and harden the film. Without the use of activated SnO2 ligands, deposited nanowire networks are highly insulating, and become conductive only after annealing at above 200 °C. The sheet resistance values of representative films are shown in Figure 4(a). The capability to form transparent conductive networks in a single deposition step that remain useful over a wide range of processing temperatures provides a high degree of versatility for designing thin film device fabrication procedures.

Figure 5(a) shows the sheet resistance and transmission of a number of nanocomposite films deposited from inks containing different nanowire concentrations. The deposited films show excellent conductivity at transmission values up to 85%, and then rapidly increase in sheet resistance as the network begins to reach its connectivity limit. The optimum performance of these networks at low to moderate transmission values is a consequence of the relatively large nanowire diameters, which scatter a noticeable amount of light even when the conditions required for current percolation are just barely met. Nonetheless, the sheet resistance and transmission of the completed nanocomposite networks place them within an acceptable range for applications in a variety of optoelectronic devices. Figure 5(b) shows the wavelength dependent transmission spectra of several nanowire networks, which transmit light well out into the region. The presence of high transmission values out to wavelengths well above 1300 nm, where ITO or other conductive oxide layers would typically begin to show parasitic absorption, is due to the use of semiconducting SnO2 ligands, which is complimentary to the broad spectrum transmission of the silver nanowire network itself.

Avoiding the use of highly doped nanoparticles has the potential to provide optical advantages, but can create difficulties when attempting to make electrical contact to neighboring device layers. In order to investigate their functionality in thin film devices, we have incorporated AgNW/SnO2 nanocomposite films as electrodes in amorphous silicon (a-Si) solar cells. Two contact structures were used during fabrication: one with the nanocomposite film directly in contact with the p-i-n absorber structure and one with a 10 nm Al:ZnO (AZO) layer present to assist in forming Ohmic contact with the device. The I-V characteristics of the resulting devices are shown in Figure 6(a).

The thin AZO contact layers typically show sheet resistance values greater than 2.5 kΩ/⧠, and so cannot be responsible for long range lateral current transport within the electrode structure. However, their presence is clearly beneficial in improving contact between the nanocomposite electrode and the absorber material, as the SnO2 matrix material is evidently not conductive enough to form a high quality contact with the p-type side of the a-Si stack. We hope that future modifications to the AgNW/SnO2 composite, or perhaps the use of islands of high conductivity material such as a discontinuous layer of doped nanoparticles will allow for the deposition of completed electrode stacks that provide both rapid fabrication and good performance.

Figure 6(b) contains the top view image of a completed device. The enhanced viscosity of the nanowire/sol-gel composite inks allows for films to be blade coated onto substrates with a variety of surface properties without reductions in network uniformity. In contrast with traditional back electrodes deposited in vacuum environments, the nanocomposite can be blade coated into place in a single pass under atmospheric conditions and dried within moments. We anticipate that the use of sol-gel mixtures to enhance wetting and dispersibility may prove useful in the formulation of other varieties of semiconducting and metallic inks for deposition onto a variety of substrate structures.

3. Conclusions

In summary, we have successfully exchanged the insulating ligands that normally surround as-synthesized AgNWs with shells of substantially more conductive SnO2 nanoparticles. The exchange of one set of ligands for the other is mediated by

17

the presence of ammonia during the mixing process, which appears to be necessary for the effective removal of the PVP ligands that initially cover the nanowire surface. The resulting nanowire/nanoparticle mixtures allow for the deposition of nanocomposite films that require no annealing or other post-treatments to function as high quality transparent conductors with transmission and sheet resistance values of 85% and 10 Ω/ , respectively. Networks formed in this manner can be deposited quickly and easily in open air, and have been demonstrated as an effective n-type electrode in a-Si solar cells when a thin interfacial layer is deposited first to ensure good electronic contact⧠ with the rest of the device. The ligand management strategy described here could potentially be useful in any number of material systems that presently suffer from highly insulating materials that reside on the surface of otherwise high performance nano and microstructures.

4. Experimental Details Tin oxide nanoparticle synthesis. Tin chloride pentahydrate was dissolved in ethylene glycol by stirring for several hours at a concentration of 10 grams per 80 mL to serve as a stock solution. In a typical synthesis reaction, 10 mL of the SnCl4·5H2O stock solution is added to a 100 mL flask and stirred at room temperature. Still at room temperature, 250 mg ammonium acetate and 500 mg ammonium acetate were added in powder form to regulate the solution pH and to serve as coordinating agents for the growing oxide nanoparticles. 30 ml of water was then added, and the flask was heated to 90 °C for 1 to 2 hours in an oil bath, during which the solution took on a cloudy white color. The gelled nanoparticles were then washed twice in ethanol in order to keep trace amounts of ammonia present in the solution. Additional washing cycles would deactivate the SnO2, and then require the addition of ammonia to coordinate with as-synthesized AgNWs.

Silver nanowire synthesis. Copper(ii) chloride dihydrate was first dissolved in ethylene glycol at

1 mg/ml to serve as a stock solution for nanowire seed formation. 20 ml of ethylene glycol was then added into a 100 ml flask, along with 200 µL of copper chloride solution. the mixture was then heated to 150 °C while stirring at 325 rpm, and .35g of PVP (MW 55,000) was added. In a small separate flask, .25 grams of silver nitrate was dissolved in 10 ml ethylene glycol by sonicating for approximately 2 minutes, similar to the method described here.22 The silver nitrate solution was then injected into the larger flask over approximately 15 minutes, and the reaction was allowed to progress for 2 hours. After the reaction had reached completion, the various steps were repeated without cooling down. 200 µL of copper chloride solution and .35g PVP were added in a similar manner to the first reaction cycle, and another .25g silver nitrate were dissolved via ultrasonics and injected over 15 minutes. The second reaction cycle was allowed to progress for another 2 hours, before the flask was cooled and the reaction products were collected and washed three times in ethanol.

Nanocomposite ink formation. After the synthesis of the two types of nanostructures is complete,

18

the double washed SnO2 nanoparticles and triple-washed nanowires can be combined at a variety of weight ratios to form the completed nanocomposite ink. The dispersibility of the mixture is improved when more

SnO2 is used, although the sheet resistance of the final networks will begin to increase if they contain excessive SnO2. AgNW agglomeration during mixing is most easily avoided if the SnO2 and AgNW solutions are first diluted to the range of 10 to 20 mg/ml in ethanol, with the SnO2 solution being added first to an empty vial and the AgNW solution added afterwards. The dilute mixture was then be allowed to settle overnight, and the excess solvent removed to concentrate the wires to a concentration that is appropriate for blade coating.

Film and electrode deposition. The completed nanocomposite ink was deposited onto any desired substrates using a razor blade and scotch tape spacer. The majority of the substrates used in this study were

Corning soda lime glass, but the combined inks also deposited well on silicon, SiO2, and any other substrates tested. Electrode deposition onto a-Si substrates was accomplished by masking off the desired cell area with tape, and then depositing over the entire region. The p-i-n a-Si stacks and 10 nm AZO contact layers were deposited using PECVD and sputtering, respectively.

ACKNOWLEDGMENTS The authors would like to acknowledge the use of the Electron Imaging Center for Nanomachines

(EICN) located in the California NanoSystems Institute at UCLA.

REFERENCES [1] Sun, Y.; Gates, B.; Mayers, B.; Xia, Y., Crystalline silver nanowires by soft solution

processing. Nano Lett. 2002, 2, 165-168.

[2] Kim, T.; Kim, Y. W.; Lee, H. S.; Kim, H.; Yang, W. S.; Suh, K. S., Uniformly

interconnected silver-nanowire networks for transparent film heaters. Adv. Funct.

Mater. 2013, 23, 1250-1255.

[3] Hu, L.; Wu, H.; Cui, Y., Metal nanogrids, nanowires, and nanofibers for transparent

electrodes. MRS Bull. 2011, 36, 760-765.

19

[4] van de Groep, J.; Spinelli, P.; Polman, A., Transparent conducting silver nanowire

networks. Nano Lett. 2012, 12, 3138-3144.

[5] Yang, L.; Zhang, T.; Zhou, H.; Price, S. C.; Wiley, B. J.; You, W., Solution-processed

flexible polymer solar cells with silver nanowire electrodes. ACS Appl. Mater.

Interfaces 2011, 3, 4075-4084.

[6] Scardaci, V.; Coull, R.; Lyons, P. E.; Rickard, D.; Coleman, J. N., Spray deposition of

highly transparent, low-resistance networks of silver nanowires over large areas. Small

2011, 7, 2621-2628.

[7] Wiley, B.; Sun, Y.; Xia, Y., Synthesis of silver nanostructures with controlled shapes

and properties. Acc. Chem. Res. 2007, 40, 1067-1076.

[8] Korte, K. E.; Skrabalak, S. E.; Xia, Y., Rapid synthesis of silver nanowires through a

cucl- or cucl2-mediated polyol process. J. Mater. Chem. 2008, 18, 437-441.

[9] Anuj, R. M.; Akshay, K.; Chongwu, Z., Large scale, highly conductive and patterned

transparent films of silver nanowires on arbitrary substrates and their application in

touch screens. Nanotechnology 2011, 22, 245201.

[10] Lee, J.-Y.; Connor, S. T.; Cui, Y.; Peumans, P., Solution-processed metal nanowire

mesh transparent electrodes. Nano Lett. 2008, 8, 689-692.

[11] Zhu, R.; Chung, C.-H.; Cha, K. C.; Yang, W.; Zheng, Y. B.; Zhou, H.; Song, T.-B.;

Chen, C.-C.; Weiss, P. S.; Li, G.; Yang, Y., Fused silver nanowires with metal oxide

nanoparticles and organic polymers for highly transparent conductors. ACS Nano 2011,

5, 9877-9882.

[12] Chung, C.-H.; Song, T.-B.; Bob, B.; Zhu, R.; Duan, H.-S.; Yang, Y., Silver nanowire

composite window layers for fully solution-deposited thin-film photovoltaic devices.

Adv. Mater. 2012, 24, 5499-5504.

20

[13] Kim, A.; Won, Y.; Woo, K.; Kim, C.-H.; Moon, J., Highly transparent low resistance

zno/ag nanowire/zno composite electrode for thin film solar cells. ACS Nano 2013, 7,

1081-1091.

[14] Ajuria, J.; Ugarte, I.; Cambarau, W.; Etxebarria, I.; Tena-Zaera, R. n.; Pacios, R.,

Insights on the working principles of flexible and efficient ito-free organic solar cells

based on solution processed ag nanowire electrodes. Sol. Energy Mater. Sol. Cells

2012, 102, 148-152.

[15] Tokuno, T.; Nogi, M.; Karakawa, M.; Jiu, J.; Nge, T.; Aso, Y.; Suganuma, K.,

Fabrication of silver nanowire transparent electrodes at room temperature. Nano Res.

2011, 4, 1215-1222.

[16] Lim, J.-W.; Cho, D.-Y.; Jihoon, K.; Na, S.-I.; Kim, H.-K., Simple brush-painting of

flexible and transparent ag nanowire network electrodes as an alternative ito anode for

cost-efficient flexible organic solar cells. Sol. Energy Mater. Sol. Cells 2012, 107,

348-354.

[17] De, S.; Higgins, T. M.; Lyons, P. E.; Doherty, E. M.; Nirmalraj, P. N.; Blau, W. J.;

Boland, J. J.; Coleman, J. N., Silver nanowire networks as flexible, transparent,

conducting films: Extremely high dc to optical conductivity ratios. ACS Nano 2009, 3,

1767-1774.

[18] Hu, L.; Kim, H. S.; Lee, J.-Y.; Peumans, P.; Cui, Y., Scalable coating and properties of

transparent, flexible, silver nanowire electrodes. ACS Nano 2010, 4, 2955-2963.

[19] Garnett, E. C.; Cai, W.; Cha, J. J.; Mahmood, F.; Connor, S. T.; Greyson Christoforo,

M.; Cui, Y.; McGehee, M. D.; Brongersma, M. L., Self-limited plasmonic welding of

silver nanowire junctions. Nat. Mater. 2012, 11, 241-249.

21

[20] Yu, Z.; Zhang, Q.; Li, L.; Chen, Q.; Niu, X.; Liu, J.; Pei, Q., Highly flexible silver

nanowire electrodes for shape-memory polymer light-emitting diodes. Adv. Mater.

2011, 23, 664-668.

[21] Song, T.-B.; Chen, Y.; Chung, C.-H.; Yang, Y.; Bob, B.; Duan, H.-S.; Li, G.; Tu,

K.-N.; Huang, Y., Nanoscale joule heating and electromigration enhanced ripening of

silver nanowire contacts. ACS Nano 2014, 8, 2804-2811.

[22] Lee, P.; Lee, J.; Lee, H.; Yeo, J.; Hong, S.; Nam, K. H.; Lee, D.; Lee, S. S.; Ko, S. H.,

Highly stretchable and highly conductive metal electrode by very long metal nanowire

percolation network. Adv. Mater. 2012, 24, 3326-3332.

[23] Lee, J. H.; Lee, P.; Lee, D.; Lee, S. S.; Ko, S. H., Large-scale synthesis and

characterization of very long silver nanowires via successive multistep growth. Cryst.

Growth Des. 2012, 12, 5598-5605.

[24] Bob, B.; Song, T.-B.; Chen, C.-C.; Xu, Z.; Yang, Y., Nanoscale dispersions of gelled

Sno2: Material properties and device applications. Chem. Mater. 2013, 25, 4725-4730.

22

Figure 1. Process flow diagram showing the synthesis of AgNWs and SnO2 nanoparticles followed by stirring in the presence of ammonium salts to create the final nanocomposite ink. Transparent conducting films were produced by blade coating the completed inks onto the desired substrate.

23

Figure 2. (a,c,d) SEM images of as-synthesized AgNWs at various magnifications. (b,e,f) SEM images of nanocomposite films, showing the tendency of the SnO2 nanoparticles to coat the entire outer surface of the AgNWs, increasing their apparent diameter and giving them a soft appearance.

24

Figure 3. Schematic diagrams and TEM images of (a) a single untreated AgNW, (b) an AgNW in the presence of uncoupled SnO2 (all ammonium ions removed), and (c) an AgNW with a coordinating

SnO2 shell. Scale bars in images (a), (b), and (c) are 300 nm, 400 nm, and 600 nm, respectively. (d,e)

Optical images of AgNW and SnO2 nanoparticle dispersions mixed in varying amounts (d) before and

(e) after settling for 24 hours. The numbers associated with each solution represent the AgNW:SnO2 concentrations in mg/ml. The uncoupled solution contains AgNWs and non-coordinating SnO2 nanoparticles, and shows settling behavior similar to the pure AgNW and pure SnO2 solutions. (f)

Normalized Ag and Sn EDX signal mapped across the diameter of a single nanowire, with the inset showing the scanning path across an isolated wire.

25

Figure 4. Sheet resistance versus temperature for films deposited using (red) AgNWs that have been washed three times in ethanol and (blue) mixtures of AgNW and SnO2 with weight ratio of 2:1. The annealing time at each temperature value was approximately 10 minutes. The large sheet resistance values of the bare AgNWs when annealed below 200 °C is typical for nanowires fabricated using copper chloride seeds, which clearly illustrate the impact of SnO2 coordination at low treatment temperatures.

26

Figure 5. (a) Sheet resistance and transmission data for samples deposited from solutions of varying nanostructure concentration. Each of these samples were fabricated starting from the same nanocomposite ink, which was then diluted to a range of concentrations while maintaining the same

AgNW to SnO2 weight ratio. (b) Transmission spectra of several transparent conducting networks chosen from the plot in plot (a).

27

Figure 6. (a) I-V characteristics of devices made with AgNW/SnO2 rear electrodes with (blue) and without (red) a 10 nm AZO contact layer. The dramatic double diode effect is likely a result of a significant barrier to charge injection at the electrode/a-Si interface. (b) Top view SEM image of the

AgNW/SnO2 composite films on top of the textured a-Si absorber. (c) Schematic cross section of the a-Si device architecture used in solar cell fabrication. The thickness of the thin AZO contact layer is exaggerated for clarity.

28