Self Limiting PT EALD

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Self-terminating growth of Pt by electrochemical deposition Yihua Liu, Dincer Gokcen, Ugo Bertocci and Thomas. P.

Moffat* Material Measurement Laboratory National Institute of Standards and Technology Gaithersburg, Maryland 20899 Abstract A self-terminating rapid electrodeposition process for controlled growth of Pt monolayer films from a K2PtCl4-NaCl electrolyte has been developed that is tantamount to wet atomic layer deposition (ALD). Despite the deposition overpotential being in excess of 1 V, Pt deposition is quenched at potentials just negative of proton reduction by an alteration of the double layer structure induced by a saturated surface coverage of underpotential deposited hydrogen, (Hupd). The surface is reactivated for Pt deposition by stepping the potential to more positive values where Hupd is oxidized and fresh sites for adsorption of PtCl42- become available. Periodic pulsing of the potential enables sequential deposition of two dimensional (2-D) Pt layers to fabricate films of desired thickness relevant to a range of advanced technologies.

*thomas.moffat@nist.gov

One sentence summary: An unanticipated process for atomic layer deposition of Pt is detailed whereby potential control of adsorbed H enables sequential deposition of metal monolayers from aqueous solutions.

Pt is a key constituent in a wide range of heterogeneous catalysts, but its high cost constrains development of important alternative energy conversion systems such as low temperature fuel cells (1-3). A variety of strategies are being explored to enhance catalyst performance and minimize Pt loadings. These range from alloying to nanoscale engineering of core-shell and related architectures that typically involve spontaneous processes such as dealloying and segregation to form Pt-rich surface layers (4, 5). The deposition of 2-D Pt layers, that are also of interest in thin film electronics and magnetic materials, is non-trivial due to the step-edge barrier to interlayer transport that results in roughening or 3-D mound formation (6). In situ scanning tunnel microscopy (STM) of Pt electrodeposition at moderate overpotentials reveals that metal nucleation and growth on Au proceeds by formation of 3-D clusters at defect sites on single crystal surfaces (7). At small overpotentials, X-ray scattering indicates that smooth Pt monolayers can be electrodeposited on Au (111) although a long growth time of 2000 s is required (8). Voltammetric studies show a potential dependent transition between 2-D island versus 3D multilayer growth although it is only possible to obtain a partial Pt monolayer coverage in the 2-D growth regime (9). To circumvent these difficulties, surface limited place exchange reactions are being explored. For instance galvanic displacement of an underpotential deposited (upd) metal monolayer, typically Cu, occurs by the desired Pt group metal with the exchange resulting in a sub-monolayer coverage of the noble metal (10, 11). The process can be repeated to form multiple layers using a variant known as electrochemical atomic layer epitaxy (ECALE) (12). The multistep process typically requires an exchange of electrolytes and some care to control (or avoid) the trapping of the less noble metal as a minor alloying constituent within the film. The reversible nature

of many upd reactions makes it difficult to control deposition processes especially when considering sub-nanometer scale films. Robust additive fabrication schemes are facilitated by irreversible processes analogous to vapor phase deposition of thin films at low temperatures, although kinetic factors often constrain the quality of the resulting films. Prior analytical studies of Pt deposition have largely limited the applied potential to values positive of Hupd and proton reduction. One intriguing exception is Pt deposition from a pH 10, Pt(NH3)2(H2O)22+ - NaHPO4 electrolyte, where inhibition of the reaction was evident as the potential was scanned into the Hupd region, although the magnitude and thus significance of the effect was not examined (13). Herein, we show that formation of a saturated Hupd layer exerts a remarkable quenching or self terminating effect on Pt deposition, restricting it to a high coverage of 2-D Pt islands. When repeated, by using a pulsed potential waveform to periodically oxidize the Hupd layer, sequential deposition of discrete Pt layers can be achieved. The process is thus analogous to ALD but with a rapid potential cycle replacing the time consuming displacement and replacement of the ambient reactant. This report focuses on Pt deposition experiments performed at room temperature in aqueous solutions consisting of 0.5 mol/L NaCl and 0.003 mol/L K2PtCl4 with pH values ranging between 2.5 and 4. Beyond this particular electrolyte, self-terminating Pt deposition was observed over a wide range of pH and Cl- concentrations and was not dependent on the oxidation state (2+, 4+) of the Pt halide precursors. To isolate the partial current associated with only the growth process, an electrochemical quartz crystal microbalance (EQCM) was used to track Pt deposition on

an Au electrode as the potential was swept in the negative direction. Voltammetry in Fig. 1a shows the onset of Pt deposition at 0.25 VSSCE followed by a substantial current rise to a maximum at -0.32 VSSCE that is close to diffusion limited PtCl42- reduction. Beyond the peak, the deposition rate decreases smoothly as the mass transfer boundary layer thickness expands. A sharp drop in the current occurs as the potential moves negative of -0.5 VSSCE, eventually reaching a minimum near -0.7 VSSCE followed by an increase due to hydrogen evolution from water. The gravimetrically determined (EQCM) metal deposition rate reveals that the sharp drop below -0.5 VSSCE corresponds to complete quenching of metal deposition. This remarkable self-termination or passivation process occurs despite the large applied overpotential (> 1 V) available for driving the deposition reaction. The gravimetric data is used to reconstruct the partial voltammogram for Pt deposition a two electrons process. Good agreement with the measured voltammogram indicates the current efficiency of Pt deposition is close to 100 % as the potential is swept toward the diffusion limited value. As the current peak is approached, an apparent loss in efficiency is observed, due to non-uniform deposition that develops as the PtCl42depletion gradient sets up a convective flow field that spans the electrode. In contrast to the EQCM, voltammetry with a rotating disk electrode (RDE) provides uniform mass transport that yields a more symmetric peak (Fig. 1b). The contribution of the proton reduction reaction is isolated by performing voltammetry in the absence of the Pt complex. Merging of the respective voltammograms at negative potentials indicates that quenching of the metal deposition reaction is coincident with the onset of the H2 evolution reaction. The overlap of the diffusion limited proton reduction current also

indicates the absence of significant homogeneous reaction between the generated H2 and PtCl42-, excluding this as an explanation for the quenching of the Pt deposition reaction. The two electron reduction of PtCl42- to Pt is not expected to depend on pH, and the onset of significant Pt deposition from PtCl42- at 0.0 VSSCE shown in Fig. 1d supports this contention. In contrast, sharp acceleration of the deposition rate below -0.2 VSSCE is clearly pH dependent. This correlates with the onset of Hupd evident in PtCl42--free voltammetry (Fig. 1c). Chronocoulometry studies indicate that the transition between a halide and a hydrogen covered Pt surface occurs in the same regime (14). The metal deposition rate increases with Hupd coverage reaching a peak value that is independent of pH, while the peak potential shifts by -0.059 V/pH reflecting the importance of H surface chemistry in controlling the Pt deposition process. The onset of proton reduction in the absence of PtCl42-, marked by the dotted line in Fig. 1b, occurs at the essentially the same potential. Thus, the peak deposition rate occurs at the hydrogen reversible potential. Moving to more negative potentials, the metal deposition rate declines rapidly and within 0.1 V of its peak value the current merges with that attributable solely to diffusion limited proton reduction, indicating complete quenching of the Pt deposition reaction. Importantly, transient studies of Hads on Pt indicate that the coverage does not reach saturation at the reversible hydrogen potential but rather occurs 0.1 V below the reversible value (15). This is precisely the potential regime where the metal deposition reaction is fully quenched. Cyclic voltammetry reveals that the passivation process is reversible with reactivation coincident with the onset of Hupd oxidation (Fig. 1S). Selftermination of the metal deposition reaction arises from perturbation of the double layer structure that accompanies Hads saturation of the Pt surface. Recent theoretical work

indicates that the water structure adjacent to a hydrogen covered Pt(111) surface is significantly altered with the centroid of the O atoms within the first water layer being displaced by more than 0.1 nm from the metal surface as the water-water interactions in the first layer become stronger (16). In a related development, an EQCM study of Pt in sulfuric acid has identified a potential of minimal mass near the reversible potential of hydrogen reactions (17). The gravimetric measurements reflect the impact of Hupd on the adjacent water structure that leads to a minimum in coupling between the electrode and electrolyte, consistent with the recent theoretical result. In addition to Hupd perturbation of the water structure, the quenching of metal deposition reaction occurs at potentials negative of the Pt point of zero charge (pzc) where anions would have been desorbed (14). The above combination exerts a remarkable effect whereby PtCl42- reduction is completely quenched while diffusion limited proton reduction continues unabated. Self-terminating Pt deposition was also examined under potentiostatic conditions. Optical micrographs of a selection of films after 500 s deposition at various potentials are shown as inserts in Fig. 1a. Only the lower half of Au-coated Si(100) wafer was immersed in solution with differences in reflectivity and color indicate the anomalous dependence of deposition on potential; specifically a 33 nm thick Pt film was deposited at -0.4 VSSCE while a nearly invisible much thinner layer was grown at -0.8 VSSCE. X-ray photoelectron spectroscopy (XPS) was used to further quantify the composition and thickness of Pt grown as a function of deposition time and potential on (111) textured Au. For films deposited at -0.8 VSSCE, a representative spectrum with the 4f doublets for the metallic states of Au and Pt is shown in Fig. 2 (insert). The ratio of the Pt and Au peak areas was used to calculate the Pt thickness assuming it forms a uniform

overlayer (18). For deposition times up to 1000 s, the measured thickness varies between 0.21 nm and 0.25 nm, congruent with the deposition of a Pt monolayer with a thickness comparable to the (111) d-spacing of Pt. Monolayer formation is complete within the first second of stepping the potential to -0.8 VSSCE and the absence of further growth confirms the self-terminating nature of the deposition reaction. Beyond 1000 s, an additional increment of Pt deposition is evident. Inspection of the surface with scanning electron microscopy revealed a sparse coverage of spherically shaped Pt particles on the surface attributable to H2 induced precipitation, a process requiring some heterogeneity and extended incubation to nucleate. Particle formation can be avoided by using shorter deposition times. Scanning tunneling microscopy was used to directly observe the Pt overlayer morphology (Fig. 3). Analysis was facilitated by using a flame annealed Au (111) surface with isolated surface steps, 0.24+/-0.02 nm in height, that serve as fiduciary markers (Fig. 3a). Pt deposition results in three distinct levels of contrast that reflect the surface height with the lowest level being the original Au terraces (Fig. 3b). The same three-level structure is observed independently of deposition time up to 500 s (Fig. 3c). The middle contrast level corresponds to a high density of Pt islands that cover ~ 85 % of the Au surface with a step height of ~ 0.24 nm consistent with XPS results. Inspection using a higher rendering contrast reveals a ~10 % coverage of a second layer of small Pt islands with a step height ranging between 0.23 nm to 0.26 nm (Fig. 3d). Step positions

associated with the flame annealed substrate are preserved with negligible expansion or overgrowth of the 2-D Pt islands occurring beyond the original step edge. The lateral span of the Pt islands lies in the range of 2.02+/-0.38 nm corresponding to an area of

4.23+/-1.97 nm2. Incipient coalescence of the islands is constrained by surrounding (dark) narrow channels, 2.1+/-0.25 nm wide, that account for the remaining Pt-free portion of the first layer. The reentrant channels correspond to open Au terrace sites that are surrounded by adjacent Pt islands in what amounts to a huge increase in step density relative to the original substrate, the net geometric or electronic effect of which is to block further Pt deposition. The chemical nature of the inter-island region is assayed by exploiting the distinctive voltammetry of Pt and Au with respect to Hupd and oxide formation and reduction as detailed in Fig. 2S Similar three level Pt overlayers have been observed for monolayer films produced by molecular beam epitaxy (MBE) deposition at 0.05 ML/min (19). Pt-Au intermixing driven by the decrease in surface energy that accompanies Au surface segregation was evident. In the present work, Pt monolayer formation is effectively complete within 1 s giving a growth rate three orders of magnitude greater than the MBESTM study. Exchange of the deposited Pt with the underlying Au substrate is expected to be less developed although intermixing and possible chemical contrast is evident on limited sections of the surface that are correlated with the original faulted geometry of the partially reconstructed Au surface. Upon lifting of the reconstruction, the excess Au atoms expelled mark the original fault location as linear one dimensional surface defects in the Pt overlayer (Fig. 3e). Simplifying, a schematic of the self-terminating Pt deposition process in Fig. 3f indicates that the Hupd accompanying incremental expansion of the 2-D Pt islands serves to hinder the development of a second Pt layer, presumably by perturbation the overlying water structure.16 This rapid process results in a much

higher island coverage than has been obtained by other methods such as galvanic exchange reactions. As the saturated Hupd coverage is the agent of termination, reactivation for further Pt deposition is possible by removing the upd layer by sweeping or stepping the potential to positive values, e.g. > +0.2 VSSCE, where negligible Pt deposition occurs. Sequential pulsing between +0.4 VSSCE and -0.8 VSSCE enables Pt monolayer deposition to be controlled in a digital manner. EQCM was used to track the mass gain showing two net increments per cycle (Fig. 4a). The mass gain is attributed to a combination Pt deposition (486 ng/cm2 for a monolayer of Pt(111)), anion adsorption and desorption (41 ng/cm2 for 7 x1014 Cl- ion /cm2, 117 ng/cm2 for a 0.14 fractional coverage of PtCl42-) (7, 20) and coupling to other double layer components such as water. The anionic mass increments are expected to be asymmetric for the first cycle on the Au surface but once it is covered subsequent cycles only involve Pt surface chemistry. After correcting for the electroactive surface area of the Au electrode (Areal/Ageometric=1.2 derived from reductive desorption of Au oxide in perchloric acid) the net mass gain for each cycle indicates that close to a pseudomorphic layer of Pt is deposited. XPS analysis of Pt films grown for various deposition cycles gives remarkably good agreement with EQCM data (Fig. 4b). The ability to rapidly manipulate potential and double layer structure, as opposed to exchange of reactants, offers simplicity, substantially improved process efficiency, and far greater process speed than other surface limited deposition methods.

References 1. F. T. Wagner, R. Lakshmanan, M. F. Mathias, J. Phys. Chem. Lett. 1, 2204 (2010).

2. M. K. Debe, Nature 486, 43 (2012). 3. M. T. M. Koper, Ed., Fuel Cell Catalysis, A Surface Science Approach (Wiley & Sons, Hoboken, NJ, 2009). 4. V. R. Stamenkovic, et al. Nat. Mater. 6, 241 (2007). 5. R. R. Adzic, et al. Top. Catal. 46, 249 (2007). 6. T. Michely, J. Krug, Islands, Mounds and Atoms, Patterns and Processes in Crystal Growth Far from Equilibrium (Springer Series in Surface Science, V42, New York, 2003). 7. H. F. Waibel, M. Kleinert, L. A. Kibler, D. M. Kolb, Electrochim. Acta 47, 1461 (2002). 8. T. Kondo, et al. Electrochim. Acta 55, 8302 (2010). 9. I. Bakos, S. Szabo, T. Pajkossy, J. Solid State Electrochem. 15, 2453 (2011). 10. S. R. Brankovic, J. X. Wang, R. R. Adzic, Surf. Sci. 474, L173 (2001). 11. D. Gokcen, S.-E. Bae, S. R. Brankovic, Electrochim. Acta 56, 5545 (2011). 12. B. W. Gregory, J. L. Stickney, J. Electroanal. Chem. 300, 543 (1991). 13. A. J. Gregory, W. Levason, R. E. Noftle, R. Le Penven, D. Pletcher, J. Electroanal. Chem. 399, 105 (1995). 14. N. Garcia-Araez, V. Climent, E. Herrero, J. Feliu, J. Lipkowski, J. Electroanal. Chem. 582, 76 (2005). 15. D. Strmcnik, D. Tripkovic, D. van der Vliet, V. Stamenkovic, N. M. Markovic, Electrochem. Commun. 10, 1602 (2008). 16. T. Roman, A. Gro, Structure of water layers on hydrogen-covered Pt electrodes Cataly. Today, available on line July 12, in press

10

17. G. Jerkiewicz, G. Vatankhah, S. Tanaka, J. Lessard, Langmuir 27, 4220 (2011). 18. P. J. Cumpson, M. P. Seah, Surf. Interface Anal. 25, 430 (1997). 19. M. O. Pedersen, et al. Surf. Sci. 426, 395 (1999). 20. Y. Nagahara, et al. J. Phys. Chem B. 108, 3224 (2004).

Figure 1. Gravimetric and voltammetric measurements (2 mV/s) of Pt deposition from a NaCl-PtCl42- solution using either (a) a static EQCM or (b) (c) (d) an Au RDE (400 rpm). The inserts in (a) are optical images of Pt films grown on 1 cm wide Au-coated Si(100) wafers for 500 s at the indicated potentials. (b), (c), and (d) present the effect of pH on PtCl42- reduction and the background reactions associated with the supporting electrolyte. Background reactions were examined using a Pt RDE.

Figure 2. XPS derived thickness (red squares) of Pt films as a function of deposition time at -0.8 VSSCE on Au-coated Si wafers from a pH=4 solution. The Au and Pt lines correspond to the (111) d-spacing of the respective bulk metals.

Figure 3. (a) STM images of representative Au(111) surface with monoatomic steps. (bc) 2D Pt layers obtained after (b) 5 and (c) 500 second deposition at -0.8 VSSCE. (d) High contrast image of 2-D Pt layer morphology on Au(111). (e) Linear defects in Pt layer associated with lifting of the reconstructed Au substrate. (f) A schematic of Hupd terminated Pt deposition on Au(111).

11

Figure 4. Sequential deposition of Pt monoatomic layers by pulsed deposition in a pH 4 solution. (a) Mass change accompanying each pulse. (b) EQCM mass increase is converted to thickness and compared with XPS measurements. XPS analysis of the EQCM specimen ()and a series of Pt films deposited on Au-coated Si wafers.

12

Figure 1

13

Figure 2

14

Figure3

15

Figure 4

16

Supplementary Information Self-terminating growth of Pt by electrochemical deposition Yihua Liu, Dincer Gokcen, Ugo Bertocci and Thomas. P. Moffat Material Measurement Laboratory National Institute of Standards and Technology Gaithersburg, Maryland 20899

Material and methods Kinetic study of PtCl42- reduction Linear scan voltammetry (LSV) at a sweep rate of 2 mV/s was employed to study the reduction kinetics of PtCl42- from solutions consisting of 0.50 mol/L NaCl and 0.003 mol/L K2PtCl4. The influence of pH, adjusted by HClO4 and NaOH additions, was

examined in a closed electrochemical cell filled with a Ar-saturated 100 ml PtCl42solution. A Pt plate counter electrode was held in a glass tube filled with 0.50 mol/L NaClO4 connected to the PtCl42- electrolyte through a fine glass frit. The reference electrode was a sodium chloride saturated calomel electrode (SSCE). The working electrode was either a mechanically polished Au rotating disk electrode (RDE) spun at 400 rpm or a static Au-coated quartz crystal electrode. The background activity, i.e. proton reduction, on a Pt RDE was determined using in 0.5 mol/L NaCl solutions at various pH values. All solutions were prepared from analytical-grade chemicals dissolving in 18 Mcm water. The same grade of water was used for rinsing and cleaning activities. All glassware was cleaned by soaking for one hour in aqua regia followed by extensive rinsing.

Growth of Pt monolayer and multilayer The potential pulse method outlined in the main text was utilized to grow Pt monolayer and multilayer films from a pH 4 PtCl42- solution. Au-coated Si(100) wafer fragments were used for XPS studies while an annealed Au (111) single crystal surfaces was used for STM characterizations. The Au-coated Si (100) surfaces were prepared by e-beam evaporation of a 150 nm Au layer on a 5 nm Ti seeded polished Si (100) surface. Immediately before use the Au-coated Si (100) wafers were soaked for a minute in a piranha* solution made of 3/1 volume ratio of concentrated H2SO4 (70%) and H2O2 (30%). The Au (111) surfaces were prepared using Claviliers flame annealing technique the Au (111) surfaces were annealed using H2 flame for 10 minutes and then cooled to room temperature gradually (21). Followed the electrodeposition, the deposits were

promptly rinsed with 18 M water, dried in a stream of N2, and subjected to the characterizations immediately.
Warning: Piranha solution should be handled with caution: most probably when it has been mixed with significant quantities of oxidizable organic materials detonation may occur. Likewise, working solution should not be sealed from atmosphere due to gas evolution. Accordingly used solution should be properly disposed with appropriate care.

XPS The XPS (Kratos AXIS Ultra DLD) (22) was operated at a base pressure of 3 x 10-10 torr using a monochromated AlK source. The Casa XPS program was used in evaluating the peak areas of Au 4f and Pt 4f spectra using Shirleys algorithm for background subtraction. The peak area ratio was used to calculate the thickness of the overlayer, d, after accounting for the elemental sensitivity factors (si) (sAu = 6.25, sPt = 5.575) and the attenuation length of the photoelectrons in the Pt overlayer (AL =1.252 nm) (18).

I /s d = AL cos ln 1 + Pt Pt I Au / s Au

[1]

For reference purposes the 111 d-spacing for bulk Pt (ao=0.39240 nm) is 0.227 nm, which is slightly less than 0.235 nm of Au (ao=0.40786 nm).

STM STM tips were made of etching Pt/Ir (90:10) wires in 2:1 CaCl2:H2O solution at 25V-AC potential. Etched STM tips were rinsed with water and acetone. All highresolution STM measurements were performed using a Digital Instruments Nanoscope III controller with an A-type scanner at constant current mode (Itip<5nA). Plane-fit STM images were not subjected to any other filtering options. Images were analyzed using both Nanoscope and WSXM software (23). Average step height values, lateral dimensions (both x and y directions) and sizes of the Pt islands are computed using autonomous technique provided by the software for various images recorded at the different regions of the sample surface. Standard deviation values (+/-) are quoted with average values to reflect variances observed in the different images.

EQCM The EQCM experiment was performed using AT cut quartz crystals coated with a 150 nm Au layer and an adhesion layer of 5 nm Ti (Maxtek). The same cleaning procedure was followed as described for Au-coated Si(100) surfaces. The electrode was maintained under potential control for the duration of the experiment. Specifically, the experiments began with a 80 mL Ar-saturated 0.5 mol/L NaCl pH 4 solution with the potential set to 0.400 VSSCE. After stability was established a 1 mL aliquot of

concentrated K2PtCl4 solution was added to give a final PtCl42- concentration of 0.003 mol/L after being homogenized by magnetic stirring. Once the signal drift became insignificant, the potential pulses were applied.

Supplementary results

Reversibility of the self terminating deposition process

Figure 1S. Cyclic voltammetry reveals the reversible nature of suppressed and reactivated Pt deposition from a pH 3.5 solution of 0.5 mol/L NaCl + 0.003 mol/L K2PtCl4 (400 rpm, 2 mV/s).

Voltammetric examination of Pt overlayer on Au The chemical nature of the inter-island region is assayed by exploiting the distinctive voltammetry of Pt and Au with respect to Hupd and oxide formation and reduction. In 0.1 mol/L HClO4 Hupd features are evident at 0.050 VRHE E 0.400 VRHE. The wave shape is consistent with that for Pt(111) although the magnitude 108 C/cm2

+/- 5 is less than 146 C/cm2 due to finite size effects (Fig. 2S).24 This is also similar to the Hupd results observed for Pt rich Pt1-xAux surface alloys grown on Pt(111).25 Oxidation of the surface shows two distinct reduction waves for Pt oxide at 0.67 VRHE and Au oxide at 1.14 VRHE, with the former being more pronounced. The peak potential for the Au oxide reduction is shifted to more negative values compared to pure Au due to finite size effects. With due consideration of the background current for a fully consolidated Pt deposit, the charge associated with the Au oxide formation and reduction on the monolayer Pt film electrode corresponds to ~ 11 % of the Au substrate being accessible to the electrolyte. This is in reasonable agreement with the STM coverage determination.

Figure 2S. Cyclic voltammetry shows Hupd as well as oxide formation and reduction in 0.1 mol/L HClO4 on Au-coated Si surfaces before and after the growth of a Pt monolayer.

References 21. J. Clavilier, R. Faure, G. Guinet, R. Durand, J. Electroanal. Chem. 107, 205 (1980) 22. Identification of commercial products in this paper was done to specify the experimental procedure. In no case does this imply endorsement or recommendation by the National Institute of Standards and technology.

23. I. Horcas, et al. Rev. Sci. Instrum. 78, 013705 (2007) . 24. M.T.M. Koper, Electrochimica Acta, 56, 10645 (2011) 25. A. Bergbreiter, O. B. Alves and H.E. Hoster, Chem Phys Chem, 11, 1505 (2010).

You might also like