Opto-Electronic Advances, Volume. 6, Issue 6, 220170(2023)

Hot electron electrochemistry at silver activated by femtosecond laser pulses

Oskar Armbruster1,2、*, Hannes Pöhl2, and Wolfgang Kautek2、*
Author Affiliations
  • 1University of Natural Resources and Life Sciences, Vienna, Department of Bionanosciences, Muthgasse 11, 1190 Vienna, Austria
  • 2University of Vienna, Department of Physical Chemistry, Währinger Straße 42, 1090 Vienna, Austria
  • show less

    A silver microelectrode with a diameter of 30 µm in an aqueous K2SO4 electrolyte was irradiated with 55 fs and 213 fs laser pulses. This caused the emission of electrons which transiently charged the electrochemical double layer. The two applied pulse durations were significantly shorter than the electron-phonon relaxation time. The laser pulse durations had negligible impact on the emitted charge, which is incompatible with multiphoton emission. On the other hand, the observed dependence of emitted charge on laser fluence and electrode potential supports the thermionic emission mechanism.

    Keywords

    Introduction

    The photoelectric effect describes the ejection of an electron from a metal by a high-energy photon (photon energy > work function of the metal) 1, 2. High intensity irradiation with low-energy photons can also result in electron emission by the multiphoton photoelectric effect3-6.

    Additionally, low-energy photons can generate hot electrons which are in thermal nonequilibrium with the host lattice7-9. Electron-electron scattering leads to equilibration on a sub-picosecond time scale, which allows the application of the two-temperature model10. After the electron-phonon relaxation time, which is on the order of several picoseconds, electrons return to thermal equilibrium with the lattice by electron-phonon-scattering. Due to the comparatively low heat capacity of the electron gas, electron temperatures of the order of 103 K to 104 K are easily achieved while maintaining a cold lattice. These hot electrons can be thermally emitted into the environment as a current burst with a duration on the order of the electron-phonon relaxation time11-13.

    For a Gaussian beam, the absorbed volumetric energy density is

    U(r,z)=2βEw2πμexp(2(rw)2)exp(zμ),  

    where E is the pulse energy, β is the fraction of absorbed radiation, w is the Gaussian beam radius, and μ = 142 nm is the effective absorption length of silver8. For low electron temperatures up to some 5000 K, the electron heat capacity of silver Ce is approximately CeγeTe14 where γe = 62.8 J m−3 K−2 is the heat capacitance coefficient of silver14. Via Ce = dU/dTe, the electron temperature Te thus is

    Te(r,z)=2U(r,z)γe.  

    Figure 1 shows the electron temperature calculated according to Eq. (2).

    Calculated electron temperature Te of silver for various incident laser fluences F according to Equation (2) with γe = 62.8 J m−3 K−2, μ = 142 nm, and β = 0.15.

    Figure 1.Calculated electron temperature Te of silver for various incident laser fluences F according to Equation (2) with γe = 62.8 J m−3 K−2, μ = 142 nm, and β = 0.15.

    A metal under irradiation in contact with an electrolyte emits electrons which undergo thermalization and solvation within a time span of 100 fs to 1 ps corresponding to the fast Debye relaxation time in water15. In an electron acceptor-free solution, solvated electrons may then diffuse back to the electrode and are reinjected. If, however, a sufficient number of electron acceptors (such as H3O+, N2O, or NO3) are present, the electrons can be scavenged16.

    The work function from a metallic electrode into an electrolyte solution ϕ is the difference between the conduction band edge of the solution and the Fermi energy of the metal. ϕ depends on the electrode potential φ versus a reference potential and on the work function at that reference potential ϕ0 so that

    ϕ=ϕ0+qeφ,  

    with the elementary charge qe.

    Thermionic electron emission from a heated electron gas in nonequilibrium with the host lattice has been calculated using a one-dimensional interface potential17. The emitted charge density q is mainly determined by the electron gas temperature Te at the surface and the electrode potential contained in the work function17. At Te < Tc (~ 1500 K11) tunneling is predominant, whereas at Te > Tc over the barrier emission prevails17. When Te > Tc, the charge emitted from the electron gas into the electrolyte is

    qA0Te2kBTeϕτeexp(ϕkBTe),  

    with the Richardson constant A0 ≈ 120 A cm−2 K−1, the Boltzmann constant kB, the duration of the current pulse which is equal to the electron-phonon relaxation time τe, and the work function in an electrolyte ϕ (Eq. (3)). Equation (4) describes the exponential dependence of q on the pulse energy E (Eqs. (1) and (2)) and the electrode potential φ (Eq. (3)). In the tunneling case, Te < Tc, the emitted charge is negligible and cannot be measured experimentally17.

    Due to the Gaussian laser beam shape, the electron temperature Te and hence the emitted charge density q is radially nonconstant. Hence, the total emitted charge Q is

    Q=2π0reqrdr,  

    where re is the electrode radius.

    At relatively high electrolyte concentrations, the electrochemical diffuse double-layer can be neglected. The latter represents the space charge in the electrolyte18.

    There has been vivid interest in laser electrochemical phenomena and applications. Laser-induced electrochemical deposition of metals on metals relies possibly on thermal and defect generation effects19. When semiconductor substrates are chosen, locally photogenerated electrons can reduce metal ions resulting in metallic surface structures20-22. Laser-induced electrochemical de- and repassivation investigations are of interest in in-situ corrosion research studies23-25. The thermoemission of nonequilibrium electrons can be a unique technique for the generation of picosecond current pulses which cannot be realized by conventional instrumentation17. This allows short-lived intermediates of electrode reactions to be studied.

    Experimental

    A block diagram of the laser-electrochemical system is depicted in Fig. 2. The electrochemical cell consists of a silver working microelectrode (μWE, re = 15 μm radius) and a macroscopic platinum wire (1 mm diameter, 15 mm length) as counter electrode (CE). The fabrication of the extremely small μWE with a diameter of only 30 µm required an elaborate procedure described in the Supplementary information. A saturated Hg/Hg2SO4 reference electrode (not shown in Fig. 2) was employed to monitor the steady-state electrochemical potential during laser irradiation and the impedance spectroscopy (see below) experiments. Argon purged 0.2 M aqueous K2SO4 was used as the electrolyte with a pH adjusted to 2.2 with concentrated H2SO4.

    Via a large-valued resistor (R: 1 MΩ), the μWE was kept at the desired steady-state potential φ, controlled by a digital-to-analog converter (DAQ; DAQPad-6020E DAC, National Instruments). In order to facilitate measuring the small hot electron emission-induced signal Δφ(t) (order of millivolts) on the dc-bias background φ, a high pass amplifier (HPA) was employed. It provided an input impedance of 1012 Ω, dc voltage suppression, an amplification of ~102, and a bandwidth of 1.5 MHz. The hot electron-induced transient potential change output Δφ(t) of the HPA was recorded by a digital storage oscilloscope (OSC; WaveRunner 64Xi, LeCroy).

    Block diagram of the laser-electrochemical system. μWE: silver working microelectrode; CE: platinum counter electrode; fs-CPO: femtosecond chirped pulse oscillator; E: laser pulse energy; τ: pulse duration; OBJ: microscope objective; R: 1 MΩ resistor; DAQ: digital-to-analog converter output; φ: applied electrode potential; HPA: high pass amplifier; Δφ: transient voltage signal; OSC: digital storage oscilloscope.

    Figure 2.Block diagram of the laser-electrochemical system. μWE: silver working microelectrode; CE: platinum counter electrode; fs-CPO: femtosecond chirped pulse oscillator; E: laser pulse energy; τ: pulse duration; OBJ: microscope objective; R: 1 MΩ resistor; DAQ: digital-to-analog converter output; φ: applied electrode potential; HPA: high pass amplifier; Δφ: transient voltage signal; OSC: digital storage oscilloscope.

    A Ti:Sapphire Chirped Pulse Oscillator (fs-CPO; modified Femtosource XL, Femtolasers Produktions GmbH) delivered ultrashort pulses at a central wavelength of 800 nm 1.55 eV with a repetition rate of 11 MHz and pulse durations of either 55 fs or 213 fs. Pulses were picked with a Pockels cell at a rate of 50 Hz. The laser pulse energy E was adjusted using a half-wave plate/polarizer combination and delivered through a microscope objective (OBJ; EC Plan-NEOFLUAR 10×/0.3, Zeiss). The laser beam exiting the microscope objective was characterized with a custom knife edge beam profiler26, yielding a beam waist radius of w0 = (1.6 ± 0.1) µm and a Rayleigh length of zR = (11.7 ± 0.4) µm. Using the method reported in ref.27, the µWE was placed in front of the beam waist to achieve a beam radius of w = 10 µm.

    The applied laser peak fluences F0 were varied between 30 mJ cm−2 and 70 mJ cm−2, well below (factor of ~6) the multi-pulse damage threshold of silver immersed in the electrolyte, as determined by the A−ln(E) approach28-30. φ was varied between −500 mV and +500 mV vs. the standard hydrogen electrode (SHE), which covers a potential region well within the polarizability range of the used silver microelectrode31 as confirmed by repeated cyclic voltammetry sweeps.

    Results and discussion

    Owing to the comparatively large time constant of the electrochemical cell, the measurable quantity is the integral of the emission current, i.e., the emitted charge (Eq. (4)) which is observable as the transient voltage change Δφ(t) of the µWE. Figure 3 shows exemplary Δφ(t) transients triggered by a fs laser pulse irradiating the µWE. The extracted peak voltage change of Δφ0 = (19.7 ± 0.1) mV is highlighted in the inset of Fig. 3(a).

    (a) Exemplary Δφ(t) transient voltage measurement averaged over 100 single laser pulses. τ = (55 ± 1) fs, F0 = (64 ± 7) mJ cm-2, ϕ = (–460 ± 40) mV vs. SHE. The inset enlarges the region around t = 0 and shows the extracted peak voltage change Δφ0 = (19.7 ± 0.1) mV. (b) Comparison of Δφ(t) transient voltage measurements recorded at various laser peak fluences F0.

    Figure 3.(a) Exemplary Δφ(t) transient voltage measurement averaged over 100 single laser pulses. τ = (55 ± 1) fs, F0 = (64 ± 7) mJ cm-2, ϕ = (–460 ± 40) mV vs. SHE. The inset enlarges the region around t = 0 and shows the extracted peak voltage change Δφ0 = (19.7 ± 0.1) mV. (b) Comparison of Δφ(t) transient voltage measurements recorded at various laser peak fluences F0.

    The electrons emitted from the electrode into the electrolyte get solvated and quickly reduce the solvent, i.e., reduce the solvated H+ (hydronium ions). This leads to a recharging of the electrochemical double layer, characterized by its capacitance Cdl. That causes a transient change in the potential of the working electrode with amplitude Δφ0. Thus, the emitted charge is Q = Δφ0Cdl.

    Impedance spectroscopy was employed for the determination of the double layer capacitanceCdl(f, φ) as a function of frequency f and the dc bias potential φ. Figure 4 shows an exemplary impedance trace measured in situ prior to recording the Δφ(t) transient. Due to the large difference in surface area between the μWE and the CE (~104), the electrolyte and counter electrode contributions to the overall impedance can be neglected. To account for slight deviations from the ideal capacitive 1/f behavior (which is commonly observed in electrode impedance measurements), a constant phase element model (CPE)32, 33 was applied to determine Cdl. In this specific example, a frequency dependent double layer capacitance of Cdl(f) = [(0.40 ± 0.02) (f/1 Hz)(−0.064 ± 0.003)] nF was derived. For a frequency of 2.5 kHz, which approximates the inverse transient decay time of Δφ(t), this translates to a capacitance of (0.21 ± 0.01) nF. This particular example (τ = (55 ± 1) fs, F0 = (64 ± 7) mJ cm−2, φ = (−460 ± 40) mV vs. SHE) thus yields Q = (4.2 ± 0.2) pC.

    Imaginary part X of an exemplary impedance measurement corresponding to the Δφ(t) transient in Fig. 3.The solid line shows the constant phase element fit.

    Figure 4.Imaginary part X of an exemplary impedance measurement corresponding to the Δφ(t) transient in Fig. 3.The solid line shows the constant phase element fit.

    Figure 5 shows the logarithmic plot of the emitted hot electron charge Q over the applied electrode bias φ and the peak laser fluence F0 for two pulse durations, τ = (55 ± 1) fs and τ = (213 ± 1) fs. Despite the significant difference in terms of pulse duration (factor of ~4), emitted charges are largely unaffected. This also means that multiphoton excitation can be ruled out. The emitted charges range from 1 fC to 5 pC. This corresponds to 5×103 to 3×107 electrons emitted from the metal into the electrolyte solution. Thus, the external quantum efficiency is very low and ranges from 2×10−8 to 6×10−5. Considering the electron-phonon relaxation time (700 fs34 for Ag) as the duration of the current pulse and the Gaussian beam area w2π as the emission surface, current densities exceeded 2×106 A cm−211, 12, 35.

    Logarithmic plot of the emitted hot electron charge Q for two pulse durations, τ = (55 ± 1) fs (red circles) and τ = (213 ± 1) fs (blue stars), plotted over the applied electrode bias φ and peak laser fluence F0. The gray surface is calculated from Eq. (5) but for a constant factor with γe = 62.8 J m−3 K−2, μ = 142 nm, β = 0.15, and φ0 = 3.41 eV at 0 V vs. SHE.

    Figure 5.Logarithmic plot of the emitted hot electron charge Q for two pulse durations, τ = (55 ± 1) fs (red circles) and τ = (213 ± 1) fs (blue stars), plotted over the applied electrode bias φ and peak laser fluence F0. The gray surface is calculated from Eq. (5) but for a constant factor with γe = 62.8 J m−3 K−2, μ = 142 nm, β = 0.15, and φ0 = 3.41 eV at 0 V vs. SHE.

    A comparison with previous work shows that the maximum charge density below the damage threshold of Ag is comparable with the present results, i.e., ca. 1 µC cm−211, 12, 35. However, the method of the Gaussian beam radius determination strongly deviated which affects the reported intensity scale. The previous studies relied on the simple evaluation of the diameter of the blackened area on a photographic paper. This is known to overestimate the actual Gaussian beam radius. In the present work, it was determined by the accurate knife edge method26, 36.

    The strongly superlinear, almost exponential, dependence of the emitted chargeQ, and thus current density, on the laser fluence F0 is in accordance with the thermoemission model expressed by Eq. (5). The two employed pulse durations are significantly shorter than the electron-phonon relaxation time (~ 700 fs34). With a weak temperature dependence of reflectivity and energy deposition depth, the local peak electron temperature Te will depend only on the local fluence F(r) according to Eq. (2). The gray surface in Fig. 5 is calculated according to Eq. (5) using the following parameters. For silver, the work function in aqueous solution at the point of zero charge is 2.77 eV37. The point of zero charge is −640 mV vs. SHE38. With this, the work function is φ0 = 3.41 eV at 0 V vs. SHE. The effective absorption length was reported as μ = 142 nm due to ballistic electron transport8. The electron heat capacity coefficient was reported to be γe = 62.8 J m−3 K−214. For a moderately polished surface, a fraction of absorbed radiation of β = 0.15 can be assumed. On the other hand, the laser pulse durations had negligible impact on the emitted charge, which is incompatible with multiphoton emission.

    The photonic excitation of the metallic Fermi sea leads to a broadened Fermi-Dirac distribution n(ε) in the metal electrode (Fig. 6). The scale of ε is opposite to the electrode potential scale φ. The electron transfer at the interface to the electrolyte takes place according to the Franck Condon principle at constant energy39, 40. The Fermi level εF in the absence of irradiation, which is electrochemically controlled by the applied potential φ, is positioned such that electron transfer to the electron scavenger in the electrolyte is thermodynamically negligible. The Fermi sea, which is heated upon laser irradiation, can inject electrons at a higher energy level than the equilibrium state into the fluctuating states of the hydronium ions. This corresponds to a more negative electrode potential φ. The initial reduction product of this process is hydrogen in statu nascendi, atomic H0. According to the Franck Condon principle mentioned above, this extremely strong reducing agent can re-inject an electron into the metal. This became possible because the Fermi distribution returned to its equilibrium state after the electron-phonon relaxation.

    Schematic electronic Fermi-Dirac distribution n on the metal and density of states (DOS) as function of the electronic energy ε of H3O+ and H0.

    Figure 6.Schematic electronic Fermi-Dirac distribution n on the metal and density of states (DOS) as function of the electronic energy ε of H3O+ and H0.

    Conclusions

    Hot electrons in thermal nonequilibrium with the metal lattice were generated with low-energy photons from a femtosecond pulse laser. Two pulse durations, τ = (55 ± 1) fs and τ = (213 ± 1) fs significantly shorter than the electron-phonon relaxation time were employed with little effect on the emitted charge. A multiphoton emission scheme should depend strongly on intensity and thus pulse duration. This is however not observed, discouraging the interpretation of multiphoton photoemission. Electrons were emitted into an adjacent acidic K2SO4 electrolyte as giant current pulses with durations of the order of the electron-phonon relaxation time. Current densities reached 2×106 A cm−2. Emitted electrons reacted with the solvent by reducing the protons to hydrogen. The transient electrode potential change based on the recharging of the electrochemical double layer was registered. The strongly superlinear dependence of the emitted charge on the laser fluence is supported by the thermoemission model. The photonic excitation of the metallic Fermi sea leads to a broadened Fermi-Dirac distribution in the metal electrode. The former injects electrons at a higher energy level, i.e., at a more negative electrode potential, than the equilibrium value into the fluctuating states of the hydronium ions. The reduction products can re-inject electrons into the metal because the Fermi distribution has returned to its equilibrium state.

    [1] H Hertz. Ueber einen einfluss des ultravioletten lichtes auf die electrische entladung. Ann Phys, 267, 983-1000(1887).

    [2] P Lenard. Ueber die lichtelektrische wirkung. Ann Phys, 313, 149-198(1902).

    [3] AT Georges. Theory of the multiphoton photoelectric effect: a stepwise excitation process. Phys Rev B, 51, 13735-13738(1995).

    [4] JP Girardeau-Montaut, C Girardeau-Montaut. Theory of ultrashort nonlinear multiphoton photoelectric emission from metals. Phys Rev B, 51, 13560-13567(1995).

    [5] A Damascelli, G Gabetta, A Lumachi, L Fini, F Parmigiani. Multiphoton electron emission from Cu and W: an angle-resolved study. Phys Rev B, 54, 6031-6034(1996).

    [6] AT Georges. High-order multiphoton photoelectric effect at midinfrared laser wavelengths. Phys Rev A, 66, 063412(2002).

    [7] M Bonn, DN Denzler, S Funk, M Wolf, SS Wellershoff et al. Ultrafast electron dynamics at metal surfaces: competition between electron-phonon coupling and hot-electron transport. Phys Rev B, 61, 1101-1105(2000).

    [8] J Hohlfeld, SS Wellershoff, J Güdde, U Conrad, V Jähnke et al. Electron and lattice dynamics following optical excitation of metals. Chem Phys, 251, 237-258(2000).

    [9] D Bäuerle. Laser Processing and Chemistry(2011).

    [10] SI Anisimov, BL Kapeliovich, TL Perel’man. Electron emission from metal surfaces exposed to ultrashort laser pulses. J Exp Theor Phys, 39, 375-377(1974).

    [11] AG Krivenko, W Kautek, J Krüger, VA Benderskii. Subpicosecond emission from mercury and silver into electrolyte solution: an experimental study. Russ J Electrochem, 33, 394-400(1997).

    [12] J Krüger, W Kautek, AG Krivenko, VA Benderskii. Gigantic hydrogen-ion discharge currents initiated by a subpicosecond laser. Russ J Electrochem, 34, 1068-1075(1998).

    [13] W Kautek, O Armbruster. Non-thermal material response to laser energy deposition. In Castillejo M, Ossi PM, Zhigilei L, eds. Lasers in Materials Science, 43-66(2014).

    [14] ZB Lin, LV Zhigilei, V Celli. Electron-phonon coupling and electron heat capacity of metals under conditions of strong electron-phonon nonequilibrium. Phys Rev B, 77, 075133(2008).

    [15] WJ Ellison, K Lamkaouchi, JM Moreau. Water: a dielectric reference. J Mol Liq, 68, 171-279(1996).

    [16] AM Brodsky, YV Pleskov. Electron photoemission at a metal-electrolyte solution interface. Prog Surf Sci, 2, 1-73(1972).

    [17] VA Benderskii, AV Benderskii. Laser Electrochemistry of Intermediates(1995).

    [18] JOM Bockris, AKN Reddy, ME Gamboa-Aldeco. Modern Electrochemistry 2A: Fundamentals of Electrodics 2nd ed(2000).

    [19] VA Benderskii, IO Efimov, AG Krivenko. Short-pulse laser activation of metal electrodes. J Electroanal Chem Interfacial Electrochem, 315, 29-64(1991).

    [20] S Tamir, J Zahavi. Laser‐induced gold deposition on a silicon substrate. J Vac Sci Technol A, 3, 2312-2315(1985).

    [21] W Kautek, N Sorg, W Paatsch. Laser-induced electrodeposition of transition metals on silicon. Electrochim Acta, 36, 1803-1810(1991).

    [22] N Sorg, W Kautek, W Paatsch. Etching pretreatment and galvanic Cu enhancement of laser-deposited ultrathin Ni structures on p-Si. Ber Bunsenges Phys Chem, 95, 1501-1507(1991).

    [23] R Oltra, GM Indrianjafy, M Keddam, H Takenouti. Laser depassivation of a channel flow double-electrode: a new technique in repassivation studies. Corros Sci, 35, 827-832(1993).

    [24] TO Nagy, U Pacher, A Giesriegl, L Soyka, G Trettenhahn et al. Laser-induced electrochemical de- and repassivation investigations on plasma-oxidized aluminium alloys. Appl Surf Sci, 302, 184-188(2014).

    [25] TO Nagy, MJJ Weimerskirch, U Pacher, W Kautek. Repassivation investigations on aluminium: physical chemistry of the passive state. Z Phys Chem, 230, 1303-1327(2016).

    [26] JM Khosrofian, BA Garetz. Measurement of a Gaussian laser beam diameter through the direct inversion of knife-edge data. Appl Opt, 22, 3406-3410(1983).

    [27] O Armbruster, A Naghilou, H Pöhl, W Kautek. In-situ and non-destructive focus determination device for high-precision laser applications. J Opt, 18, 095401(2016).

    [28] JM Liu. Simple technique for measurements of pulsed Gaussian-beam spot sizes. Opt Lett, 7, 196-198(1982).

    [29] H Howard, AJ Conneely, GM O'Connor, TJ Glynn. Investigation of a method for the determination of the focused spot size of industrial laser beams based on the drilling of holes in mylar film. Proc SPIE, 4876, 541-552(2003).

    [30] J Krüger, W Kautek. Ultrashort pulse laser interaction with dielectrics and polymers. In Lippert TK, ed. Polymers and Light, 247-290(2004).

    [31] M Pourbaix. Atlas of Electrochemical Equilibria in Aqueous Solutions 2nd ed(1974).

    [32] GJ Brug, ALG van den Eeden, M Sluyters-Rehbach, J Sluyters. The analysis of electrode impedances complicated by the presence of a constant phase element. J Electroanal Chem Interfacial Electrochem, 176, 275-295(1984).

    [33] T Pajkossy. Impedance spectroscopy at interfaces of metals and aqueous solutions — surface roughness, CPE and related issues. Solid State Ion, 176, 1997-2003(2005).

    [34] RHM Groeneveld, R Sprik, A Lagendijk. Femtosecond spectroscopy of electron-electron and electron-phonon energy relaxation in Ag and Au. Phys Rev B, 51, 11433-11445(1995).

    [35] AG Krivenko, J Krüger, W Kautek, VA Benderskii. Subpicosecond-pulse-laser-induced electron emission from mercury and silver into aqueous electrolytes. Ber Bunsenges Phys Chem, 99, 1489-1494(1995).

    [36] J Krüger, W Kautek. Ultrashort pulse laser interaction with dielectrics and polymers. In Lippert TK, ed. Polymers and Light, 247-290(2004).

    [37] YM Zolotovitskii, LI Korshunov, VA Benderskii. Electron work function from metals in a liquid dielectric. Bull Acad Sci USSR, Div Chem Sci, 21, 760-763(1972).

    [38] R Holze. Table 3.1. Electrode potentials of zero charge of metal electrodes in contact with electrolyte solutions. In Lechner MD, ed. Electrochemical Thermodynamics and Kinetics, 223-272(2007).

    [39] H Gerischer. Über den ablauf von redoxreaktionen an metallen und an halbleitern. I. Allgemeines zum elektronenübergang zwischen einem festkörper und einem redoxelektrolyten. Z Phys Chem, 26, 223-247(1960).

    [40] H Gerischer. Über den ablauf von redoxreaktionen an metallen und an halbleitern. II. Metall-elektroden. Z Phys Chem, 26, 325-338(1960).

    Tools

    Get Citation

    Copy Citation Text

    Oskar Armbruster, Hannes Pöhl, Wolfgang Kautek. Hot electron electrochemistry at silver activated by femtosecond laser pulses[J]. Opto-Electronic Advances, 2023, 6(6): 220170

    Download Citation

    EndNote(RIS)BibTexPlain Text
    Save article for my favorites
    Paper Information

    Category: Research Articles

    Received: Aug. 24, 2022

    Accepted: Jan. 16, 2023

    Published Online: Oct. 8, 2023

    The Author Email: Armbruster Oskar (;), Kautek Wolfgang (;)

    DOI:10.29026/oea.2023.220170

    Topics