Plasmon-induced charge transfer has been studied for the development of plasmonic photodiodes and solar cells. There are two mechanisms by which a plasmonic nanoparticle can transfer charge to an adjacent material: indirect transfer following plasmon decay and direct transfer as a way of plasmon decay. Using single-particle dark-field scattering and photoluminescence imaging and spectroscopy of gold nanorods on various substrates, we identify linewidth broadening and photoluminescence quantum yield quenching as key spectroscopic signatures that are quantitatively related to plasmon-induced interfacial charge transfer. We find that dark-field scattering linewidth broadening is due to chemical interface damping through direct charge injection via plasmon decay. The photoluminescence quantum yield quenching reveals additional mechanistic insight into electron–hole recombination as well as plasmon generation and decay within the gold nanorods. Through these two spectroscopic signatures, we identify charge transfer mechanisms at TiO2 and indium doped tin oxide interfaces and uncover material parameters contributing to plasmon-induced charge transfer efficiency, such as barrier height and resonance energy.
INTRODUCTION
There is continuing intense interest in plasmonic nanoparticle–semiconductor hybrid structures for photovoltaic and photocatalytic devices, utilizing the large absorption cross section of plasmonic nanoparticles and their ability to generate tunable, energetic charge carriers to inject into the adjacent material.1–6 The predominately proposed mechanism of charge transfer is via hot carriers generated after plasmon decay, followed by their subsequent transfer to the adjacent material.7–9 However, theoretical work and some experimental evidence also suggest plasmon-induced direct charge transfer (PICT) as a possible mechanism, where electrons involved in the plasmon resonance decay via injection into the adjacent material.10–13 If PICT is the dominant pathway, it is expected to provide higher charge injection quantum efficiencies than the hot-carrier injection pathway as PICT avoids competing internal relaxation in the metal.10,11 Thus, quantifying the contributions of PICT and hot-carrier transfer to the overall yield is important to device design.
Many plasmon enhanced photocatalytic and photovoltaic studies rely on completed devices to measure their performance and are carried out on bulk samples.14–16 Recently, methods for single-particle characterization of plasmonic photovoltaic devices have been developed, utilizing either an electrochemical cell or electron microscopy.17,18 However, there is still a need for all-optical methods that are capable of predicting device performance at multiple design stages, which is minimally invasive and does not require a completed device.13
Here, we use dark-field scattering (DFS) and photoluminescence (PL) spectroscopies to identify the mechanisms and material parameters involved in charge transfer from gold nanorods (AuNRs) into the conduction band of an adjacent semiconductor: indium doped tin oxide (ITO) and amorphous titanium dioxide (TiO2) compared to quartz (SiO2). These substrates were chosen since ITO and TiO2 have different Schottky barrier heights, ITO and TiO2 are widely used and studied semiconductor materials, and SiO2 is an effective insulator that serves as a standard for no charge transfer. Additionally, our studies offer lower bounds on the spectral signatures of charge transfer efficiency, as the contact area between the AuNRs and the semiconductors is minimal in our sample geometry of AuNRs on a planar substrate compared to overcoated or core–shell AuNRs.
METHODS
Sample preparation
Three different AuNR sizes, nominally 23 × 64, 30 × 75, and 39 × 96 nm and referred to as small, medium, and large, respectively, were used for all measurements. The medium AuNRs were prepared by seed mediated growth according to Ref. 19 and the small and large AuNRs were prepared according to Ref. 20. The AuNR size distributions were published previously.21–23 The AuNRs were spin coated onto indexed SiO2, ITO coated glass, and TiO2 substrates. All substrates were cleaned prior to spin coating by O2 plasma etching. We made no attempts to remove the surfactants from the AuNRs as that could introduce differences between substrates during O2 plasma cleaning. While the surfactants are known to create a ∼3.2 nm dielectric shell around the AuNRs,24 previous studies have demonstrated that charge transfer is still observed in electrochemical experiments with the surfactant still present on the AuNR surface.22,25 The TiO2 coated substrates were made by atomic layer deposition (Ultratech Savannah ALD) of 12 nm TiO2 onto glass slides from the tetrakis(dimethylamido)titanium(IV) precursor for 300 cycles at a rate of 0.04 nm/cycle. Substrate indexing was performed by electron beam evaporation of 20 nm Ti through a transmission electron microscopy grid (Ted Pella) shadow mask, resulting in a labeled alphanumeric grid on the substrate.
Dark-field scattering and photoluminescence imaging and spectroscopy
Single-particle DFS spectroscopy was performed according to previously published studies using an inverted microscope (Zeiss, Axio Observer.D1m).21,26 Briefly, incandescent light (Zeiss, HAL 100) was focused by a dark-field oil immersion condenser through the substrate onto a sample region moved by a piezo-sample scanning stage (Physik Instrumente, P-517.3CL). The scattered light was collected with a 50× air space objective [Zeiss, EC Epiplan-Neofluar, 0.8 numerical aperture (NA)], spatially filtered through a 50 µm pinhole (Thorlabs), and either imaged on an avalanche photodiode (APD, PerkinElmer, SPCM-AQRH-15) for obtaining the AuNR position or dispersed through a spectrometer (Shamrock, SR193i-A) and detected with a charge-coupled device (CCD) camera (Andor, iDus 420 BEX2-DD) for acquiring spectra. The raw spectra were corrected for the illumination, background, and collection efficiency by subtracting the background from the raw spectrum and dividing the result by the dark count subtracted incandescent light spectrum.
Correlated single-particle PL imaging and spectroscopy were performed on the same AuNRs as investigated by DFS imaging and spectroscopy according to previously published procedures.21,26 Briefly, the AuNRs were illuminated in an epifluorescence configuration through the 50× 0.8 NA air space objective with a 488 nm continuous wave laser (Coherent Obis) having a power of 0.155 MW/cm2 at the sample plane. The emission was collected by the same objective, transmitted through a 50/50 beam splitter (Chroma), 488 nm notch filter (Semrock), and 496 nm long pass filter (Semrock). The pinhole used in DFS was removed for PL measurements. The PL was either imaged on the APD or dispersed through the spectrometer and detected with the CCD camera as in the case of the DFS measurements. The raw PL spectrum was corrected for instrument efficiency and background according to a previous method.27 Briefly, the background (a PL spectrum of an area with no AuNRs) was subtracted from the raw AuNR PL spectrum and corrected for the wavelength-dependent detection efficiency of the setup.
Quantum yield calculation
The quantum yields (QYs) of the individual AuNRs were calculated according to a previously published method.27 Here, the QY is defined as the ratio of emitted photons to absorbed photons. First, the raw photon counts were determined by integrating the photon counts for localized AuNRs within a 10 × 10 µm2 scanned APD image. The number of emitted photons was calculated by correcting the raw photon counts for the wavelength-dependent efficiencies of the optics and filters used and photon detection efficiency of the APD. The number of absorbed photons was calculated using the excitation power of the 488 nm laser and the absorption cross section of the individual AuNRs as obtained from finite difference time domain (FDTD) simulations. Details regarding the FDTD simulations are discussed in the section titled Electromagnetic simulations. The energy absorbed was converted into photons absorbed by dividing by the energy of the incident photons (2.54 eV). Finally, the spectrally resolved QY was calculated by multiplying the QY with the integral normalized PL spectrum. The Purcell factor scaled QY was calculated by integrating the QY spectrum at the resonance energy ±3 nm and dividing by the Purcell factor calculated according to the procedure given below. By dividing by the Purcell factor, we removed the resonance energy, linewidth, and mode volume dependencies from the PL QY.
Purcell factor
The Purcell factor was calculated according to a previous publication.21 Briefly, the Purcell factor can be derived experimentally using the following definition:28
where Eres is the scattering resonance energy, Γ is the linewidth, and V is the mode volume. The mode volume was calculated according to the following equation:28,29
The integral is calculated over all space, where |E| is the magnitude of the electric field determined from FDTD simulations discussed in the section titled Electromagnetic simulations, ω is the frequency, ϵ is the complex dielectric function of gold (taken from Johnson and Christy), and γ is the Drude damping, set to 0.07 eV.30,31
Electromagnetic simulations
Absorption cross section
The absorption cross sections for the QY calculations were determined using FDTD simulations (Lumerical FDTD Solutions) of AuNRs with sizes similar to those measured by scanning electron microscopy (SEM). A total of 302 unique AuNR sizes were simulated for each substrate with ranges of 16–46 nm for the width and 56–115 nm for the length. The AuNRs were modeled according to previously published methods as hemisphere capped cylinders using the gold dielectric function measured by Johnson and Christy31 sitting on a semi-infinite substrate of either SiO2 (refractive index of 1.52), ITO (refractive indices from Ref. 32), or TiO2 (refractive indices from Ref. 33).21 The choice of gold dielectric dataset is not expected to change any trends in the absorption cross section calculation since the 488 nm absorption is far from the AuNR resonances. The simulated absorption cross sections at 488 nm were interpolated with a 1 nm spacing according to a cubic spline. The average absorption cross section of a ±3 nm patch around each experimentally measured AuNR’s size was used as the AuNR’s absorption cross section for QY calculations. The average calculated absorption cross sections had a standard error of less than 10%.
Mode volume
The electric field components used in the mode volume calculations were determined using FDTD simulations. The electric field distributions at the longitudinal plasmon resonance were obtained using 3D frequency domain field profile monitors around the AuNR. The computational mesh size was set to 1 nm, rather than the default 0.3 nm, to decrease the computational cost of using a 3D frequency domain field profile monitor. As above, the size of each AuNR from SEM was used to find simulation results within a ±3 nm size window.
Scanning electron microscopy
Correlated SEM micrographs were obtained after performing all DFS and PL measurements to determine the sizes of individual AuNRs. The indexed grids on the samples enabled for single-particle DFS and PL measurements to be correlated with SEM images for the same individual AuNRs. The SEM micrographs were collected on a FEI Quanta 400 environmental scanning electron microscope with field emission gun operated in the low-vacuum mode at a voltage of 30 kV.
RESULTS
We studied the DFS and 488 nm, off-resonant, interband excited PL spectra of single AuNRs from three different size distributions spin coated onto SiO2, ITO, and TiO2 substrates (Fig. S1). We assume similar interfacial contacts between the AuNRs and the different substrates, as supported by the overlapping AuNR size distributions for all substrates (Fig. S1). As demonstrated previously, the PL and DFS resonance energies follow a linear relationship with a slight blueshift in the PL resonance compared to DFS [Fig. 1(a)].34,35 This correlation continues to hold for the AuNRs on ITO and TiO2, demonstrating that the substrate does not affect this resonance energy relationship between DFS and PL. The scattering linewidth, as extracted through Lorentzian fitting, depends on the resonance energy, size, and substrate, as illustrated in Fig. 1(b). The substrate effect on linewidth is twofold. First, the resonance energy redshifts for AuNRs of the same size but on ITO or TiO2 compared to SiO2 [Fig. 1(c)] because of the higher refractive indices of those substrates. This spectral shift results in a linewidth narrowing as the plasmon is moved away from the interband region, as illustrated by the gray shaded area in Fig. 1(b), illustrating the resonance energy dependence of bulk damping based on the complex dielectric function of bulk gold.30,36 Second, chemical interface damping (CID) leads to a broadening of the DFS linewidth, although this effect is not readily apparent and requires further analysis that removes the bulk and radiation damping contributions to the linewidth, as shown below.37
Spectral characteristics of the AuNRs on different substrates. (a) PL emission maximum vs DFS maximum for each AuNR. Black asterisks indicate the three representative AuNRs in [insets of (c)]. The black line corresponds to a 1:1 correlation. Error estimates from Lorentzian fitting are ±1 nm for DFS λmax and ±2 nm for PL λmax, roughly the size of the inner dimensions of the diamond markers. (b) DFS linewidth vs DFS maximum for each AuNR. The shaded area represents the bulk damping contribution to the linewidth. Error estimates from Lorentzian fitting are ±1 nm for DFS λmax and ±1 meV for DFS Γ, roughly the size of the inner dimensions of the diamond markers. (c) DFS spectra of three representative AuNRs each 32 × 90 ± 3 nm. Inset: SEM micrographs of the AuNRs. The image width is 150 nm. (d) PL spectra of the same three representative AuNRs. The vertical bars illustrate the integrated area for the scaled QY analysis at λmax ± 3 nm.
Spectral characteristics of the AuNRs on different substrates. (a) PL emission maximum vs DFS maximum for each AuNR. Black asterisks indicate the three representative AuNRs in [insets of (c)]. The black line corresponds to a 1:1 correlation. Error estimates from Lorentzian fitting are ±1 nm for DFS λmax and ±2 nm for PL λmax, roughly the size of the inner dimensions of the diamond markers. (b) DFS linewidth vs DFS maximum for each AuNR. The shaded area represents the bulk damping contribution to the linewidth. Error estimates from Lorentzian fitting are ±1 nm for DFS λmax and ±1 meV for DFS Γ, roughly the size of the inner dimensions of the diamond markers. (c) DFS spectra of three representative AuNRs each 32 × 90 ± 3 nm. Inset: SEM micrographs of the AuNRs. The image width is 150 nm. (d) PL spectra of the same three representative AuNRs. The vertical bars illustrate the integrated area for the scaled QY analysis at λmax ± 3 nm.
In addition to linewidth changes due to different interfaces, we see a clear PL QY dependence on the substrate for the three example AuNRs presented in Fig. 1(d). We use PL spectroscopy to probe the hot carrier dynamics within the AuNRs under the assumption that PL is due to the Purcell factor enhanced radiative recombination of charge carriers or plasmon emission.26,38–41 The substrate effect on QY is also twofold. First, the resonance quality factor (Eres/Γ)28,42–44 affects the QY as it is directly related to the Purcell factor that enhances the radiative recombination of charge carriers within the AuNR or the plasmon density of states.21,26,40,41,45 Thus, the substrate modifies the QY through resonance energy shifts and changes in linewidth (due to both bulk damping and CID). We account for these effects by scaling the QY by the Purcell factor (Fig. S2).21 Second, the QY depends on the radiative recombination of charge carriers or plasmon emission.26,38,46 If the excited charge carriers are injected into the semiconductor material at the interface, we expect a reduction in the PL QY.26
Based on the energetic alignment of the AuNR and semiconductor bands (Table I and Fig. 2), we expect to observe changes to the plasmon damping in the DFS linewidth due to PICT. PICT provides an additional plasmon decay channel affecting the scattering linewidth via CID,47 where the total linewidth is described by the phenomenological model, Γ = ΓBulk + ΓRad + ΓCID. The contributing components to the total linewidth includes ΓBulk, which stands for the resonance energy dependent bulk gold damping having a constant intraband contribution of 73 meV and an interband component beginning around 690 nm and increasing with shorter wavelengths. Γrad represents the volume dependent radiation damping, and ΓCID stands for CID.47–50 Each term, except for CID, can be accounted for from the dielectric function or size of the nanoparticle and can therefore be subtracted from the total linewidth to obtain the contribution from CID only. The CID term is a catchall term for interfacial effects, including resonance energy transfer, charge transfer, and dipole scattering.12 Here, we calculate the linewidth damping due to CID and relate it to the PICT mechanism of charge transfer into ITO and TiO2 compared to SiO2. We exclude an energy transfer mechanism due to the high bandgaps (>3.2 eV) for our materials compared to the 488 nm (2.54 eV) excitation source for PL and <2.2 eV plasmon resonance energy.
. | Conduction band minimum (ϕCBM) (eV) . | Schottky barrier height (ϕSB = EF − ϕCBM) (eV) . | Interband barrier height (ϕIB = EF + EIB − ϕCBM) (eV) . |
---|---|---|---|
SiO2 | 0.8 | 4.3 | 6.5 |
TiO2 | 4.2 | 0.9 | 3.1 |
ITO | 4.9 | 0.2 | 2.4 |
. | Conduction band minimum (ϕCBM) (eV) . | Schottky barrier height (ϕSB = EF − ϕCBM) (eV) . | Interband barrier height (ϕIB = EF + EIB − ϕCBM) (eV) . |
---|---|---|---|
SiO2 | 0.8 | 4.3 | 6.5 |
TiO2 | 4.2 | 0.9 | 3.1 |
ITO | 4.9 | 0.2 | 2.4 |
Illustration of the relative positions for the energies given in Table I.
We accounted for the resonance energy dependence on the plasmon linewidth by subtracting the bulk damping contribution. We calculated the bulk damping contribution according to , where γb is the Drude damping, ɛIB is the imaginary part of the dielectric function due to interband absorption, Eres is the plasmon resonance energy, and Ep is the free electron plasma energy.36 The values for ɛIB were calculated by subtracting the free electron Drude model from the imaginary part of the dielectric data [] from Ref. 30 according to using values of 9.3 eV for Ep and 73 meV for γb.30,36 Any negative ɛIB values, which have no physical meaning, were set to 0, which only occurred below the interband onset and would lead to artificial linewidth narrowing in the intraband region. The bulk damping contribution to the linewidth is represented by the gray area in Fig. 1(b).56 We note that the choice of gold dielectric data and Drude parameters affects the magnitude of ΓCID but not the substrate dependent trends (Fig. S3). We chose the particular dielectric data published in Ref. 30 to minimize any spectral dependence on the linewidth for AuNRs on SiO2, which we expect to be spectrally flat after the subtraction of the bulk damping contribution. We accounted for the size dependent radiation damping contribution according to Γrad = hκV/π, where we determined κ = 3.34 × 10−7 fs−1 nm−3, consistent with previously reported values,48,49 by fitting the data for the SiO2 substrate after subtracting the bulk linewidth contribution. Here, V is the AuNR volume as determined by scanning electron microscopy, κ is a phenomenological constant, and h is Planck’s constant. We attribute the remainder of the single-particle plasmon linewidth, after subtracting the bulk and radiation damping contributions, to CID caused by the TiO2 and ITO interfaces, as discussed next in Fig. 3(a). The quantum efficiencies of each of the damping processes are calculated and presented in Table II. Our average CID quantum efficiency (ϕCID) for TiO2 is lower than what is observed in Ref. 56, where ϕCID = 0.3. We attribute this reduction in ϕCID to the presence of surfactant between the AuNRs and the substrate and a smaller interfacial contact between the AuNRs and the TiO2 due to the sample geometry: AuNRs lying on TiO2 vs overcoated with TiO2. Thus, our results represent a lower estimate for CID compared to what should be possible with optimized samples.
Changes in single AuNR optical properties on semiconducting substrates (ITO and TiO2) compared to the insulating SiO2 substrate. (a) Bulk and radiation damping adjusted linewidth for AuNRs on each substrate. ΓCID increases on ITO (red) and TiO2 (yellow) substrates compared to SiO2 (blue). (b) Purcell factor adjusted QY substrate comparison. The PL intensity is partly quenched on ITO (red) and significantly quenched on TiO2 (yellow) compared to SiO2 (blue). The red bars are the medians of the distributions, while blue boxes represent the interquartile range of the distributions.
Changes in single AuNR optical properties on semiconducting substrates (ITO and TiO2) compared to the insulating SiO2 substrate. (a) Bulk and radiation damping adjusted linewidth for AuNRs on each substrate. ΓCID increases on ITO (red) and TiO2 (yellow) substrates compared to SiO2 (blue). (b) Purcell factor adjusted QY substrate comparison. The PL intensity is partly quenched on ITO (red) and significantly quenched on TiO2 (yellow) compared to SiO2 (blue). The red bars are the medians of the distributions, while blue boxes represent the interquartile range of the distributions.
Average quantum efficiencies of electron–hole generation (ϕeh), radiative decay (ϕrad), and charge transfer (ϕCID) for the different substrates.
. | ϕeh = ΓBulk/Γtot . | ϕrad = Γrad/Γtot . | ϕCID = ΓCID/Γtot . |
---|---|---|---|
SiO2 | 0.78 ± 0.08 | 0.20 ± 0.08 | 0.02 ± 0.05 |
ITO | 0.77 ± 0.1 | 0.18 ± 0.07 | 0.05 ± 0.1 |
TiO2 | 0.72 ± 0.1 | 0.17 ± 0.07 | 0.11 ± 0.06 |
. | ϕeh = ΓBulk/Γtot . | ϕrad = Γrad/Γtot . | ϕCID = ΓCID/Γtot . |
---|---|---|---|
SiO2 | 0.78 ± 0.08 | 0.20 ± 0.08 | 0.02 ± 0.05 |
ITO | 0.77 ± 0.1 | 0.18 ± 0.07 | 0.05 ± 0.1 |
TiO2 | 0.72 ± 0.1 | 0.17 ± 0.07 | 0.11 ± 0.06 |
The CID induced linewidth broadening (ΓCID) has a mean of 3 meV for the AuNRs on SiO2 [Fig. 3(a), blue] and is negligible, as expected, since SiO2 is an insulating material for which, based on the energetic band alignments, no charge transfer between the AuNR and SiO2 can occur (Table I). Thus, SiO2 serves as a reference case of no ΓCID due to charge transfer. ITO shows a modest increase in the mean ΓCID to 7 meV but with a large distribution [Fig. 3(a), red]. The AuNRs on TiO2 experienced the largest increase in ΓCID with a mean of 12 meV [Fig. 3(a), yellow]. This value of ΓCID for TiO2 is lower than previously measured in Ref. 56, likely because those AuNRs were overcoated with TiO2, while here the AuNRs are deposited on the surface of a TiO2 film leading to a smaller surface area contact between the AuNRs and TiO2. A one-way analysis of variance [F(2167) = 15.15, p = 8.95 × 10−7] with multiple comparisons test [Tukey honest significant difference (HSD)]57 of ΓCID for AuNRs on the different substrates finds differences between the substrates. The difference in mean ΓCID values between SiO2 and ITO (p = 0.055) is slightly above the significance threshold (p ≤ 0.05 is significant), while the differences between TiO2 and SiO2 or ITO are significantly different (p = 0.005 and p = 1.3 × 10−7, respectively). Thus, PICT, as measured by ΓCID, occurs in TiO2 and to a smaller extent in ITO. This interpretation of CID due to a PICT mechanism is consistent with observations of free carrier absorption in TiO2 following plasmon excitation.58,59
In addition to ΓCID, we find that the PL QY serves as a spectroscopic signature of electron transfer. In order to properly compare the QY between the substrates, we calculated a scaled QY by dividing the QY at the resonance energy (λmax ± 3 nm) by the Purcell factor.21 The results are summarized in Fig. 3(b). Again, SiO2 [Fig. 3(b), blue] serves as a standard of no charge transfer with a mean scaled QY of 17 × 10−13. Scaled QY quenching is visible for AuNRs on both the ITO and TiO2 substrates with mean scaled QYs of 11 × 10−13 and 7 × 10−13, respectively [Fig. 3(b), red and yellow]. We also performed a one-way analysis of variance [F(2167) = 129.9 and p = 9.5 × 10−35] between the three substrates for the scaled QY, confirming significant differences. A multiple comparisons test (Tukey HSD) reveals that the quenching on ITO and TiO2 seen in Fig. 3(b) is highly significant different from SiO2 (p = 9.56 × 10−10 and p = 9.56 × 10−10, respectively) with TiO2 and ITO also highly significantly different from each other (p = 9.86 × 10−10). Based on our interpretation of PL, we conclude that there is a reduction in the average number of excited charge carriers able to radiatively recombine occurs when the AuNRs are in interfacial contact with ITO or TiO2, indicating that the charges are injected into the adjacent material.
Our correlated single-particle measurements enable us to also investigate the effect of the size and resonance energy on the ΓCID and scaled QY for the different substrates. Figure S4 illustrates the AuNR size dependence of ΓCID and scaled QY. The inverse effective electron path length (1/Leff), which is directly related to the surface area to volume ratio via Leff = 4V/S, has been previously used to describe surface scattering contributions to the DFS linewidth.49,60 The AuNRs on SiO2 exhibit no correlation between 1/Leff and ΓCID or scaled PL. While previous work has reported a correlation between 1/Leff and linewidth, the range of AuNR sizes is small here and in the size range where surface scattering is relatively weak.61 Meanwhile, the AuNRs on ITO and TiO2 demonstrate a weak negative correlation between size and ΓCID, where AuNRs with larger 1/Leff have smaller ΓCID (i.e., smaller AuNRs show less CID), but no correlation was observed with the scaled PL. The observed correlation between 1/Leff and ΓCID for AuNRs on ITO and TiO2 is contrary to the relationship found in Ref. 61, although the mechanism for CID there was likely due to a change in molecular induced surface dipole moment upon adsorbate exchange.62,63 Again, the 1/Leff ranges here are small and fall in the region where radiation damping dominates over surface scattering,49,61 and the dependence is weak. One possibility for why this negative correlation exists is through substrate induced changes in the radiation damping term κ, which is dependent on the permittivity of the surrounding material.64
In Figs. 4(a)–4(c), we identify relationships between the plasmon resonance wavelength and ΓCID for AuNRs on ITO but not SiO2 and not a statistically significant dependence in TiO2. The AuNRs on SiO2 demonstrate no correlation between the scattering resonance wavelength and ΓCID, consistent again with SiO2 serving as our standard. The AuNRs on ITO exhibit a negative correlation between the resonance energy and ΓCID. The observation that AuNRs with shorter wavelength resonances experience larger ΓCID generally follows a modified Fowler–Nordheim equation,65 describing the yield of electron transfer as proportional to , where c is a geometric parameter that accounts for enhancement, set to 18 for ΓCID and 18 × 10−13 for the scaled QY (obtained by fitting the ITO data), and b is a variable offset. TiO2 does not exhibit a ΓCID wavelength dependence since the Schottky barrier is higher for TiO2 (ϕSB = 0.9 eV), and therefore, the excess energy is small, i.e., a shallow regime in the Fowler–Nordheim equation where the variance in the data is larger than any expected trend [Fig. 4(c), shaded region and solid yellow line]. For ITO, a higher excess energy leads to more efficient charge injection into the ITO for higher energy resonances, given the low Schottky barrier (ϕSB = 0.2 eV), i.e., a steep regime in the Fowler–Nordheim equation [Fig. 4(b), solid red line].11,66
Changes in single AuNR spectra on semiconducting substrates (ITO: red and TiO2: yellow) compared to the insulating SiO2 (blue) substrate. (a)–(c) Bulk and radiation damping adjusted DFS linewidth vs scattering resonance wavelength for AuNRs on SiO2 (a), ITO (b), and TiO2 (c). (d)–(f) Scaled QY vs scattering resonance wavelength. The PL intensity is partly quenched on the ITO substrate (e) and greatly quenched on the TiO2 substrate (f) compared to the insulating SiO2 (d) substrate. Dashed lines illustrate the mean (μ) of the data, and the shaded area indicates ±1 standard deviation (σ) around the mean. r is the Pearson correlation coefficient, which can have values from −1 (perfect negative linear correlation) to 1 (perfect positive linear correlation), and p is the confidence value for r where p < 0.05 is considered statistically significant. Black asterisks indicate the three representative AuNRs in Fig. 1. Solid lines are guides to the eye assuming a Fowler–Nordheim dependence.
Changes in single AuNR spectra on semiconducting substrates (ITO: red and TiO2: yellow) compared to the insulating SiO2 (blue) substrate. (a)–(c) Bulk and radiation damping adjusted DFS linewidth vs scattering resonance wavelength for AuNRs on SiO2 (a), ITO (b), and TiO2 (c). (d)–(f) Scaled QY vs scattering resonance wavelength. The PL intensity is partly quenched on the ITO substrate (e) and greatly quenched on the TiO2 substrate (f) compared to the insulating SiO2 (d) substrate. Dashed lines illustrate the mean (μ) of the data, and the shaded area indicates ±1 standard deviation (σ) around the mean. r is the Pearson correlation coefficient, which can have values from −1 (perfect negative linear correlation) to 1 (perfect positive linear correlation), and p is the confidence value for r where p < 0.05 is considered statistically significant. Black asterisks indicate the three representative AuNRs in Fig. 1. Solid lines are guides to the eye assuming a Fowler–Nordheim dependence.
For the scaled QY, we again see that SiO2 exhibits no correlation with resonance wavelength [Fig. 4(d)], thus scaling the QY by the Purcell factor proves to be an appropriate method to account for dielectric and CID effects on the QY. In contrast, for both the ITO and TiO2 substrates, we see a negative correlation in scaled QY with resonance wavelength [Figs. 4(e) and 4(f)]. This correlation is more pronounced on the ITO substrate [Fig. 4(e)]. We attribute this trend to higher rates of back electron transfer followed by radiative recombination for higher energy plasmon resonances.67 Thus, the scaled QY for lower energy resonances is quenched, while the scaled QY for the highest energy resonances is unaffected. This argument is akin to the trend in Fig. 4(b), where a higher plasmon energy results in a higher charge transfer rate in the forward direction, resulting in a loss in plasmon coherence and thus an increase in ΓCID. For the scaled QY quenching, the trend is due to a combination of forward and back electron transfer. Considering the same small energy barrier in both directions for ITO, the observed resonance energy dependence should have similar impacts for forward and back electron transfer, i.e., back electron transfer followed by radiative recombination occurs at a higher rate for plasmons with higher resonances, resulting in a negative correlation between the resonance wavelength and scaled QY. It is important to note that these observed trends for the PL QY clearly depend on the plasmon resonance even though a plasmon is not initially excited in the PL experiments using 488 nm light. We will discuss the underlying mechanism below.
We performed a similar analysis for the interband region of the PL spectrum at 525 ± 5 nm (Fig. S5). Here, the QY and not the Purcell factor scaled QY was used since the longitudinal plasmon resonance does not overlap with this spectral region and the transverse plasmon is weak and less sensitive to local refractive index changes; thus, Purcell factor enhancement is not considered.26 We expect the interband QY (QYIB) to be unaffected by the TiO2 substrate because the primary electrons generated through 488 nm interband excitation lack the energy to overcome the interband barrier height (ϕIB) between gold and TiO2 (hν < ϕIB, where hν = 2.54 eV is the photon energy and ϕIB = 3.1 eV; Table I). Indeed, AuNRs on TiO2 and SiO2 substrates have nearly identical values for QYIB with means of 1.25 × 10−8 and 1.23 × 10−8, respectively (p = 0.98). Interestingly, AuNRs on ITO experience an enhancement in QYIB with a mean of 2.36 × 10−8 (p = 5.6 × 10−10). We attribute this enhancement to interband excited charge injection since 488 nm interband excitations have a high enough energy to overcome the interband barrier height for ITO (hν > ϕIB, where ϕIB = 2.4 eV; Table I), followed by either fast back electron transfer and then radiative recombination or potentially direct radiative recombination from an electron at the bottom of the ITO conduction band to the d-band hole formed in gold. The latter mechanism implies that effectively a larger density of electronic states is available for PL through the participation of electrons in ITO and explains the increased QY. This observation and interpretation, however, merits future investigations.
DISCUSSION
The results of these studies lead us to propose several distinct mechanisms for charge transfer between AuNRs and ITO or TiO2, as schematically illustrated in Fig. 5. For AuNRs on SiO2, the plasmon energy is much smaller than the required energy to reach the SiO2 conduction band and thus no charge transfer occurs [Figs. 3(a) and 5(a)]. Additionally, when AuNRs on SiO2 are excited with a 488 nm laser, the interband excited electrons even after Auger scattering lack the necessary energy and therefore no charge transfer occurs [Figs. 3(b) and 5(e)].68,69 These properties allow SiO2 to serve as a standard for which to compare the ΓCID values and scaled QYs of TiO2 and ITO.
Schematics of PICT (a)–(c) and Auger scattered electron injection (e)–(g) for different substrates and corresponding expected spectral changes (d) and (h). (a) PICT into SiO2 is inhibited by a large Schottky barrier. (b) PICT proceeds with electron injection into TiO2. (c) Large resonance energy dependent PICT proceeds with electron injection into ITO due to a small Schottky barrier and the degenerate semiconductor nature of ITO.67,70 (d) A cartoon of DFS linewidth broadening on the semiconductor substrates due to CID via the PICT mechanism. (e) 488 nm (2.54 eV) interband laser excitation of electron–hole pairs do not result in charge transfer into SiO2. (f) Interband excited electron–hole pairs scatter to form a plasmon resonance dependent electron energy distribution that transfers into the conduction band. (g) Interband excited electron–hole pairs scatter to form a plasmon that directly injects an electron into the semiconductor conduction band. Note that mechanisms (f) and (g) are possible for both ITO and TiO2 and are not mutually exclusive. (h) A cartoon of PL QY quenching due to charge transfer reducing plasmon PL and interband PL QYIB enhancement in ITO. Energy level values for SiO2, TiO2, ITO, and Au are listed in Table I.
Schematics of PICT (a)–(c) and Auger scattered electron injection (e)–(g) for different substrates and corresponding expected spectral changes (d) and (h). (a) PICT into SiO2 is inhibited by a large Schottky barrier. (b) PICT proceeds with electron injection into TiO2. (c) Large resonance energy dependent PICT proceeds with electron injection into ITO due to a small Schottky barrier and the degenerate semiconductor nature of ITO.67,70 (d) A cartoon of DFS linewidth broadening on the semiconductor substrates due to CID via the PICT mechanism. (e) 488 nm (2.54 eV) interband laser excitation of electron–hole pairs do not result in charge transfer into SiO2. (f) Interband excited electron–hole pairs scatter to form a plasmon resonance dependent electron energy distribution that transfers into the conduction band. (g) Interband excited electron–hole pairs scatter to form a plasmon that directly injects an electron into the semiconductor conduction band. Note that mechanisms (f) and (g) are possible for both ITO and TiO2 and are not mutually exclusive. (h) A cartoon of PL QY quenching due to charge transfer reducing plasmon PL and interband PL QYIB enhancement in ITO. Energy level values for SiO2, TiO2, ITO, and Au are listed in Table I.
For AuNRs on TiO2, the band alignments suggest that plasmon resonance energies above 0.9 eV can inject electrons from the AuNR into the TiO2 conduction band [Fig. 5(b)]. We observe PICT spectroscopically via an increase in the ΓCID compared to AuNRs on SiO2 [Fig. 3(a)]. Additionally, we measure no resonance energy dependence for ΓCID [Fig. 4(c)], although this observation may be due to either being in a shallow region of a Fowler–Nordheim dependence or is similar to the lack of an energy dependence observed in Ref. 10. When the AuNRs on TiO2 are excited with a 488 nm laser, the interband excited electrons lack the energy to overcome the Schottky barrier, similar to the AuNRs on SiO2 [Fig. 5(e)]. The inability of the initially excited d-band electrons to inject into the TiO2 conduction band is supported by the lack of change in QYIB compared to SiO2 (Fig. S5). Unlike the AuNRs on SiO2, there is a large scaled QY quenching at the plasmon resonance [Fig. 3(b)], indicating that while the excited d-band electrons are unable to directly overcome the Schottky barrier, Auger scattered electrons can overcome the Schottky barrier. These secondary electrons can have a maximum energy of Ef + hν but are not necessarily dependent on the plasmon resonance energy. However, we nevertheless observe a weak resonance energy dependence for the scaled QY [Fig. 4(f)] and, therefore, postulate that the plasmon resonance must play a role in interfacial charge transfer although we are exciting off resonance at 488 nm. This resonance energy dependence, which is more pronounced for ITO (discussed below), can be described in two ways: (1) Auger scattered electrons have an energy distribution controlled by the density of electronic states modified by the plasmon, and thus, the Auger scattered electron energy maximum is Ef + Eres [Fig. 5(f)]; or (2) charge carrier scattering generates a plasmon that can directly inject an electron with energy Ef + Eres through the PICT pathway [Fig. 5(g)].35,40,41,46 The resonance energy dependencies of these mechanisms can lead to the observed trend in Fig. 4(f), where higher energy resonances result in electron injection higher in the TiO2 conduction band, therefore enabling more efficient back electron transfer followed by PL.
Invoking a plasmon in this way, effectively corresponding to the reverse process of plasmon decay, has been suggested before and becomes especially necessary for the ITO substrate.40,41 For AuNRs on ITO, the PICT pathway is available for plasmon resonances with energies above 0.2 eV [Fig. 5(c)]. Similar to AuNRs on TiO2, we observe PICT via an increase in ΓCID compared to AuNRs on SiO2 [Fig. 3(a)]. Unlike for AuNRs on TiO2, on ITO we observe a pronounced resonance energy dependence in ΓCID [Fig. 4(b)]. We assign this observation to being in a steep region of a Fowler–Nordheim dependence compared to AuNRs on TiO2. For the AuNRs on ITO, we also observe quenching of the scaled QY compared to SiO2 [Fig. 3(b)], with a magnitude smaller than for TiO2 while now following strong plasmon resonance energy dependence [Fig. 4(e)]. When excited with a 488 nm laser, the initially created interband excited electrons have enough energy to directly overcome the barrier. If charge transfer were to only occur through these primary electrons, we should not expect a dependence on the plasmon resonance energy. However, because we clearly observe a resonance energy dependence, the same mechanisms of charge transfer leading to scaled QY quenching for AuNRs on TiO2 must be present for AuNRs on ITO, i.e., Auger scattered electron injection into the ITO conduction band either through plasmonic modification of the density of electronic states [Fig. 5(f)] or scattering induced plasmon excitation followed by PICT [Fig. 5(g)]. We, therefore, conclude that it is not just the primary interband excitation generated electrons, which are injected into the semiconductor, but secondary charge carrier scattering events mediated by the plasmon must also play a significant role to explain the observed plasmon resonance dependence of PL quenching on ITO and TiO2.
Finally, while we expected QYIB to be quenched due to charge transfer for AuNRs on ITO, we observed an enhancement of QYIB compared to AuNRs on SiO2 (Fig. S5). We hypothesize that this enhancement is either due to a larger density of electronic states leading to direct radiative recombination of an electron at the bottom of the ITO conduction band with the d-band hole formed in gold or fast back electron transfer followed by radiative recombination.
CONCLUSIONS
We have demonstrated the ability to spectroscopically detect charge transfer through single-particle DFS and PL spectroscopy. Specifically, we identified scattering linewidth broadening and scaled QY quenching as spectroscopic signatures of charge transfer following both plasmon and interband excitation. For TiO2, we observed PICT and Auger scattered charge transfer with low electron back transfer rates. For ITO, which is a degenerate semiconductor with a conduction band close to the Fermi energy of gold, we observed three mechanisms of charge transfer: resonance energy dependent PICT, Auger scattered charge transfer, and interband excited electron transfer. These processes had higher electron back transfer rates than TiO2 because of the lower Schottky barrier between ITO and gold. These two complementary spectral signatures of charge transfer have utility in predicting the performance of plasmonic photoelectric and photocatalytic devices,1,71,72 without needing to have a completed device, thus allowing evaluation of the device’s individual components, such as a hole acceptor in the absence of an electron acceptor. Further studies into these mechanisms and spectroscopic signatures include expanding the range of AuNR sizes to further characterize predicted dependencies, expanding the resonance wavelength range, and comparing other metal-oxide semiconductors possessing different barrier heights and carrier mobilities. Investigating other spectral signatures, such as changes to the absorption spectrum, could provide additional insight into plasmon-assisted charge transfer mechanisms.
SUPPLEMENTARY MATERIAL
See the supplementary material for supplemental Figs. S1–S5.
ACKNOWLEDGMENTS
This material is based upon work supported by the National Science Foundation under the NSF Center for Adapting Flaws into Features (Grant No. CHE-2124983). S.L. acknowledges the Robert A. Welch Foundation for support through the Charles W. Duncan, Jr.-Welch Chair in Chemistry (C-0002). This work was conducted, in part, using resources of the Shared Equipment Authority at Rice University.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
The manuscript was written through the contributions of all authors. All authors have given approval to the final version of the manuscript.
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.