Ternary semiconductors such as AgInS2, with their interesting photocatalytic properties, can serve as building blocks to design light harvesting assemblies. The intraband transitions created by the metal ions extend the absorption well beyond the bandgap transition. The interfacial electron transfer of AgInS2 with surface bound ethyl viologen under bandgap and sub-bandgap irradiation as probed by steady state photolysis and transient absorption spectroscopy offers new insights into the participation of conduction band and trapped electrons. Capping AgInS2 with CdS shifts emission maximum to the blue and increases the emission yield as the surface defects are remediated. CdS capping also promotes charge separation as evident from the efficiency of electron transfer to ethyl viologen, which increased from 14% to 29%. The transient absorption measurements that elucidate the kinetic aspects of electron transfer processes in AgInS2 and CdS capped AgInS2 are presented. The improved performance of CdS capped AgInS2 offers new opportunities to employ them as photocatalysts.

Attaining clean renewable energy as an alternative to unsustainable fossil fuels continues to gain traction across multidisciplinary fields. Of particular interest are ternary I–III–VI2 semiconductors, such as AgInS2 and CuInS2, which are finding applications in light-emitting diodes (LEDs),1–5 photocatalysis,6–8 batteries,9 and photovoltaic devices.10–15 The photophysical properties of I–III–VI2 semiconductors are complex and differ from their analog, cadmium-based binary chalcogenides. Binary semiconductor systems, such as CdTe and CdSe, exhibit distinct excitonic peaks, narrow photoluminescence (PL) bandwidths, and short emission lifetimes.16–18 These excitonic peaks and sharp absorption onsets are indicative of the dominant band edge transitions. These semiconductor nanocrystals also exhibit relatively small Stokes shift, which is indicative of dominance of band–band charge carrier recombination channels with minimal contribution from intraband states.16 

In contrast to II–VI semiconductors, the optical properties of I–III–VI2 systems are characterized by a lack of distinct excitonic peaks and emergence of a long tail in their absorption spectra, large Stokes shifts, broad PL bandwidths,19,20 and long PL lifetimes.20–25 The spectral features arise from surface and interstitial defects (trap states) and donor–acceptor pairs (DAPs), which act as radiative and non-radiative recombination sites.24,26–28 The surface trap states originate from the non-bonding orbitals of the undercoordinated surface atoms. Proper choice of ligands can passivate the surface of quantum dots (QDs) if they bond to the undercoordinated surface atoms. These uncoordinated atoms yield electronic states that lie within the bandgap of the semiconductor and, thus, act as a trap for electrons and holes.29 The use of a higher bandgap material as a shell to the AgInS2 core has been seen to significantly improve its photophysical properties. ZnS shell on AgInS2 is reported to have reduced surface trap centers and, thus, enhance the PL quantum yield (PLQY).19,30 Therefore, it is intriguing to study how capping AgInS2 with CdS affects the photophysical properties, particularly interfacial charge transfer processes that are relevant to photocatalysis.

To achieve efficient conversion of light energy into chemical energy, it is important to maximize charge separation by suppressing the charge recombination within the semiconductor nanocrystal as well as back electron transfer from the electron transfer product. Careful design strategy is required to suppress these undesired processes. Viologen molecules are useful probes to evaluate the interfacial electron transfer properties as they can capture electrons and yield a stable reduced viologen radical in an inert atmosphere. By tracking the formation of viologen radicals, one can elucidate kinetic factors limiting interfacial electron transfer processes.31,32 We have now employed ethyl viologen, EV2+ having one-electron reduction potential (E0 = −0.449 V vs NHE) less negative than the conduction band (CB) of the AgInS2 and AgInS2–CdS QDs (ECB ∼ −1.0 V vs NHE) can readily capture photogenerated electrons.31 Using EV2+ as a probe, we have now succeeded in elucidating the involvement of conduction band and sub-bandgap states in promoting electron transfer. The role of AgInS2–CdS heterostructure in enhancing the efficiency of charge transfer is also discussed.

AgInS2 and AgInS2–CdS QDs were synthesized by using modified hot-injection methods. Experimental details of the synthesis of AgInS2 and AgInS2–CdS QDs are presented in the supplementary material (Fig. S1). The transmission electron microscopy (TEM) images of the two QDs are shown in Figs. 1(a), 1(b), 1(d), and 1(e). The size distribution of the QDs was determined by measuring the edge length of ∼150 QDs [see Figs. 1(c) and 1(f)]. AgInS2 and AgInS2–CdS are pyramidal shaped with average edge lengths of 4.8 ± 0.6 and 5.5 ± 0.9 nm, respectively. The slight increase in size for the AgInS2–CdS can be attributed to the introduction of the desired thin CdS shell. This thin shell is able to remediate the surface defects without blocking the charge transfer at the heterostructure interface. In addition, these nearly monodispersed particles show a tail feature in absorption and broad emission, as shown in Figs. 2(a), 2(c), and 2(d), hence confirming that polydispersity is not the key source of such features in these materials.33 

FIG. 1.

TEM image for (a) and (b) AgInS2 and (d) and (e) AgInS2–CdS. Size distribution statistics of (c) AgInS2 and (f) AgInS2–CdS. ∼150 particles were sized by using ImageJ software from the TEM images.

FIG. 1.

TEM image for (a) and (b) AgInS2 and (d) and (e) AgInS2–CdS. Size distribution statistics of (c) AgInS2 and (f) AgInS2–CdS. ∼150 particles were sized by using ImageJ software from the TEM images.

Close modal
FIG. 2.

(a) Absorption spectra of (a) AgInS2 and (b) AgInS2–CdS in toluene. (b) Emission decay traces of (a) AgInS2 and (b) AgInS2–CdS monitored at 750 and 650 nm, respectively (λexc − 355 nm, see Table S1 for fitting parameters). (c) PL excitation (PLE) alongside absorbance and emission spectra of (c) AgInS2 and (d) AgInS2–CdS. PLE excited at different wavelengths to establish the origin of the optical transitions involved.

FIG. 2.

(a) Absorption spectra of (a) AgInS2 and (b) AgInS2–CdS in toluene. (b) Emission decay traces of (a) AgInS2 and (b) AgInS2–CdS monitored at 750 and 650 nm, respectively (λexc − 355 nm, see Table S1 for fitting parameters). (c) PL excitation (PLE) alongside absorbance and emission spectra of (c) AgInS2 and (d) AgInS2–CdS. PLE excited at different wavelengths to establish the origin of the optical transitions involved.

Close modal

The absorption and emission properties of AgInS2 and AgInS2–CdS are presented in Figs. 2(a)2(d). The absorption spectra [Fig. 2(a)] show absorption below 700 nm. The intraband transitions that induce tail absorption24,26–28 slightly shift to blue upon capping with CdS. Both of these nanocrystals continue to show broad emission in the visible-near IR region. These broad emissions imply that the PL originates from the recombination of carriers trapped by intraband levels and not exciton recombination.6,24,34

In accordance with the blue shift in absorbance, we also see a shift in the emission spectra from ∼750 nm (AgInS2) to ∼650 nm (AgInS2–CdS). The excitonic transition that appears near the band edge (∼500 nm) is not visible in the absorption spectrum. It is buried in the tail absorption and, hence, cannot be resolved from absorption spectra. The excitation spectra recorded at different monitoring wavelengths [Figs. 2(c) and 2(d)], however, show the origin of different optical transitions contributing to the emission. When monitored at longer wavelength (e.g., 800 nm), we see excitation spectra that closely follow the absorption spectra indicating contributions from bandgap and intrabandgap transitions. When the monitoring wavelengths are set close to the band edge (e.g., 570 nm), we see the appearance of a shoulder in the range of 450–500 nm region. This shoulder that corresponds to excitonic transition indicates a bandgap of ∼2.5 eV, similar to the one reported elsewhere.20,35 Thus, the excitation spectra presented in Figs. 2(c) and 2(d) allow us to identify the origin of emission from different transitions. Interestingly, the CdS capping seems to have little effect on these transitions that contribute to overall emission.

Ternary semiconductors such as AgInS2 exhibit both direct bandgap excitation and intraband transitions arising from the vacancies and interstitial (antisite) defects (Scheme 1).35–39 The lack of the bandgap (excitonic) emission dominance in the steady state emission spectra has been attributed to the sub picosecond decay of the bandgap states to the intraband states.35,38 Earlier reports show that the emissions of ternary QDs are dominated by donor acceptor pair (DAP) recombination. Sulfur vacancy (VS) and silver interstitial (Agint.) act as plausible defects for donor state, whereas silver vacancy (VAg) and sulfur interstitial (Sint.) act as the acceptor states.25,36,40–42

SCHEME 1.

Plausible absorption and deactivation pathways in AgInS2 QDs. (i) Absorption and recombination from band edge transitions. (ii) Recombination to defect site above the valence band (VB). (iii) Recombination from defect site below the conduction band (CB) to the VB. (iv) Absorption and recombination mediated by defects below the CB. (v) DAP recombination arising from mid-bandgap defects. Note: Solid lines indicated absorption or radiative recombination, while dotted lines indicate nonradiative recombination.

SCHEME 1.

Plausible absorption and deactivation pathways in AgInS2 QDs. (i) Absorption and recombination from band edge transitions. (ii) Recombination to defect site above the valence band (VB). (iii) Recombination from defect site below the conduction band (CB) to the VB. (iv) Absorption and recombination mediated by defects below the CB. (v) DAP recombination arising from mid-bandgap defects. Note: Solid lines indicated absorption or radiative recombination, while dotted lines indicate nonradiative recombination.

Close modal

The emission decay of AgInS2 and AgInS2–CdS nanocrystal suspensions was monitored in a nanosecond laser flash photolysis set up using 355 nm laser pulse excitation [Fig. 2(b)]. Long-lived emission lifetimes of AgInS2 (avg. lifetime of 493 ns) and AgInS2–CdS (avg. lifetime of 300 ns) recorded for these samples indicated the dominance of DAP radiative pathway for the deactivation of the excited state.24,25 Electrons from shallow donors (close to VB) recombining with shallow acceptors (close to CB) give rise to the emission in the higher energy part of the emission band, while deep seated donor and acceptor states are responsible for the emission seen in the lower energy part (red region) of the emission band. The broad emission and tail absorption are independent of shape of the nanocrystals but dependent on the ratio of Ag:In.43 Related works on single particle of the CuInS2 counterpart have also ruled out the possibility of polydispersity33 and size44 as the only major sources of broadening in emission linewidth. Therefore, trapping of charge carriers in the randomly positioned local states within the bandgap (recombination centers) is the plausible reason for the broad emission and long lifetime seen in ternary I–III–VI2 QDs.33,40,43,44

Other models depart from this defect mediated model (DAP) by invoking self-trapped exciton (STE) as another plausible mechanism in CuInS245,46 and AgInS2.44 Knowles et al.45 argue that excitation of CuInS2 nanocrystal from its ground state to a delocalized excitonic state is followed by rapid relaxation to a self-trapped excitonic state. This trapping of exciton is modulated by hole localization (a direct consequence of strong vibronic coupling and nuclear distortion at the copper state). As a result, emission from this self-trapped excitonic state was characterized by a large Stokes shift, longer lifetimes, and a broadened band shape. Similarly, a long tail in the absorption spectrum has been associated with direct excitation of this self-trapped excitonic state.44–46 However, the presence of self-trapped exciton cannot be the sole contributor to the observed emission at longer wavelengths. The absorption tail extending beyond the band edge contributing to the longer wavelength emission [excitation spectra in Fig. 2(d)] and wavelength dependent emission lifetimes suggest that intraband transitions contribute to the observed emission at longer wavelengths.

The adsorbed species play an important role in dictating the excited state interactions of semiconductor nanocrystals.6 For example, the excited semiconductor nanocrystals can undergo energy transfer or electron transfer with the adsorbed molecule depending upon the energy match and/or redox potentials.47–52 In a previous study,53,54 we probed the surface interactions of perovskite nanocrystals with viologen molecules and how these interactions dictate the interfacial electron transfer.

The interaction of AgInS2 and AgInS2–CdS nanocrystals with ethyl viologen was probed through the quenching of emission [Figs. 3(a) and 3(c)]. With increasing concentration of ethyl viologen, we see a decreased emission of nanocrystals indicating an electron transfer to adsorbed viologen species. In the absence of redox species, the QDs under bandgap excitation undergo charge carrier recombination, both radiatively and non-radiatively [reaction (1)]. When a redox molecule (viz., ethyl viologen) is present near the surface, an additional deactivation pathway of electron transfer [reaction (2)] is encountered,

QD+hνQD(h+e)QD+hν(orheat),
(1)
QD(e)+EV2+QD+EV+,
(2)

while EV2+ refers to ethyl viologen and QD represents AgInS2 or AgInS2–CdS.

FIG. 3.

PL quenching of (a) AgInS2 and (c) AgInS2–CdS QDs in toluene:ethanol (87:13% v/v) with varied concentrations of ethyl viologen (EV2+) (λexc = 355 nm). Corresponding double reciprocal plot analysis of emission quenching of (b) AgInS2 and (d) AgInS2–CdS [see expression (4)]. The apparent association constant, Kapp, was obtained from the intercept and the slope (Kapp = y − intercept/slope).

FIG. 3.

PL quenching of (a) AgInS2 and (c) AgInS2–CdS QDs in toluene:ethanol (87:13% v/v) with varied concentrations of ethyl viologen (EV2+) (λexc = 355 nm). Corresponding double reciprocal plot analysis of emission quenching of (b) AgInS2 and (d) AgInS2–CdS [see expression (4)]. The apparent association constant, Kapp, was obtained from the intercept and the slope (Kapp = y − intercept/slope).

Close modal

The quenching of the QD emission observed at relatively low concentrations of EV2+ indicates strong interaction between the QDs and the EV2+. This interaction is represented by the existence of an equilibrium between adsorbed and unadsorbed EV2+ with an apparent association constant, Kapp,

QD(e)+EV2+   Kapp  [QDEV2+],
(3)
1ϕf0ϕf(obs)=1ϕf0ϕf+1Kappϕf0ϕf[EV2+],
(4)

where the observed quantum yield [ϕf(obs)] refers to the resultant emission arising from viologen-bound (ϕf′) and unbound QD (ϕf0).55 The linearity of the double reciprocal plot of observed emission difference and EV2+ concentration [Figs. 3(b) and 3(d)] shows the validity of expression (4). From the slope and intercept of this plot, we obtain the apparent association constant (Kapp) for equilibrium (3). The Kapp values were 6.2 × 106 and 2.0 × 106 M−1 for AgInS2 and AgInS2–CdS, respectively. The high Kapp values indicate that EV2+ molecules are strongly bound to the surface of the QDs. Slightly lower Kapp observed for AgInS2–CdS (as compared to AgInS2) indicates that the influence of CdS in dictating the attachment to the QD surface as opposed to direct attachment to AgInS2 surface.

Elucidation of mechanistic and kinetic details of interfacial electron transfer at semiconductor interface is important to establish the photocatalytic properties of semiconductor nanocrystals.56,57 The photoluminescence quenching results discussed in the previous section point to the electron transfer between excited AgInS2 and AgInS2–CdS nanocrystals and EV2+. Since the product of electron transfer, EV+•, has the spectroscopic fingerprint with maxima at 605 nm (ε = 1.4 × 104 M−1 cm−1),58,59 we can directly probe the production of electron transfer product under steady state irradiation conditions. Figures 4(a) and 4(b) show the accumulation of EV+• when AgInS2 and AgInS2–CdS suspensions containing 100 μM EV2+ were irradiated with visible light (>420 nm). By employing >420 nm excitation wavelengths, both the AgInS2 core are excited without directly exciting the EV2+. Note that the CdS absorption is too small to make any direct contribution to the EV yield. Figure 4(d) shows the increase in EV concentration with a time of irradiation. The EV attains a steady state concentration after about 800 s of irradiation. The attainment of steady state concentration at longer times is reflective of forward and back electron transfer rates becoming equal,31,53,60,61

kf×[EV2+]=kb[EV+].
(5)

It is evident that the AgInS2–CdS heterostructure enables higher steady state concentration of EV reflective of better charge separation. The observed yield of electron transfer product in AgInS2–CdS is almost double that observed in AgInS2.

FIG. 4.

Different absorbance spectra recorded during steady state photolysis corresponding to EV formation due to electron transfer from (a) AgInS2 and (b) AgInS2–CdS QDs. Absorption spectra were recorded at different light exposure times. (c) Scheme showing involvement of both the exciton and intraband states in transfer of electrons to EV2+. Evolution of EV monitored via the growth of absorbance at 605 nm using (d) 420 nm, (e) 515 nm, and (F) 550 nm long pass filters. Control experiment in the absence of any photocatalyst was also carried out[(D-c)]. The initial concentration of EV2+ was 100 µM. Experiments were carried out in deaerated toluene/ethanol 87:13 v/v% with visible light excitation (Xe lamp, 420/515/550 nm long pass filters, 200 mW cm−2).

FIG. 4.

Different absorbance spectra recorded during steady state photolysis corresponding to EV formation due to electron transfer from (a) AgInS2 and (b) AgInS2–CdS QDs. Absorption spectra were recorded at different light exposure times. (c) Scheme showing involvement of both the exciton and intraband states in transfer of electrons to EV2+. Evolution of EV monitored via the growth of absorbance at 605 nm using (d) 420 nm, (e) 515 nm, and (F) 550 nm long pass filters. Control experiment in the absence of any photocatalyst was also carried out[(D-c)]. The initial concentration of EV2+ was 100 µM. Experiments were carried out in deaerated toluene/ethanol 87:13 v/v% with visible light excitation (Xe lamp, 420/515/550 nm long pass filters, 200 mW cm−2).

Close modal

We also conducted irradiation using a long pass filter of 515 nm and tracked EV growth, as shown in Figs. S2(a) and S2(b). Under these irradiation conditions, only AgInS2 can be excited, thus avoiding any direct contributions from CdS excitation. We see the formation of EV but attaining a lower steady state concentration [Fig. 4(e)] because of filtering of excitations below 515 nm. Again, we observed a higher yield of EV with AgInS2–CdS heterostructure. The presence of a CdS shell around AgInS2 facilitates better charge separation and suppresses the back electron transfer.

It should be noted that the experiments in Figs. 4(d) and 4(e), the selected excitation wavelength allowed both band edge and sub-bandgap transitions. To check the participation of the sub-bandgap states, we irradiated the samples using excitation wavelengths >550 nm. At these excitation wavelengths, direct bandgap excitation of CdS or AgInS2 is not feasible. Figure 4(f) shows EV formation as we subject AgInS2 and AgInS2–CdS to sub-bandgap excitations [corresponding spectra are presented in Figs. S2(d) and S2(e)]. Interestingly, the difference in EV yields with AgInS2 and AgInS2–CdS was relatively small. These results indicate that the electrons from the trap states are accessible and can as well be leveraged for photocatalysis [see scheme in Fig. 4(c)]. Under these different excitation wavelengths, we see a higher yield of electron transfer product with AgInS2–CdS NCs. We also carried out control experiments by exposing only EV2+ in deaerated toluene/ethanol 87:13 v/v% (i.e., without QDs) to similar irradiation conditions described above. For the >420 nm excitation wavelengths, we see negligible contribution as a result of direct excitation [see trace (c) Fig. 4(d)]. By utilizing >515 and >550 nm filters, we do not see any generation of EV [Figs. S2(c) and S2(f)].

We also independently measured quantum yields of EV formation using ferrioxalate actinometry.62,63 The details of actinometry measurements are presented in the supplementary material. Under steady state irradiation conditions (>420 nm, 30 min), we obtain a quantum yield (QY) of EV of 14.5% and 29.3% for AgInS2 and AgInS2–CdS nanocrystals, respectively. Although the initial electron transfer from excited AgInS2 could be very high (>90%), a significant fraction of transferred electrons is lost in back electron transfer (vide infra). This increase in electron transfer yield is consistent with the results obtained with selective bandgap excitation of AgInS2 core in Figs. 4(d) and 4(e). The higher EV QY observed for AgInS2–CdS further supports the argument that the heterostructure configuration of AgInS2–CdS facilitates improved charge separation.

Transient absorption spectroscopy is convenient to probe the ultrafast electron transfer events following the laser pulse excitation of AgInS2 and AgInS2–CdS nanocrystals. When excited with 400 nm laser pulse excitation, AgInS2 shows a broad bleach in the 500–750 nm region [Fig. 5(a)]. The excited state is long-lived as evident from the slow bleach recovery. Corresponding time-resolved transient absorption spectra of AgInS2–CdS NCs are shown in the supplementary material (Fig. S3). The recovery becomes faster in the presence of EV2+ [Fig. 5(b)]. This trend further confirms that the ultrafast electron transfer to EV2+ is the preferred pathway for excited state deactivation of both AgInS2 and AgInS2–CdS QDs. The fast-bleaching recovery recorded in nanosecond time scale at different EV2+concentrations for these two suspensions represents kinetics of conduction band electrons [Figs. 5(c) and 5(d)]. The residual bleach at the end of these traces was further analyzed on a longer microsecond time scale [Figs. 5(e) and 5(f)]. The bleach recovery traces were analyzed with biexponential kinetic fit and kinetic parameters are summarized in Table I (see supplementary material Tables S2 and S3 for short timescale and Tables S4 and S5 for long timescale).

FIG. 5.

Transient absorption spectra of AgInS2 suspension in toluene/ethanol (87:13% v/v): (a) without EV2+ and (b) with 4 µM EV2+. The transient absorption spectra were recorded following a 400 nm laser pulse (12 µJ/cm2) excitation. Bleach recovery kinetic traces of the (c) AgInS2 and (d) AgInS2–CdS at ∼580 nm in the absence and increasing concentrations of EV2+. Corresponding recovery kinetics at longer timescale for (e) AgInS2 and (f) AgInS2–CdS at ∼580 nm.

FIG. 5.

Transient absorption spectra of AgInS2 suspension in toluene/ethanol (87:13% v/v): (a) without EV2+ and (b) with 4 µM EV2+. The transient absorption spectra were recorded following a 400 nm laser pulse (12 µJ/cm2) excitation. Bleach recovery kinetic traces of the (c) AgInS2 and (d) AgInS2–CdS at ∼580 nm in the absence and increasing concentrations of EV2+. Corresponding recovery kinetics at longer timescale for (e) AgInS2 and (f) AgInS2–CdS at ∼580 nm.

Close modal
TABLE I.

Excited state and electron transfer properties of AgInS2 QD and AgInS2–CdS heterostructure.

ParameterAgInS2AgInS2–CdS
Emission quantum yield 0.30 0.37 
Kapp [equilibrium (3)] (M−16.2 × 106 2.0 × 106 
Avg. emission lifetime (ns) 493 300 
Forward electron transfer, ket (fast) (s−10.3–2.0 × 1011 0.2–1.3 × 1011 
ket (slow) (s−10.1–3.6 × 108 0.08–1.6 × 108 
Back electron transfer, kbet (s−12.6 × 105 1.5 × 105 
Electron transfer (EV) quantum yield (λ > 420 nm) 0.15 0.29 
(λ > 515 nm) 0.04 0.11 
(λ > 550 nm) 0.04 0.04 
ParameterAgInS2AgInS2–CdS
Emission quantum yield 0.30 0.37 
Kapp [equilibrium (3)] (M−16.2 × 106 2.0 × 106 
Avg. emission lifetime (ns) 493 300 
Forward electron transfer, ket (fast) (s−10.3–2.0 × 1011 0.2–1.3 × 1011 
ket (slow) (s−10.1–3.6 × 108 0.08–1.6 × 108 
Back electron transfer, kbet (s−12.6 × 105 1.5 × 105 
Electron transfer (EV) quantum yield (λ > 420 nm) 0.15 0.29 
(λ > 515 nm) 0.04 0.11 
(λ > 550 nm) 0.04 0.04 

In the absence of EV2+, we observe a relatively long lifetime for the recovery of the bleach for both AgInS2 and AgInS2–CdS nanocrystals. The average lifetime estimated for AgInS2 and AgInS2–CdS is 780 and 502 ns, respectively. (These lifetimes are slightly greater than the emission lifetime as the transient absorption represents various deactivation pathways.) As discussed earlier,20–25 the surface and the interstitial trap states extend the lifetime of charge carriers.24,25 The CdS shell remediates some of the surface defects that are responsible for long-lived recovery as evident from the decreased lifetime of recovery.

The traces recorded in Figs. 5(c) and 5(d) show faster bleach recovery of AgInS2 and AgInS2–CdS with increasing concentration of EV2+. More than 70% of the recovery occurs within 50 ps at the highest EV2+ concentrations. The short lifetime of bleach recovery decreases from 64.7 to 6.3 ns for AgInS2 suspension containing 8 μM EV2+ and 110.0 to 7.3 ns for AgInS2–CdS suspension containing 8 μM EV2+ suggesting rapid electron transfer in both of these systems. If we consider the observed decrease is entirely due to the electron transfer process [reaction (2)], we can obtain rate constant for electron transfer (ket) using the following expression:

ket=1/τ1/τ0,
(6)

where τ0 and τ′ are the fast processes lifetimes of bleaching recovery in the absence and presence of an electron acceptor.

The rate constants determined at 8 μM EV2+ were 2.0× 1011 and 1.3 × 1011 s−1 for AgInS2 and AgInS2–CdS, respectively, and represent electron transfer from conduction band electrons. The CdS shell around AgInS2 appears to slow down the electron transfer process. As noted in our earlier work,53,54 the surface interaction plays an important role in dictating the rate and efficiency of interfacial electron transfer. As seen in the case of emission quenching measurements, the lower association constant observed for EV2+ with AgInS2–CdS would also influence the rate of electron transfer. Strong interaction between viologen (MV2+) and CsPbBr3 nanocrystals has also been shown to exhibit extended charge separation with electrons residing in the viologen moiety and holes residing in the semiconductor nanocrystals.64 Given the long lifetime of the excited state of AgInS2, we would expect a long-lived charge separation with the stabilization of electron transfer product.

The residual bleach at the end of the initial fast recovery was tracked further at longer times [Figs. 5(e) and 5(f)]. The long-term bleach exhibited faster recovery in the presence of EV2+. This faster recovery over a long time scale arises from the transfer of trapped electrons to EV2+. It should be noted that the long-term bleach that arises from trapped charge carriers is only a small fraction, since nearly 90% of the recovery is seen within 200 ps (see Figs. S4 and S5). The rate constants from the longer time kinetics at 8 μM EV2+ were 3.6 × 108 and 1.6 × 108 s−1 for AgInS2 and AgInS2–CdS, respectively. Although the contribution of trapped electrons to EV2+ is small, it highlights the ability to participate in the interfacial electron transfer process. This observation is also in agreement with the steady state photolysis experiments in which we observe EV formation under sub-bandgap excitations [Fig. 4(f)]. The photophysical and electron transfer properties of AgInS2 and AgInS2–CdS are compared in Table I.

Despite the faster bleach recovery seen in the presence of EV2+, the residual bleach signal in the 500–700 nm region in the previous measurements prevented us from characterizing the electron transfer product (EV). We employed a nanosecond laser flash photolysis to probe the formation of EV by using 355 nm laser pulse excitation (pulse width 10 ns). The transient absorption spectra recorded with AgInS2 and AgInS2–CdS are shown in Figs. 6(a) and 6(c), respectively. The formation of the electron transfer product with its characteristic absorption of around 600 nm could be seen immediately after the laser flash excitation. A fraction of this transient absorption decays quickly (within ∼15 µs) indicating contribution from the back electron transfer. The transient absorption decay profiles [Figs. 6(b) and 6(d)] show the initial decay followed by the stabilization of the absorbance signal.

FIG. 6.

Transient absorption spectra recorded following the 355 nm laser pulse excitation of (a) AgInS2 and (c) AgInS2–CdS dispersed in toluene/ethanol (87:13% v/v) in the presence of 50 µM EV2+. Decay kinetics traces of EV for (b) AgInS2 and (d) AgInS2–CdS recorded at 610 nm.

FIG. 6.

Transient absorption spectra recorded following the 355 nm laser pulse excitation of (a) AgInS2 and (c) AgInS2–CdS dispersed in toluene/ethanol (87:13% v/v) in the presence of 50 µM EV2+. Decay kinetics traces of EV for (b) AgInS2 and (d) AgInS2–CdS recorded at 610 nm.

Close modal

It is interesting to note that the initial decay component is significantly smaller in AgInS2–CdS heterostructures, indicating smaller loss of the electron transfer product in the back electron transfer process. The initial decay observed at 600 nm can be attributed to back electron transfer between trapped holes and EV species,

AgInS2(ht)+EV+AgInS2+EV2+.
(7)

In the case of AgInS2–CdS, the CdS shell acts as a barrier to suppress the back electron transfer between EV2+ and trapped holes (Scheme 2). The fraction of residual bleach [ΔA(30 µs)/ΔA(0 µs)] in the kinetic traces in Figs. 6(c) and 6(d) corresponds to stabilization of electron transfer product. It is interesting to note that the residual bleach fraction observed for AgInS2–CdS is nearly twice that observed for AgInS2. This observation parallels the increased quantum yield of EV observed in steady state irradiation experiments with AgInS2–CdS nanocrystals (29.3%) as compared to pristine AgInS2 nanocrystals (14.5%).

SCHEME 2.

Involvement of CdS in modulating electron transfer to and from EV2+. Prevalence of back electron transfer characterizing AgInS2 core is significantly reduced in AgInS2–CdS.

SCHEME 2.

Involvement of CdS in modulating electron transfer to and from EV2+. Prevalence of back electron transfer characterizing AgInS2 core is significantly reduced in AgInS2–CdS.

Close modal

The results presented here shed light on two important issues. The first one is related to new mechanistic insights into the electron transfer processes of ternary semiconductors. Both bandgap and sub-bandgap excitations induce electron transfer as probed through EV formation under wavelength selective irradiation. The quantum yields obtained with sub-bandgap excitation (λ > 550 nm) are relatively lower compared to broadband white light (λ > 420 nm) irradiation. The transfer of conduction band electrons to EV2+ is fast, which occurs with a rate constant of 0.6–2 × 1011 s−1 and constitutes a major part of interfacial electron transfer. The contribution of trapped electrons from the interstitial defects occurs at a slower rate (0.5–2.8 × 108 s−1). The second aspect is the stabilization of electron transfer product. Both steady state irradiation and transient absorption measurements indicate suppression of back electron transfer in AgInS2–CdS heterostructures. The beneficial role that the CdS layer plays results in enhancement in the electron transfer quantum yield from 14.5% to 29.3%. As evident from the analysis of electron transfer kinetics, the interfacial electron transfer in ternary semiconductors is a kinetically driven process. Careful control of surface interaction of the acceptor molecule and suppression of back electron transfer is needed to maximize the efficiency of photocatalytic reduction. The heterostructures such as AgInS2–CdS can provide new design strategies to amplify the use of ternary semiconductors in photocatalysis.

The supplementary material includes experimental methods and procedures, actinometry, data of control experiments, and kinetic analysis.

The research described herein is supported by the Division of Chemical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciences of the U.S. Department of Energy, through Award No. DE-FC02-04ER15533. This contribution is NDRL No. 5354 from the Notre Dame Radiation Laboratory. We also acknowledge the University of Notre Dame Equipment Restoration and Renewal (ERR) program for the purchase of the Spectra Physics laser used for the transient absorption measurements.

The authors have no conflicts to disclose.

A.K. contributed to conceptualization, synthesis, measurements, data analysis, validation, visualization, and writing of the first draft and revision. P.V.K. contributed to conceptualization, method development, funding acquisition, resources, supervision, discussions, and writing.

The data that support the findings of this study are available within the article and its supplementary material.

1.
W.-S.
Song
and
H.
Yang
, “
Efficient white-light-emitting diodes fabricated from highly fluorescent copper indium sulfide core/shell quantum dots
,”
Chem. Mater.
24
(
10
),
1961
1967
(
2012
).
2.
P.-H.
Chuang
,
C. C.
Lin
, and
R.-S.
Liu
, “
Emission-tunable CuInS2/ZnS quantum dots: Structure, optical properties, and application in white light-emitting diodes with high color rendering index
,”
ACS Appl. Mater. Interfaces
6
(
17
),
15379
15387
(
2014
).
3.
B.
Chen
,
H.
Zhong
,
W.
Zhang
,
Z. A.
Tan
,
Y.
Li
,
C.
Yu
,
T.
Zhai
,
Y.
Bando
,
S.
Yang
, and
B.
Zou
, “
Highly emissive and color-tunable CuInS2-based colloidal semiconductor nanocrystals: Off-stoichiometry effects and improved electroluminescence performance
,”
Adv. Funct. Mater.
22
(
10
),
2081
2088
(
2012
).
4.
Z.
Tan
,
Y.
Zhang
,
C.
Xie
,
H.
Su
,
J.
Liu
,
C.
Zhang
,
N.
Dellas
,
S. E.
Mohney
,
Y.
Wang
,
J.
Wang
, and
J.
Xu
, “
Near-band-edge electroluminescence from heavy-metal-free colloidal quantum dots
,”
Adv. Mater.
23
(
31
),
3553
3558
(
2011
).
5.
G.
Zaiats
,
S.
Ikeda
, and
P. V.
Kamat
, “
Optimization of the electron transport layer in quantum dot light-emitting devices
,”
NPG Asia Mater.
12
(
1
),
57
(
2020
).
6.
T.
Uematsu
,
A.
Doko
,
T.
Torimoto
,
K.
Oohora
,
T.
Hayashi
, and
S.
Kuwabata
, “
Photoinduced electron transfer of ZnS–AgInS2 solid-solution semiconductor nanoparticles: Emission quenching and photocatalytic reactions controlled by electrostatic forces
,”
J. Phys. Chem. C
117
(
30
),
15667
15676
(
2013
).
7.
R.
Miranti
,
Y.
Firdaus
,
E.
Fron
,
M.
Van der Auweraer
,
J.
Parisi
, and
H.
Borchert
, “
Photoinduced charge transfer in hybrid systems of CuInS2 nanocrystals and conductive polymer
,”
J. Phys. Chem. C
119
(
39
),
22669
22680
(
2015
).
8.
J.
Ann Maria Xavier
,
G.
Devatha
,
S.
Roy
,
A.
Rao
, and
P. P.
Pillai
, “
Electrostatically regulated photoinduced electron transfer in ‘cationic’ eco-friendly CuInS2/ZnS quantum dots in water
,”
J. Mater. Chem. A
6
(
44
),
22248
22255
(
2018
).
9.
D.
Yan
,
S.
Huang
,
Y.
Von Lim
,
D.
Fang
,
Y.
Shang
,
M. E.
Pam
,
J.
Yu
,
D.
Xiong
,
X. L.
Li
,
J.
Zhang
,
Y.
Wang
,
L.
Pan
,
Y.
Bai
,
Y.
Shi
, and
H. Y.
Yang
, “
Stepwise intercalation-conversion intercalation sodiation mechanism in CuInS2 prompting sodium storage performance
,”
ACS Energy Lett.
5
(
12
),
3725
3732
(
2020
).
10.
M.
Bernechea
,
N. C.
Miller
,
G.
Xercavins
,
D.
So
,
A.
Stavrinadis
, and
G.
Konstantatos
, “
Solution-processed solar cells based on environmentally friendly AgBiS2 nanocrystals
,”
Nat. Photonics
10
(
8
),
521
525
(
2016
).
11.
H.
Zhong
,
Z.
Bai
, and
B.
Zou
, “
Tuning the luminescence properties of colloidal I–III–VI semiconductor nanocrystals for optoelectronics and biotechnology applications
,”
J. Phys. Chem. Lett.
3
(
21
),
3167
3175
(
2012
).
12.
T.-L.
Li
,
Y.-L.
Lee
, and
H.
Teng
, “
High-performance quantum dot-sensitized solar cells based on sensitization with CuInS2 quantum dots/CdS heterostructure
,”
Energy Environ. Sci.
5
(
1
),
5315
5324
(
2012
).
13.
D. H.
Jara
,
S. J.
Yoon
,
K. G.
Stamplecoskie
, and
P. V.
Kamat
, “
Size-dependent photovoltaic performance of CuInS2 quantum dot-sensitized solar cells
,”
Chem. Mater.
26
(
24
),
7221
7228
(
2014
).
14.
P. K.
Santra
,
P. V.
Nair
,
K.
George Thomas
, and
P. V.
Kamat
, “
CuInS2-sensitized quantum dot solar cell. Electrophoretic deposition, excited-state dynamics, and photovoltaic performance
,”
J. Phys. Chem. Lett.
4
(
5
),
722
729
(
2013
).
15.
C.-C.
Chang
,
J.-K.
Chen
,
C.-P.
Chen
,
C.-H.
Yang
, and
J.-Y.
Chang
, “
Synthesis of eco-friendly CuInS2 quantum dot-sensitized solar cells by a combined ex situ/in situ growth approach
,”
ACS Appl. Mater. Interfaces
5
(
21
),
11296
11306
(
2013
).
16.
N.
Pradhan
and
S.
Efrima
, “
Single-precursor, one-pot versatile synthesis under near ambient conditions of tunable, single and dual band fluorescing metal sulfide nanoparticles
,”
J. Am. Chem. Soc.
125
(
8
),
2050
2051
(
2003
).
17.
N.
Gaponik
,
D. V.
Talapin
,
A. L.
Rogach
,
K.
Hoppe
,
E. V.
Shevchenko
,
A.
Kornowski
,
A.
Eychmüller
, and
H.
Weller
, “
Thiol-capping of CdTe nanocrystals: An alternative to organometallic synthetic routes
,”
J. Phys. Chem. B
106
(
29
),
7177
7185
(
2002
).
18.
C.-H.
Chuang
,
S. S.
Lo
,
G. D.
Scholes
, and
C.
Burda
, “
Charge separation and recombination in CdTe/CdSe core/shell nanocrystals as a function of shell coverage: Probing the onset of the quasi type-II regime
,”
J. Phys. Chem. Lett.
1
(
17
),
2530
2535
(
2010
).
19.
R.
Bose
,
A.
Bera
,
M. R.
Parida
,
A.
Adhikari
,
B. S.
Shaheen
,
E.
Alarousu
,
J.
Sun
,
T.
Wu
,
O. M.
Bakr
, and
O. F.
Mohammed
, “
Real-space mapping of surface trap states in CIGSe nanocrystals using 4D electron microscopy
,”
Nano Lett.
16
(
7
),
4417
4423
(
2016
).
20.
T.
Kameyama
,
T.
Takahashi
,
T.
Machida
,
Y.
Kamiya
,
T.
Yamamoto
,
S.
Kuwabata
, and
T.
Torimoto
, “
Controlling the electronic energy structure of ZnS–AgInS2 solid solution nanocrystals for photoluminescence and photocatalytic hydrogen evolution
,”
J. Phys. Chem. C
119
(
44
),
24740
24749
(
2015
).
21.
X.
Tang
,
W. B. A.
Ho
, and
J. M.
Xue
, “
Synthesis of Zn-doped AgInS2 nanocrystals and their fluorescence properties
,”
J. Phys. Chem. C
116
(
17
),
9769
9773
(
2012
).
22.
B.
Mao
,
C.-H.
Chuang
,
J.
Wang
, and
C.
Burda
, “
Synthesis and photophysical properties of ternary I–III–VI AgInS2 nanocrystals: Intrinsic versus surface states
,”
J. Phys. Chem. C
115
(
18
),
8945
8954
(
2011
).
23.
T.
Torimoto
,
T.
Adachi
,
K.-i.
Okazaki
,
M.
Sakuraoka
,
T.
Shibayama
,
B.
Ohtani
,
A.
Kudo
, and
S.
Kuwabata
, “
Facile synthesis of ZnS–AgInS2 solid solution nanoparticles for a color-adjustable luminophore
,”
J. Am. Chem. Soc.
129
(
41
),
12388
12389
(
2007
).
24.
Y.
Hamanaka
,
T.
Ogawa
,
M.
Tsuzuki
, and
T.
Kuzuya
, “
Photoluminescence properties and its origin of AgInS2 quantum dots with chalcopyrite structure
,”
J. Phys. Chem. C
115
(
5
),
1786
1792
(
2011
).
25.
M. J.
Rao
,
T.
Shibata
,
S.
Chattopadhyay
, and
A.
Nag
, “
Origin of photoluminescence and XAFS study of (ZnS)1−x(AgInS2)x nanocrystals
,”
J. Phys. Chem. Lett.
5
(
1
),
167
173
(
2014
).
26.
S. L.
Castro
,
S. G.
Bailey
,
R. P.
Raffaelle
,
K. K.
Banger
, and
A. F.
Hepp
, “
Synthesis and characterization of colloidal CuInS2 nanoparticles from a molecular single-source precursor
,”
J. Phys. Chem. B
108
(
33
),
12429
12435
(
2004
).
27.
X.
Liu
,
X.
Peng
,
T.
Xu
,
Y.
Li
,
H.
Wei
,
Z.
Guan
, and
F.
Wang
, “
Broad photoluminescence in ternary Ag-In-S and quaternary Ag-In-Zn-S nanoparticles: The role of Zn incorporation
,”
Mater. Res. Express
8
(
8
),
085002
(
2021
).
28.
Y.
Hamanaka
,
T.
Ogawa
,
M.
Tsuzuki
,
K.
Ozawa
, and
T.
Kuzuya
, “
Luminescence properties of chalcopyrite AgInS2 nanocrystals: Their origin and related electronic states
,”
J. Lumin.
133
,
121
124
(
2013
).
29.
V.
Pinchetti
,
M.
Lorenzon
,
H.
McDaniel
,
R.
Lorenzi
,
F.
Meinardi
,
V. I.
Klimov
, and
S.
Brovelli
, “
Spectro-electrochemical probing of intrinsic and extrinsic processes in exciton recombination in I–III–VI2 nanocrystals
,”
Nano Lett.
17
(
7
),
4508
4517
(
2017
).
30.
Y.
Hamanaka
,
K.
Ozawa
, and
T.
Kuzuya
, “
Enhancement of donor–acceptor pair emissions in colloidal AgInS2 quantum dots with high concentrations of defects
,”
J. Phys. Chem. C
118
(
26
),
14562
14568
(
2014
).
31.
F.
Zhao
,
Q.
Li
,
K.
Han
, and
T.
Lian
, “
Mechanism of efficient viologen radical generation by ultrafast electron transfer from CdS quantum dots
,”
J. Phys. Chem. C
122
(
30
),
17136
17142
(
2018
).
32.
H.
Zhu
,
N.
Song
,
H.
Lv
,
C. L.
Hill
, and
T.
Lian
, “
Near unity quantum yield of light-driven redox mediator reduction and efficient H2 generation using colloidal nanorod heterostructures
,”
J. Am. Chem. Soc.
134
(
28
),
11701
11708
(
2012
).
33.
H.
Zang
,
H.
Li
,
N. S.
Makarov
,
K. A.
Velizhanin
,
K.
Wu
,
Y.-S.
Park
, and
V. I.
Klimov
, “
Thick-shell CuInS2/ZnS quantum dots with suppressed ‘blinking’ and narrow single-particle emission line widths
,”
Nano Lett.
17
(
3
),
1787
1795
(
2017
).
34.
T.
Takahashi
,
A.
Kudo
,
S.
Kuwabata
,
A.
Ishikawa
,
H.
Ishihara
,
Y.
Tsuboi
, and
T.
Torimoto
, “
Plasmon-enhanced photoluminescence and photocatalytic activities of visible-light-responsive ZnS-AgInS2 solid solution nanoparticles
,”
J. Phys. Chem. C
117
(
6
),
2511
2520
(
2013
).
35.
Y. J.
Park
,
J. H.
Oh
,
N. S.
Han
,
H. C.
Yoon
,
S. M.
Park
,
Y. R.
Do
, and
J. K.
Song
, “
Photoluminescence of band gap states in AgInS2 nanoparticles
,”
J. Phys. Chem. C
118
(
44
),
25677
25683
(
2014
).
36.
S. M.
Kobosko
and
P. V.
Kamat
, “
Indium-rich AgInS2–ZnS quantum dots—Ag-/Zn-dependent photophysics and photovoltaics
,”
J. Phys. Chem. C
122
(
26
),
14336
14344
(
2018
).
37.
T.
Chevallier
,
G.
Le Blevennec
, and
F.
Chandezon
, “
Photoluminescence properties of AgInS2–Zns nanocrystals: The critical role of the surface
,”
Nanoscale
8
(
14
),
7612
7620
(
2016
).
38.
K. V.
Yumashev
,
A. M.
Malyarevich
,
P. V.
Prokoshin
,
N. N.
Posnov
,
V. P.
Mikhailov
,
V. S.
Gurin
, and
M. V.
Artemyev
, “
Optical transient bleaching/absorption of surface-oxidized CuInS2 nanocrystals
,”
Appl. Phys. B
65
(
4
),
545
548
(
1997
).
39.
R. R.
Schnepf
,
J. J.
Cordell
,
M. B.
Tellekamp
,
C. L.
Melamed
,
A. L.
Greenaway
,
A.
Mis
,
G. L.
Brennecka
,
S.
Christensen
,
G. J.
Tucker
,
E. S.
Toberer
,
S.
Lany
, and
A. C.
Tamboli
, “
Utilizing site disorder in the development of new energy-relevant semiconductors
,”
ACS Energy Lett.
5
(
6
),
2027
2041
(
2020
).
40.
D. H.
Jara
,
K. G.
Stamplecoskie
, and
P. V.
Kamat
, “
Two distinct transitions in CuxInS2 quantum dots. Bandgap versus sub-bandgap excitations in copper-deficient structures
,”
J. Phys. Chem. Lett.
7
(
8
),
1452
1459
(
2016
).
41.
K.
Hattori
,
K.
Akamatsu
, and
N.
Kamegashira
, “
Electrical properties of polycrystalline chalcopyrite AgInS2 films
,”
J. Appl. Phys.
71
(
7
),
3414
(
1992
).
42.
S. H.
You
,
K. J.
Hong
,
C. J.
Youn
,
T. S.
Jeong
,
J. D.
Moon
,
H. S.
Kim
, and
J. S.
Park
, “
Origin of point defects in AgInS2/GaAs epilayer obtained from photoluminescence measurement
,”
J. Appl. Phys.
90
(
8
),
3894
(
2001
).
43.
T.
Torimoto
,
Y.
Kamiya
,
T.
Kameyama
,
H.
Nishi
,
T.
Uematsu
,
S.
Kuwabata
, and
T.
Shibayama
, “
Controlling shape anisotropy of ZnS–AgInS2 solid solution nanoparticles for improving photocatalytic activity
,”
ACS Appl. Mater. Interfaces
8
(
40
),
27151
27161
(
2016
).
44.
O.
Stroyuk
,
A.
Raevskaya
,
F.
Spranger
,
O.
Selyshchev
,
V.
Dzhagan
,
S.
Schulze
,
D. R. T.
Zahn
, and
A.
Eychmüller
, “
Origin and dynamics of highly efficient broadband photoluminescence of aqueous glutathione-capped size-selected Ag–In–S quantum dots
,”
J. Phys. Chem. C
122
(
25
),
13648
13658
(
2018
).
45.
K. E.
Knowles
,
H. D.
Nelson
,
T. B.
Kilburn
, and
D. R.
Gamelin
, “
Singlet–triplet splittings in the luminescent excited states of colloidal Cu+:CdSe, Cu+:InP, and CuInS2 nanocrystals: Charge-transfer configurations and self-trapped excitons
,”
J. Am. Chem. Soc.
137
(
40
),
13138
13147
(
2015
).
46.
Y.
Han
,
S.
He
,
X.
Luo
,
Y.
Li
,
Z.
Chen
,
W.
Kang
,
X.
Wang
, and
K.
Wu
, “
Triplet sensitization by ‘self-trapped’ excitons of nontoxic CuInS2 nanocrystals for efficient photon upconversion
,”
J. Am. Chem. Soc.
141
(
33
),
13033
13037
(
2019
).
47.
K. R.
Gopidas
and
P. V.
Kamat
, “
Photochemistry on surfaces. 4. Influence of support material on the photochemistry of an adsorbed dye
,”
J. Phys. Chem.
93
(
17
),
6428
6433
(
1989
).
48.
J. Z.
Zhang
, “
Interfacial charge carrier dynamics of colloidal semiconductor nanoparticles
,”
J. Phys. Chem. B
104
(
31
),
7239
7253
(
2000
).
49.
A.
Hagfeldt
and
M.
Grätzel
, “
Light-induced redox reactions in nanocrystalline systems
,”
Chem. Rev.
95
(
1
),
49
68
(
1995
).
50.
J. B.
Asbury
,
E.
Hao
,
Y.
Wang
,
H. N.
Ghosh
, and
T.
Lian
, “
Ultrafast electron transfer dynamics from molecular adsorbates to semiconductor nanocrystalline thin films
,”
J. Phys. Chem. B
105
(
20
),
4545
4557
(
2001
).
51.
T.
Shishido
,
K.
Teramura
, and
T.
Tanaka
, “
Photo-induced electron transfer between a reactant molecule and semiconductor photocatalyst: In situ doping
,”
Catal. Surv. Asia
15
(
4
),
240
258
(
2011
).
52.
A. J.
Morris-Cohen
,
M. T.
Frederick
,
L. C.
Cass
, and
E. A.
Weiss
, “
Simultaneous determination of the adsorption constant and the photoinduced electron transfer rate for a CdS quantum dot–viologen complex
,”
J. Am. Chem. Soc.
133
(
26
),
10146
10154
(
2011
).
53.
A.
Kipkorir
,
J.
DuBose
,
J.
Cho
, and
P. V.
Kamat
, “
CsPbBr3–CdS heterostructure: Stabilizing perovskite nanocrystals for photocatalysis
,”
Chem. Sci.
12
(
44
),
14815
14825
(
2021
).
54.
J. T.
DuBose
and
P. V.
Kamat
, “
Surface chemistry matters. How ligands influence excited state interactions between CsPbBr3 and methyl viologen
,”
J. Phys. Chem. C
124
(
24
),
12990
12998
(
2020
).
55.
P. V.
Kamat
,
J. P.
Chauvet
, and
R. W.
Fessenden
, “
Photoelectrochemistry in particulate systems. 4. Photosensitization of a TiO2 semiconductor with a chlorophyll analog
,”
J. Phys. Chem.
90
(
7
),
1389
1394
(
1986
).
56.
J. T.
DuBose
and
P. V.
Kamat
, “
Probing perovskite photocatalysis. Interfacial electron transfer between CsPbBr3 and ferrocene redox couple
,”
J. Phys. Chem. Lett.
10
(
20
),
6074
6080
(
2019
).
57.
R.
Yanagi
,
T. S.
Zhao
,
D.
Solanki
,
Z. H.
Pan
, and
S.
Hu
, “
Charge separation in photocatalysts: Mechanisms, physical parameters, and design principles
,”
ACS Energy Lett.
7
,
432
(
2021
).
58.
C. L.
Bird
and
A. T.
Kuhn
, “
Electrochemistry of the viologens
,”
Chem. Soc. Rev.
10
(
1
),
49
82
(
1981
).
59.
L.
Michaelis
and
E. S.
Hill
, “
The viologen indicators
,”
J. Gen. Physiol.
16
(
6
),
859
873
(
1933
).
60.
A.
Lagatti
,
L.
Tarpani
,
E.
Fiacchi
,
L.
Bussotti
,
L.
Latterini
, and
P.
Foggi
, “
Charge transfer dynamics between MPA capped CdTe quantum dots and methyl viologen
,”
J. Photochem. Photobiol., A
346
,
382
389
(
2017
).
61.
M.
Chhetri
and
P. V.
Kamat
, “
Vectorial charge transfer across bipolar membrane loaded with CdS and Au nanoparticles
,”
J. Phys. Chem. C
125
(
12
),
6870
6876
(
2021
).
62.
C. G.
Hatchard
and
C. A.
Parker
, “
A new sensitive chemical actinometer—II. Potassium ferrioxalate as a standard chemical actinometer
,”
Proc. R. Soc. A: Math. Phys.
235
(
1203
),
518
536
(
1956
).
63.
V. L.
Bridewell
,
R.
Alam
,
C. J.
Karwacki
, and
P. V.
Kamat
, “
CdSe/CdS nanorod photocatalysts: Tuning the interfacial charge transfer process through shell length
,”
Chem. Mater.
27
(
14
),
5064
5071
(
2015
).
64.
S. M.
Kobosko
,
J. T.
DuBose
, and
P. V.
Kamat
, “
Perovskite photocatalysis. Methyl viologen induces unusually long-lived charge carrier separation in CsPbBr3 nanocrystals
,”
ACS Energy Lett.
5
(
1
),
221
223
(
2020
).

Supplementary Material