The successful design and device integration of nanoscale heterointerfaces hinges upon precise manipulation of both ground- and excited-state charge carrier (electron and hole) densities. However, it is particularly challenging to quantify these charge carrier densities in nanoscale materials, leading to uncertainties in the mechanisms of many carrier density-dependent properties and processes. Here, we demonstrate a method that utilizes steady-state and transient absorption spectroscopies to correlate monolayer MoS2 electron density with the easily measured metric of excitonic optical absorption quenching in a variety of mixed-dimensionality s-SWCNT/MoS2 heterostructures. By employing a 2D phase-space filling model, the resulting correlation elucidates the relationship between charge density, local dielectric environment, and concomitant excitonic properties. The phase-space filling model is also able to describe existing trends from the literature on transistor-based measurements on MoS2, WS2, and MoSe2 monolayers that were not previously compared to a physical model, providing additional support for our method and results. The findings provide a pathway to the community for estimating both ground- and excited-state carrier densities in a wide range of TMDC-based systems.

Two-dimensional (2D) layered semiconductors, such as transition metal dichalcogenides (TMDCs), have been studied intensively for over forty years.1,2 The successful isolation and growth of monolayer TMDCs in the past ten years ushered in the development and studies of TMDC heterostructures for use in optoelectronic, catalytic, and quantum information processing applications,3–7 due to the emergence of a direct bandgap and other beneficial properties, such as spin–valley locking.8,9 Despite these advances, several fundamental knowledge gaps remain for monolayer TMDCs and their heterostructures. In general, the optimization of semiconductor heterostructures depends on the knowledge of both ground- and excited-state carrier densities, which dictate the steady-state and transient properties of optoelectronic devices. Quantitative assessment of charge carrier concentrations can be challenging for monolayer TMDCs, since traditional methods rely upon an accurate knowledge of properties, such as dielectric constant,10 effective mass or mobility,11 and device capacitance,12 all of which can be challenging to measure and are impacted in uncertain ways by the local environment.13 

Several decades ago, the III–V community developed a powerful, but simple, method based on optical absorption spectroscopy to quantify carrier densities and carrier-density dependent properties in GaAs quantum wells, an early 2D excitonic material. This model, based on the phase space filling effect, related the carrier density in the two-dimensional layer to the resulting quenching of the strong excitonic optical transitions.14,15 Recent studies have also applied similar models to semiconducting single-walled carbon nanotubes (s-SWCNTs) as a model one-dimensional (1D) excitonic semiconductor.16,17 A key to this phase-space filling model is the knowledge, or estimate, of the exciton size, since that determines the fraction of exciton oscillator strength that is quenched by a particular charge carrier density. In turn, this exciton size is influenced by the reduced exciton mass and the dielectric constant experienced by the exciton. Applying such a model to monolayer TMDCs is an attractive possibility but is made challenging by the large experimental and theoretical variations in reported dielectric constants (that can depend on sample substrate, thickness, and growth methods)10,18,19 and excitonic constants.11,20,21

Applying a phase-space filling model to TMDCs requires a means to systematically vary the monolayer carrier density, a series of experimentally measured absorption or reflection spectra at each carrier density, and a reliable way to quantify the carrier density for each spectrum. Carrier density can be modulated by the gate-voltage in a transistor, via molecular or substitutional ground-state doping, or via dynamic processes, such as photoinduced charge transfer. Some transistor-based studies have reported trends for the dependence of exciton oscillator strength, measured by reflectance, on carrier density in MoS2,22 WS2,23 and MoSe224 monolayers, but these studies have not simulated those results with a physically relevant model, such as a phase-space filling dependence. The carrier density in these studies is extracted by considering the gate oxide capacitance and the difference between the applied gate voltage and the gate voltage associated with the charge neutrality point. However, the carrier density is dictated by the total gate capacitance, equal to the sum of the gate oxide capacitance and the channel quantum capacitance. It has been pointed out that the monolayer TMDC quantum capacitance can dominate the total capacitance, leading to errors in extracting carrier density by using the gate oxide capacitance alone.12 With these considerations in mind, the need remains for a rigorous analysis aimed at quantifying carrier density in monolayer TMDCs by optical spectroscopy.

In this study, we quantify carrier density in monolayer MoS2 by using an “internal standard,” in this case a one-dimensional (1D) electron donor in a series of mixed-dimensionality (1D/2D) donor/acceptor heterostructures (see Fig. 1),25,26 and employ a 2D phase-space filling model to simulate the dependence of MoS2 exciton bleach on carrier density. This study expands upon the concept recently introduced by Sulas-Kern et al.,25 where the optical cross section for charges in a semiconducting single-walled carbon nanotube (s-SWCNT) thin film was used to quantify charge separation quantum yield in a photoexcited s-SWCNT/MoS2 heterostructure. That proof-of-concept demonstration was limited in that it only utilized a single value of excited-state charge yield and only a single study at the time27 had estimated a charge carrier (hole) cross section in the highly enriched (6,5) s-SWCNTs used as electron donors. Since that study, Eckstein et al. utilized spectro-electrochemistry and a 1D phase-space filling model to quantify hole density in (6,5) s-SWCNTs,16 and we have performed additional studies on mixed-dimensionality s-SWCNT/MoS2 heterostructures.26,28 Furthermore, an experimental diamagnetic shift study20 that came out shortly after Sulas-Kern et al.25 provided a rigorously determined reduced exciton mass for high-quality MoS2 that allows us to better constrain the 2D phase-space filling model.

FIG. 1.

Schematic of the method employed here to quantify the impact of charge carrier (electron) density on MoS2 A exciton quenching (ground-state bleach or GSB). Pump–probe transient absorbance spectroscopy follows the photoinduced charge transfer in a photoexcited bilayer (example shown) or trilayer containing MoS2 and a (6,5) s-SWCNT thin film (the internal standard). Following photoinduced charge transfer, the following steps are used to quantify this dependence: (1) The known absorption cross sections of (6,5) s-SWCNT charge-related TA features are used to calculate the s-SWCNT hole density (nh). (2) Based on charge balance, the electron density in MoS2 (ne) is equivalent to nh. (3) The GSB magnitude in MoS2 can then be correlated directly to ne.

FIG. 1.

Schematic of the method employed here to quantify the impact of charge carrier (electron) density on MoS2 A exciton quenching (ground-state bleach or GSB). Pump–probe transient absorbance spectroscopy follows the photoinduced charge transfer in a photoexcited bilayer (example shown) or trilayer containing MoS2 and a (6,5) s-SWCNT thin film (the internal standard). Following photoinduced charge transfer, the following steps are used to quantify this dependence: (1) The known absorption cross sections of (6,5) s-SWCNT charge-related TA features are used to calculate the s-SWCNT hole density (nh). (2) Based on charge balance, the electron density in MoS2 (ne) is equivalent to nh. (3) The GSB magnitude in MoS2 can then be correlated directly to ne.

Close modal

With two independent estimates now available for (6,5) s-SWCNTs,16,27 we use this s-SWCNT “internal standard” to rigorously correlate MoS2 electron yield and exciton quenching in a variety of photoexcited s-SWCNT/MoS2 heterostructures. This method allows us to greatly constrain the potential range of physical properties used in a 2D phase-space filling model for the MoS2 component of the heterostructure. The resulting correlation between electron density and MoS2 carrier-induced exciton quenching can be used by the community to estimate both ground- and excited-state carrier densities in a wide range of MoS2-based systems, such as photoexcited heterostructures, (photo)transistors, and redox-doped or electrochemically doped monolayers. Our results also provide well-constrained ranges for fundamental MoS2 monolayer properties, such as exciton size and dielectric constant.

Figure 2(a) displays the representative steady-state absorption spectra of the mixed-dimensionality heterostructures probed in this study. Both types of heterostructures employ highly enriched (6,5) s-SWCNTs as a photoexcited electron donor and MoS2 as a photoexcited electron acceptor. The first heterostructure is a s-SWCNT/MoS2 bilayer,25,26 and the second is a WSe2/s-SWCNT/MoS2 trilayer that forms a charge transfer cascade.28 Strong excitonic absorption features related to both the (6,5) s-SWCNTs and TMDC components are labeled in the figure.

FIG. 2.

(a) Steady-state and (b) transient absorbance spectra of a neat (6,5) s-SWCNT (black), a s-SWCNT/MoS2 bilayer (orange), and a WSe2/s-SWCNT/MoS2 trilayer (blue). The labels identify salient excitonic absorption features for the s-SWCNTs (S11 and S22) and TMDCs (A and B excitons), along with the s-SWCNT trion induced absorbance (X+). (c) Exemplary spectral fits (Voigt line shapes) applied to the MoS2 steady-state (top panel) and transient (bottom panel) absorbance spectra to extract A0 and ΔA, respectively. These values are used to quantify the extent of exciton bleach (ΔA/A0) for a particular photoinduced carrier density.

FIG. 2.

(a) Steady-state and (b) transient absorbance spectra of a neat (6,5) s-SWCNT (black), a s-SWCNT/MoS2 bilayer (orange), and a WSe2/s-SWCNT/MoS2 trilayer (blue). The labels identify salient excitonic absorption features for the s-SWCNTs (S11 and S22) and TMDCs (A and B excitons), along with the s-SWCNT trion induced absorbance (X+). (c) Exemplary spectral fits (Voigt line shapes) applied to the MoS2 steady-state (top panel) and transient (bottom panel) absorbance spectra to extract A0 and ΔA, respectively. These values are used to quantify the extent of exciton bleach (ΔA/A0) for a particular photoinduced carrier density.

Close modal

Our study employs femtosecond pump–probe transient absorption (TA) spectroscopy to follow photoinduced charge transfer in these heterostructures. Specific excitonic and charge-related features in TMDCs and SWCNTs have revealed that SWCNTs aid in the spatial separation of charges in these TMDC–SWCNT based heterostructures, resulting in exceptionally long (>μs) charge-separated lifetimes.25,26,28 In this study, we use the magnitude of these transient spectral features following the completion of photoinduced charge transfer to quantify charge carrier quantum yield and ultimately to develop the phase-space filling “calibration curve” for monolayer MoS2 (Fig. 1). A judicious choice of the pump photon energy is employed to predominantly excite the s-SWCNT or TMDC components of the heterostructures.

Figure 2(b) displays the TA spectra, taken at 5 ps pump–probe delay, for a representative s-SWCNT thin film, a s-SWCNT/MoS2 bilayer, and a WSe2/s-SWCNT/MoS2 trilayer. TA features in the visible range can be assigned to the ground-state bleach (GSB) of the MoS2 and WSe2 excitonic transitions and the GSB of the (6,5) S22 excitonic transition. Features in the near-infrared (NIR) range can be assigned to the (6,5) S11 GSB and the trion induced absorption (IA), a feature that is only present when separated charge carriers (electrons or holes) are present on the s-SWCNTs. These distinct spectral features can be used to quantitatively track the temporal evolution of exciton dissociation, charge diffusion, and charge recombination in the mixed-dimensionality heterostructures.25,26,28

The charge transfer quantum yield (ϕCT) is defined as the number of separated charges (i.e., holes and electrons in separate materials) produced per photogenerated exciton. We estimate charge transfer yields for our type II heterostructures using two separate methods that utilize (1) an empirically determined trion absorption cross section that uses a heuristic model of the dependence of exciton bleaching on redox-doping induced carrier density27 and (2) a correlation between electrochemically modulated charge density and exciton bleach that utilizes a 1D phase-space filling model.16 Each method utilizes a similar overall methodology of tracking the dependence of the ground-state bleach (GSB) and trion IA on carrier density, but they utilize different frameworks for predicting the impact of the SWCNT density of states on this dependence. Until now, the two studies have not been compared to determine how well-matched their values may be for estimating SWCNT (or SWCNT heterostructure) carrier densities.

For each method described above, we first use singular value decomposition, accompanied by the associated kinetic rate equations, to simulate the full two-dimensional array of time-dependent TA spectra.25,26,28 These simulations produce concentration trajectories for the relevant species (e.g., excitons in photoexcited s-SWCNTs; charges in s-SWCNTs and TMDCs following exciton dissociation) evolving in time following the pump pulse. This method allows us to isolate and quantify charge-related s-SWCNT and MoS2 TA features.

Within the model proposed by Dowgiallo et al. (method 1 from above; supplementary material, Sec. 3.1), the intensity of the trion induced absorption can be normalized to the ground state S11 absorption (ODX(+/)/ODS11) to calculate the s-SWCNT carrier density. The trion optical density used to calculate s-SWCNT hole density derives from the TA spectrum associated with the concentration profiles where a given interfacial charge transfer reaction has reached completion (Figs. S1–S2). The s-SWCNT charge carrier concentration is then given by
(1)
where Nh/e is the hole or electron density/nm of tube length.
The second method (supplementary material, Sec. 3.2) relies on the 1D phase-space filling model developed by Eckstein et al. to correlate (6,5) s-SWCNT charge density to fractional exciton bleach.16 The fractional exciton bleach is defined as
(2)
where AS11 GSB and AS11 abs are the area under the curve for the S11 ground state bleach from the TA spectrum and S11 ground state absorption, respectively (see Figs. S3–S5 of the supplementary material).

Both these methods allow us to extract the s-SWCNT carrier density in units of nm−1. Since we are ultimately concerned with deriving an areal carrier density within the MoS2 layer, we convert this s-SWCNT carrier density to units of cm−2 by using the absorbance of the s-SWCNT film, the empirical absorbance coefficient for (6,5) s-SWCNTs, and the number of carbon atoms per nm in a (6,5) s-SWCNT (see Sec. 2 of the supplementary material). For s-SWCNT/MoS2 bilayers, the calculated s-SWCNT hole density equals the MoS2 electron density, since there is only one exciton dissociation interface and one exciton splits into one electron and one hole. The WSe2/s-SWCNT/MoS2 trilayer is more complex because excitons can be dissociated at two separate interfaces to produce electrons in MoS2 and holes in WSe2 [see Fig. 2(c) of the supplementary material]. Thus, the concentration trajectories for each species must be used to calculate the fraction of s-SWCNT holes produced by exciton dissociation at the s-SWCNT/MoS2 interface and the fraction of s-SWCNT electrons produced by exciton dissociation at the s-SWCNT/WSe2 interface (Figs. S1 and S2). The MoS2 electron density is then equivalent to the calculated s-SWCNT hole density.

Table I compares the s-SWCNT and MoS2 carrier densities extracted from each of the methods described above. The two methods agree reasonably well, with percent difference ranging from ∼3% to 37% (excluding the bilayer excited at 532 nm). The good agreement between the methods proposed by Dowgiallo et al.27 and Eckstein et al.16 suggest that each method is viable for estimating s-SWCNT ground- or excited-state carrier density and provides additional confidence in the extracted MoS2 carrier densities used in the following calibration curve.

TABLE I.

Carrier density calculated from both methods discussed above for the MoS2/SWCNT/WSe2 trilayer and multiple MoS2/SWCNT bilayer samples and excitation wavelength corresponding to either hole transfer (HT) or electron transfer (ET) and their relative difference.

λexcFluenceMoS2MoS2 nebMoS2 nec
Sample(nm)(μJ/cm2)RxnA GSBa(cm−2)(cm−2)%Diffd
Bilayer 1 1000 50.0 ET 0.049 1.05 × 1012 7.19 × 1011 37 
Bilayer 1 440 4.60 HT 0.039 8.34 × 1011 5.81 × 1011 36 
Bilayer 2 1000 1.06 ET 0.0076 3.41 × 1011 3.53 × 1011 3.6 
Bilayer 2 440 3.35 HT 0.022 6.19 × 1011 8.68 × 1011 34 
Bilayer 3 1000 1.50 ET 0.000 97 4.10 × 1011 4.84 × 1011 16 
Bilayer 3 532 50.0 HT 0.1132 2.14 × 1012 1.23 × 1012 54 
Trilayer 1000 7.41 ET 0.0032 7.43 × 1010 1.05 × 1011 34 
λexcFluenceMoS2MoS2 nebMoS2 nec
Sample(nm)(μJ/cm2)RxnA GSBa(cm−2)(cm−2)%Diffd
Bilayer 1 1000 50.0 ET 0.049 1.05 × 1012 7.19 × 1011 37 
Bilayer 1 440 4.60 HT 0.039 8.34 × 1011 5.81 × 1011 36 
Bilayer 2 1000 1.06 ET 0.0076 3.41 × 1011 3.53 × 1011 3.6 
Bilayer 2 440 3.35 HT 0.022 6.19 × 1011 8.68 × 1011 34 
Bilayer 3 1000 1.50 ET 0.000 97 4.10 × 1011 4.84 × 1011 16 
Bilayer 3 532 50.0 HT 0.1132 2.14 × 1012 1.23 × 1012 54 
Trilayer 1000 7.41 ET 0.0032 7.43 × 1010 1.05 × 1011 34 
a

Percentage bleach, i.e., ΔA/A0, where f(N)/f(0) = 1−(ΔA/A0).

b

Method of Dowgiallo et al.27 used to calculate s-SWCNT nh.

c

Method of Eckstein et al.16 used to calculate s-SWCNT nh.

d

Difference = [ne# − ne@]/[(ne# + ne@)/2] × 100.

Turning to the MoS2 components of the heterostructures, we can now correlate the extracted MoS2 carrier densities to the associated MoS2 A exciton bleach values. To produce the phase-space filling correlation we ultimately desire,14,15 we must calculate the extent of MoS2 exciton bleaching for a given photoinduced carrier density. This calculation is achieved by performing a spectral deconvolution of the steady-state and transient absorbance spectra in the region of the MoS2 A exciton, as shown in Fig. 2(c) (see also Figs. 6 and 7 of the supplementary material). All steady-state absorbance features have a positive sign, so once the peak for the A exciton is extracted via spectral deconvolution, the area of this peak is used for the original A exciton absorbance (A0). The TA spectrum can be more complex, since spectral shifts and transfer of oscillator strength to trion absorption can induce positive features in addition to the negative GSB.29 As such, the spectrum is fit to a combination of both negative and positive peaks and the total area of the ground-state bleach (GSB) is taken as the summation of the areas (absolute values) of each of these peaks.29 

The phase-space filling effect dictates that electrons in the MoS2 conduction band remove states that contribute oscillator strength to excitonic transitions, thus decreasing oscillator strength (f).14,15 A known solution to the Schrödinger equation can be used when the exciton is represented as an isolated state in an ideal quantum well, where the functional form of the wavefunction is minimally affected by carrier density, but the carriers perturb the exciton size and binding energy via screening (supplementary material, Sec. 4 and Fig. S8).14 Here, the magnitude of exciton bleaching can be related to physical properties in the TMDC, such as the dielectric constant (ɛ) and reduced exciton mass (μ), both of which impact the effective Bohr radius (a0),
(3)
where N is the density of charge carriers and Nc = 2/πa02 is the critical carrier density at which 50% of the oscillator strength is quenched. Nc in turn depends on the exciton size,
(4)
where is the reduced Planck constant and e is the elementary charge. In Fig. 3, we use a reduced exciton mass (μ) of 0.275me for MoS2 (where me is the electron mass), based on the recent study by Goryca et al.20 With a well-constrained empirical value for μ, the exciton size and corresponding dependence of exciton bleach on carrier density depend primarily on the local dielectric environment, so Fig. 3 displays multiple plots with dielectric constants ranging from 4 to 16.5.3,25
FIG. 3.

Charge density (N) dependence of the exciton oscillator strength (f) from the phase-space filling model, plotted for varying dielectric constants (ɛ) produced in MoS2 by exciton dissociation at the MoS2/SWCNT interface in either bilayers or trilayers. Bilayers prepared on MgF2 are open squares, and all other samples are on sapphire and correspond to filled circles. Going from right to left, the dielectric constant varies from 4 to 16.5 in steps of 0.5. The red and green points refer to data extracted by using the methods of Eckstein et al. and Dowgiallo et al., respectively, to calculate the photoinduced s-SWCNT carrier density (see Table I). Additional details of the constants used in the phase-space filling model are given in the text. Panel (a) provides a zoomed in view of the dependence focusing on the individual data points, while panel (b) provides the full dependence that can be used by the community for experiments producing any carrier density, associated exciton bleach, and the FETs experimental data points from Kravets and co-workers.22 

FIG. 3.

Charge density (N) dependence of the exciton oscillator strength (f) from the phase-space filling model, plotted for varying dielectric constants (ɛ) produced in MoS2 by exciton dissociation at the MoS2/SWCNT interface in either bilayers or trilayers. Bilayers prepared on MgF2 are open squares, and all other samples are on sapphire and correspond to filled circles. Going from right to left, the dielectric constant varies from 4 to 16.5 in steps of 0.5. The red and green points refer to data extracted by using the methods of Eckstein et al. and Dowgiallo et al., respectively, to calculate the photoinduced s-SWCNT carrier density (see Table I). Additional details of the constants used in the phase-space filling model are given in the text. Panel (a) provides a zoomed in view of the dependence focusing on the individual data points, while panel (b) provides the full dependence that can be used by the community for experiments producing any carrier density, associated exciton bleach, and the FETs experimental data points from Kravets and co-workers.22 

Close modal

The experimental data fall within a dielectric constant range of 5.5–12.5, with a best fit of ∼8.0 for the apparent local dielectric constant. To rationalize this range, we consider the average dielectric constant, ɛavg = 1/2 (ɛtop + ɛbottom), of the materials surrounding the MoS2 layer. While this approach is commonly applied to understand Coulomb screening, phase space filling, and bandgap renormalization in monolayer TMDCs, some studies use the static dielectric constant of the adjacent materials,30 while other studies utilize the high-frequency dielectric constants.20,31 If we first consider the static dielectric constants of sapphire [ɛ0 = 9–9.4 (in-plane), ɛ0 = 11.6–11.8 (out-of-plane), and ɛavg = 10.5]32,33 and SWCNTs (ɛ0 = 4.0),34–36 we arrive at ɛavg = 7.3. Turning to the high-frequency dielectric constants, using ɛ0 = 4.0 for the s-SWCNT film and ɛ = 3.05 for sapphire33 leads to ɛavg = 3.5. While there are insufficient reports of the high-frequency dielectric constant of highly enriched (near-monochiral) s-SWCNTs in the literature, a recent report suggests that the dielectric function is relatively flat in the far-infrared regime.37 

At first glance, this analysis suggests that the use of static dielectric constants might be most appropriate for describing phase-space filling in these photoexcited heterojunctions. To test this, we studied s-SWCNT/MoS2 bilayers prepared on MgF2 (ɛ0 = 5.2, ɛ = 1.7, square data points in Fig. 3)32,38 instead of sapphire. The average dielectric environment between MgF2 and s-SWCNTs would be ∼ɛavg = 4.1 and ɛavg = 2.4, using static and high-frequency dielectric constants, respectively. The MgF2 results all fall near the ɛ = 5.5 fit line, in reasonable agreement with the average static dielectric constants, but incommensurate with the high-frequency dielectric constants. As such, we tentatively conclude that phase-space filling in these photoexcited monolayer MoS2 heterostructures is best described by considering the static dielectric constants of the surrounding materials.

To put our results into context with the existing literature, we have also plotted the results of several studies that correlated exciton quenching with the carrier density of monolayer TMDC transistor channels. As shown in Fig. 3(b), the bounds of the phase-space filling model employed in our study do a good job of describing existing monolayer MoS2 transistor-based measurements.22 This comparison provides additional confidence in our results and may also suggest that using the gate oxide capacitance to extract carrier density is a reasonable approximation for the transistors studied by Kravets et al.22 Similarly, separately derived phase-space filling simulations do a good job of describing the transistor-based trends found for WS223 and MoSe224 monolayers (Fig. S9).

We can also connect our results to the prevailing literature exploring the carrier density-dependent renormalization of electronic (quasiparticle) bandgap and exciton binding energy in monolayer MoS2. Bandgap renormalization is known to be substantial in doped TMDCs and has been studied extensively for MoS2.39–41 Since the phase-space filling model depends on the exciton binding energy, we turn to the study of Yao et al. that discriminated between the impacts of injected charge carriers on the electronic bandgap and exciton binding energy.42 In Fig. S10, we plot the carrier density-dependent change in exciton binding energy, normalized to the original binding energy, for the best-fit and upper/lower bounds of the phase space filling simulation in Fig. 3, along with the same trend found by Yao et al. (green data points). The data from Yao et al.42 fall nicely within the bounds predicted by our phase-space simulation, providing additional confidence in our results.

Returning to Fig. 3(b), while the data at low carrier densities are most consistent with an apparent local dielectric constant in the range of ∼5.5, the data move progressively toward fit lines corresponding to higher apparent dielectric constants (up to ∼12.5) as carrier density increases. This trend suggests that, although it appears to capture the expected excitonic transition bleaching by charge carriers quite well, the 2D phase-space filling model may not capture all of the underlying physics at play in monolayer MoS2. Since our spectroscopic data and data from transistor-based measurements both seem to fall off slightly faster than might be predicted by the phase-space filling model, we consider it most likely that the behavior is intrinsic to TMDCs and does not reflect a shortcoming of either type of measurement methods.

The phase-space filling curve elucidated here (with upper/lower bounds) can be used by the community to determine monolayer MoS2 carrier density in a number of different experiments, as long as the samples/devices can be interrogated by transmission or reflection-based optical spectroscopy. These include experiments and devices where steady-state carrier densities are injected into the monolayer via electrostatic [e.g., (photo)transistors]43,44 or electrochemical gating (e.g., ion-gated transistors or spectro-electrochemical cells)40 or by adsorbed molecular redox dopants,45 as well as time-resolved experiments where excited-state carrier densities are modulated by dynamic processes, such as photoinduced charge transfer.3 To recreate the full phase-space filling model shown in Fig. 3, one can use Eqs. (3) and (4) and the most appropriate estimate for the dielectric environment [ɛavg = 1/2 (ɛtop + ɛbottom)] of the monolayer sample, to generate the relevant dependence. We suggest the use of the recently derived reduced exciton mass μ = 0.275me20 and, based on the empirical data shown here, the static dielectric constants of the surrounding materials.

In this study, we develop a scaling curve that correlates the loss of A exciton oscillator strength in monolayer MoS2 to the 2D carrier density by means of a s-SWCNT internal standard in a series of photoinduced mixed-dimensionality type II heterostructures. The trends in Fig. 3 highlight that determining an appropriate value for the dielectric constant experienced by an exciton in a monolayer MoS2/s-SWCNT system is not trivial. The scaling curve utilizes a phase-space filling model originally developed for 2D quantum wells and incorporates empirically valid values for the physical properties of 2D MoS2 excitons. The consistency between various s-SWCNT/MoS2 samples within the phase-space filling model supports the use of similar carrier density scaling curves proposed recently by Eckstein et al. and Dowgiallo et al., expanding the relatively underdeveloped assessment of charge carrier concentration in monolayer TMDCs. This study also demonstrates that the well-defined charge-associated transient absorption spectral features s-SWCNTs provide a useful guidepost for understanding how to reliably quantify charge transfer quantum yield and the role of dielectric environment on charge separation and recombination when paired with a variety of TMDCs.

The supplementary material contains materials and methods and calculations for Nexc, charge yield, and the phase-space filling model.

This work was authored by the National Renewable Energy Laboratory, operated by the Alliance for Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. This study was supported by the Solar Photochemistry Program, Division of Chemical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciences, U.S. DOE. The views expressed in the article do not necessarily represent views of the DOE or the U.S. Government.

The authors have no conflicts to disclose.

Alexis R. Myers: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Dana B. Sulas-Kern: Data curation (equal); Investigation (equal); Writing – review & editing (equal). Rao Fei: Data curation (equal); Investigation (equal); Writing – review & editing (equal). Debjit Ghoshal: Data curation (equal); Formal analysis (equal); Investigation (equal); Writing – review & editing (equal). M. Alejandra Hermosilla-Palacios: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Jeffrey L. Blackburn: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
B. A.
Parkinson
,
T. E.
Furtak
,
D.
Canfield
,
K.-K.
Kam
, and
G.
Kline
, “
Evaluation and reduction of efficiency losses at tungsten diselenide photoanodes
,”
Faraday Discuss. Chem. Soc.
70
,
233
245
(
1980
).
2.
H.
Tributsch
and
J. C.
Bennett
, “
Electrochemistry and photochemistry of MoS2 layer crystals. I
,”
J. Electroanal. Chem. Interfacial Electrochem.
81
,
97
111
(
1977
).
3.
D. B.
Sulas-Kern
,
E. M.
Miller
, and
J. L.
Blackburn
, “
Photoinduced charge transfer in transition metal dichalcogenide heterojunctions – towards next generation energy technologies
,”
Energy Environ. Sci.
13
,
2684
2740
(
2020
).
4.
M. I. B.
Utama
,
A.
Dasgupta
,
R.
Ananth
,
E. A.
Weiss
,
T. J.
Marks
, and
M. C.
Hersam
, “
Mixed-dimensional heterostructures for quantum photonic science and technology
,”
MRS Bull.
48
,
905
913
(
2023
).
5.
R.
Yang
,
J.
Fan
, and
M.
Sun
, “
Transition metal dichalcogenides (TMDCs) heterostructures: Optoelectric properties
,”
Front. Phys.
17
,
43202
(
2022
).
6.
T.
Sun
,
H.
Zhang
,
X.
Wang
,
J.
Liu
,
C.
Xiao
,
S. U.
Nanayakkara
,
J. L.
Blackburn
,
M. V.
Mirkin
, and
E. M.
Miller
, “
Nanoscale mapping of hydrogen evolution on metallic and semiconducting MoS2 nanosheets
,”
Nanoscale Horiz.
4
,
619
624
(
2019
).
7.
Z.
Li
,
N. H.
Attanayake
,
J. L.
Blackburn
, and
E. M.
Miller
, “
Carbon dioxide and nitrogen reduction reactions using 2D transition metal dichalcogenide (TMDC) and carbide/nitride (MXene) catalysts
,”
Energy Environ. Sci.
14
,
6242
6286
(
2021
).
8.
K.
Wang
,
K.
De Greve
,
L. A.
Jauregui
,
A.
Sushko
,
A.
High
,
Y.
Zhou
,
G.
Scuri
,
T.
Taniguchi
,
K.
Watanabe
,
M. D.
Lukin
,
H.
Park
, and
P.
Kim
, “
Electrical control of charged carriers and excitons in atomically thin materials
,”
Nat. Nanotechnol.
13
,
128
132
(
2018
).
9.
R.
Krishnan
,
S.
Biswas
,
Y.-L.
Hsueh
,
H.
Ma
,
R.
Rahman
, and
B.
Weber
, “
Spin-valley locking for in-gap quantum dots in a MoS2 transistor
,”
Nano Lett.
23
,
6171
6177
(
2023
).
10.
S.-L.
Li
,
K.
Tsukagoshi
,
E.
Orgiu
, and
P.
Samorì
, “
Charge transport and mobility engineering in two-dimensional transition metal chalcogenide semiconductors
,”
Chem. Soc. Rev.
45
,
118
151
(
2016
).
11.
A.
Thilgam
, “
Exciton complexes in low dimensional transition metal dichalcogenides
,”
J. Appl. Phys.
116
,
053523
(
2014
).
12.
N.
Ma
and
D.
Jena
, “
Carrier statistics and quantum capacitance effects on mobility extraction in two-dimensional crystal semiconductor field-effect transistors
,”
2D Mater.
2
,
015003
(
2015
).
13.
A.
Raja
,
A.
Chaves
,
J.
Yu
,
G.
Arefe
,
H. M.
Hill
,
A. F.
Rigosi
,
T. C.
Berkelbach
,
P.
Nagler
,
C.
Schüller
,
T.
Korn
et al, “
Coulomb engineering of the bandgap and excitons in two-dimensional materials
,”
Nat. Commun.
8
,
15251
(
2017
).
14.
D.
Huang
,
J.-I.
Chyi
, and
H.
Morkoç
, “
Carrier effects on the excitonic absorption in GaAs quantum-well structures: Phase-space filling
,”
Phys. Rev. B
42
,
5147
5153
(
1990
).
15.
A.
Cameron
,
P.
Riblet
, and
A.
Miller
, “
Broadening, screening, and phase space filling in GaAs multiple quantum wells revisted
,” in
Quantum Electronics and Laser Science Conference
(
IEEE
,
1996
), Vol.
10
.
16.
K. H.
Eckstein
,
F.
Oberndorfer
,
M. M.
Achsnich
,
F.
Schöppler
, and
T.
Hertel
, “
Quantifying doping levels in carbon nanotubes by optical spectroscopy
,”
J. Phys. Chem. C
123
,
30001
30006
(
2019
).
17.
M. A.
Hermosilla-Palacios
,
M.
Martinez
,
E. A.
Doud
,
T.
Hertel
,
A. M.
Spokoyny
,
S.
Cambré
,
W.
Wenseleers
,
Y.-H.
Kim
,
A. J.
Ferguson
, and
J. L.
Blackburn
, “
Carrier density and delocalization signatures in doped carbon nanotubes from quantitative magnetic resonance
,”
Nanoscale Horiz.
9
,
278
284
(
2024
).
18.
S.
Park
,
N.
Mutz
,
T.
Schultz
,
S.
Blumstengel
,
A.
Han
,
A.
Aljarb
,
L.-J.
Li
,
E. J.
List-Kratochvil
,
P.
Amsalem
, and
N.
Koch
, “
Direct determination of monolayer MoS2 and WSe2 exciton binding energies on insulating and metallic substrates
,”
2D Mater.
5
,
025003
(
2018
).
19.
A.
Laturia
,
M. L.
Van de Put
, and
W. G.
Vandenberghe
, “
Dielectric properties of hexagonal boron nitride and transition metal dichalcogenides: From monolayer to bulk
,”
npj 2D Mater. Appl.
2
,
6
(
2018
).
20.
M.
Goryca
,
J.
Li
,
A.
Stier
,
T.
Taniguchi
,
K.
Watanabe
,
E.
Courtade
,
S.
Shree
,
C.
Robert
,
B.
Urbaszek
,
X.
Marie
, and
S.
Crooker
, “
Revealing exciton masses and dielectric properties of monolayer semiconductors with high magnetic fields
,”
Nat. Commun.
10
,
4172
(
2019
).
21.
S.
Latini
,
K. T.
Winther
,
T.
Olsen
, and
S.
Thygesen
, “
Interlayer excitons and band alignment in MoS2/hBN/WSe2 van der Waals heterostructures
,”
Nano Lett.
17
,
938
945
(
2017
).
22.
V. G.
Kravets
,
F.
Wu
,
G. H.
Auton
,
T.
Yu
,
S.
Imaizumi
, and
A.
Grigorenko
, “
Measurements of electrically tunable refractive index of MoS2 monolayer and its usage in optical modulators
,”
npj 2D Mater. Appl.
3
,
36
(
2019
).
23.
Y.
Yu
,
Y.
Yu
,
L.
Huang
,
H.
Peng
,
L.
Xiong
, and
L.
Cao
, “
Giant gating tunability of optical refractive index in transition metal dichalcogenide monolayers
,”
Energy Environ. Sci.
12
,
1648
1656
(
2017
).
24.
M.
Li
,
S.
Biswas
,
C.
Hail
, and
H.
Atwater
, “
Refractive index modulation in monolayer molybdenum diselenide
,”
Nano Lett.
21
,
7602
7608
(
2021
).
25.
D. B.
Sulas-Kern
,
H.
Zhang
,
Z.
Li
, and
J. L.
Blackburn
, “
Microsecond charge separation at heterojunctions between transition metal dichalcogenide monolayers and single-walled carbon nanotubes
,”
Mater. Horiz.
6
,
2103
2111
(
2019
).
26.
D.
Sulas-Kern
,
H.
Zhang
,
K.
Li
, and
J.
Blackburn
,
Nanoscale
13
,
8188
8198
(
2021
).
27.
A.-M.
Dowgiallo
,
K.
Mistry
,
J.
Johnson
, and
J.
Blackburn
,
ACS Nano
8
,
8573
8581
(
2014
).
28.
A. R.
Myers
,
Z.
Li
,
M. K.
Gish
,
J. D.
Earley
,
J. C.
Johnson
,
M. A.
Hermosilla-Palacios
, and
J. L.
Blackburn
, “
Ultrafast charge transfer cascade in a mixed-dimensionality nanoscale trilayer
,”
ACS Nano
18
,
8190
8198
(
2024
).
29.
J.
Kim
,
C.
Jin
,
B.
Chen
,
H.
Cai
,
T.
Zhao
,
P.
Lee
,
S.
Kahn
,
K.
Watanabe
,
T.
Taniguchi
,
S.
Tongay
et al, “
Observation of ultralong valley lifetime in WSe2/MoS 2 heterostructures
,”
Sci. Adv.
3
,
e1700518
(
2017
).
30.
Z.
Qiu
,
M.
Trushin
,
H.
Fang
,
I.
Verzhbitskiy
,
S.
Gao
,
E.
Laksono
,
M.
Yang
,
P.
Lyu
,
J.
Li
,
J.
Su
,
M.
Telychko
,
K.
Watanabe
,
T.
Taniguchi
,
J.
Wu
,
A. H. C.
Neto
,
L.
Yang
,
G.
Eda
,
S.
Adam
, and
J.
Lu
, “
Giant gate-tunable bandgap renormalization and excitonic effects in a 2D semiconductor
,”
Sci. Adv.
5
,
eaaw2347
(
2019
).
31.
A. V.
Stier
,
N. P.
Wilson
,
G.
Clark
,
X.
Xu
, and
S. A.
Crooker
, “
Probing the influence of dielectric environment on excitons in monolayer WSe2: Insight from high magnetic fields
,”
Nano Lett.
16
,
7054
7060
(
2016
).
32.
J.
Fontanella
,
C.
Andeen
, and
D.
Schuele
, “
Low-frequency dielectric constants of α-quartz, sapphire, MgF2, and MgO
,”
J. Appl. Phys.
45
,
2852
2854
(
1974
).
33.
A. K.
Harman
,
S.
Ninomiya
, and
S.
Adachi
, “
Optical constants of sapphire (α-Al2O3) single crystals
,”
J. Appl. Phys.
76
,
8032
8036
(
1994
).
34.
V.
Perebeinos
,
J.
Tersoff
, and
P.
Avouris
, “
Scaling of excitons in carbon nanotubes
,”
Phys. Rev. Lett.
92
,
257402
(
2004
).
35.
Y.
Miyauchi
,
R.
Saito
,
K.
Sato
,
Y.
Ohno
,
S.
Iwasaki
,
T.
Mizutani
,
J.
Jiang
, and
S.
Maruyama
, “
Dependence of exciton transition energy of single-walled carbon nanotubes on surrounding dielectric materials
,”
Chem. Phys. Lett.
442
,
394
399
(
2007
).
36.
K.
Eckstein
and
T.
Hertel
, “
Electronic structure and scaling of Coulomb defects in carbon nanotubes from modified Hückel calculations
,”
J. Phys. Chem. C
127
,
23760
23767
(
2023
).
37.
T.
Nishihara
,
A.
Takakura
,
M.
Shimasaki
,
K.
Matsuda
,
T.
Tanaka
,
H.
Kataura
, and
Y.
Miyauchi
, “
Empirical formulation of broadband complex refractive index spectra of single-chirality carbon nanotube assembly
,”
Nanophotonics
11
,
1011
1020
(
2022
).
38.
J. M.
Siqueiros
,
R.
Machorro
, and
L. E.
Regalado
, “
Determination of the optical constants of MgF2 and ZnS from spectrophotometric measurements and the classical oscillator method
,”
Appl. Opt.
27
,
2549
2553
(
1988
).
39.
A.
Faridi
,
D.
Culcer
, and
R.
Asgari
, “
Quasiparticle band-gap renormalization in doped monolayer MoS2
,”
Phys. Rev. B
104
,
085432
(
2021
).
40.
R.
Almaraz
,
T.
Sayer
,
J.
Toole
,
R.
Austin
,
Y.
Farah
,
N.
Trainor
,
J. M.
Redwing
,
A.
Krummel
,
A.
Montoya-Castillo
, and
J.
Sambur
, “
Quantifying interfacial energetics of 2D semiconductor electrodes using in situ spectroelectrochemistry and many-body theory
,”
Energy Environ. Sci.
16
,
4522
4529
(
2023
).
41.
G. M.
Carroll
,
H.
Zhang
,
J. R.
Dunklin
,
E. M.
Miller
,
N. R.
Neale
, and
J.
Van De Lagemaat
, “
Unique interfacial thermodynamics of few-layer 2D MoS2 for (photo)electrochemical catalysis
,”
Energy Environ. Sci.
12
,
1648
1656
(
2019
).
42.
K.
Yao
,
A.
Yan
,
S.
Kahn
,
A.
Suslu
,
Y.
Liang
,
E.
Barnard
,
S.
Tongday
,
A.
Zettl
,
N.
Borys
, and
P. J.
Schuck
, “
Optically discriminating carrier-induced quasiparticle band gap and exciton energy renormalization in monolayer MoS2
,”
Phys. Rev. Lett.
119
,
087401
(
2017
).
43.
J.
Pak
,
I.
Lee
,
K.
Cho
,
J.-K.
Kim
,
H.
Jeong
,
W.-T.
Hwang
,
G. H.
Ahn
,
K.
Kang
,
W. J.
Yu
,
A.
Javey
et al, “
Intrinsic optoelectronic characteristics of MoS2 phototransistors via a fully transparent van der Waals heterostructure
,”
ACS Nano
13
,
9638
9646
(
2019
).
44.
K. F.
Mak
,
K.
He
,
C.
Lee
,
G. H.
Lee
,
J.
Hone
,
T. F.
Heinz
, and
J.
Shan
, “
Tightly bound trions in monolayer MoS2
,”
Nat. Mater.
12
,
207
211
(
2013
).
45.
K. P.
Dhakal
,
D. L.
Duong
,
J.
Lee
,
H.
Nam
,
M.
Kim
,
M.
Kan
,
Y. H.
Lee
, and
J.
Kim
, “
Confocal absorption spectral imaging of MoS2: Optical transitions depending on the atomic thickness of intrinsic and chemically doped MoS2
,”
Nanoscale
6
,
13028
13035
(
2014
).
Published open access through an agreement with National Renewable Energy Laboratory