• Journal of Semiconductors
  • Vol. 46, Issue 4, 042102 (2025)
Chenlin Wang1, Haixiao Zhao2, Xian Zhao2,3, Baoqing Sun1,2..., Jie Lian1,2 and Yuan Gao1,2,*|Show fewer author(s)
Author Affiliations
  • 1School of Information Science and Engineering, Shandong University, Qingdao 266237, China
  • 2Key Laboratory of Laser & Infrared System (Shandong University), Ministry of Education, Shandong University, Qingdao 266237, China
  • 3Center for Optics Research and Engineering (CORE), Shandong University, Qingdao 266237, China
  • show less
    DOI: 10.1088/1674-4926/24100011 Cite this Article
    Chenlin Wang, Haixiao Zhao, Xian Zhao, Baoqing Sun, Jie Lian, Yuan Gao. Layer-dependent optical and dielectric properties of CdSe semiconductor colloidal quantum wells characterized by spectroscopic ellipsometry[J]. Journal of Semiconductors, 2025, 46(4): 042102 Copy Citation Text show less

    Abstract

    Semiconductor colloidal quantum wells (CQWs) with atomic-precision layer thickness are rapidly gaining attention for next-generation optoelectronic applications due to their tunable optical and electronic properties. In this study, we investigate the dielectric and optical characteristics of CdSe CQWs with monolayer numbers ranging from 2 to 7, synthesized via thermal injection and atomic layer (c-ALD) deposition techniques. Through a combination of spectroscopic ellipsometry (SE) and first-principles calculations, we demonstrate the significant tunability of the bandgap, refractive index, and extinction coefficient, driven by quantum confinement effects. Our results show a decrease in bandgap from 3.1 to 2.0 eV as the layer thickness increases. Furthermore, by employing a detailed analysis of the absorption spectra, accounting for exciton localization and asymmetric broadening, we precisely capture the relationship between monolayer number and exciton binding energy. These findings offer crucial insights for optimizing CdSe CQWs in optoelectronic device design by leveraging their layer-dependent properties.

    Introduction

    Semiconductor colloidal quantum wells (CQWs), particularly those based on cadmium selenide (CdSe), are emerging as a class of materials with considerable potential in the field of optoelectronics due to their unique optical and electronic properties[13]. These materials exhibit atomic-precision layer thickness, allowing for highly tunable characteristics, such as bandgap[4], refractive index, and extinction coefficient[5, 6], which are essential for the design of high-performance optoelectronic devices. The ability to fine-tune these properties through control of the number of monolayers offers new opportunities for integrating CdSe CQWs into advanced applications, including photodetectors[7, 8], solar cells[9], lasers[10, 11], and quantum emitters[12, 13].

    In recent years, two-dimensional (2D) materials have attracted widespread attention due to their modulated electronic structures and unique lattice symmetries. Examples of such materials include perovskites[14], black phosphorus[15], and transition metal dichalcogenides[16], all of which demonstrate highly desirable properties, such as a wide spectral response range, high carrier mobility, optical gain characteristics, and rapid response speeds. However, achieving these properties within a single material system, while maintaining environmental stability, remains a major challenge, hindering their broader application. Semiconductor CQWs, particularly those made from CdSe, offer a solution to this challenge by providing a material system that can be precisely tailored to meet the demands of various optoelectronic applications[1].

    CdSe CQWs, as typical Ⅱ−Ⅵ semiconductors, exhibit tunable bandgaps ranging from approximately 3.5 eV in single-layer CQWs to 1.6 eV in multi-layer configurations. This tunability arises from quantum confinement effects, where the thickness of the CQW strongly influences the electronic and optical properties of the material. The addition of shell layers further enhances these tunable properties, making CQWs a flexible platform for device development[1720]. Moreover, CdSe CQWs demonstrate high fluorescence quantum yields, narrow emission spectra, and carrier mobilities reaching up to 100 cm²∙(V·s)−1[21, 22]. These properties make them ideal candidates for applications requiring efficient light absorption and emission, such as in quantum emission devices[23, 24], solar concentrators[25], and spintronic technologies[26].

    The synthesis of CdSe CQWs is typically achieved through solution-based methods, with thermal injection being a widely used approach. This method involves the rapid injection of precursors into a high-temperature long-chain ligand solvent, enabling precise control over the thickness and composition of each atomic layer[27]. The ability to finely tune the number of monolayers and the chemical structure of CdSe CQWs[28] through this synthesis method significantly enhances their optoelectronic performance, allowing for better device optimization[29, 30]. However, the performance of these devices is strongly dependent on the intrinsic dielectric and optical properties of the CQWs, which include the dielectric function and complex refractive index (n* = n + ik). These parameters govern the interaction between light and matter in the CQWs and display a strong dependence on the number of monolayers.

    The refractive index (n) and extinction coefficient (k) are critical material properties that influence the propagation and absorption of light within optoelectronic devices[31]. In CdSe CQW-based photodetectors, solar cells, and lasers, the complex refractive index directly impacts key performance characteristics such as the absorption spectrum, quantum yield, and lasing threshold. For example, in CQW-based lasers, the refractive index determines the mode gain and loss, which in turn affects the lasing threshold[32, 33]. Similarly, in photodetectors and solar cells, the refractive index influences the absorption depth and efficiency of the device[34, 35]. However, precise knowledge of the complex refractive index is often lacking, which can lead to inaccurate simulations and predictions of device performance.

    Various techniques have been employed to determine the dielectric function and complex refractive index of 2D materials, including Kramers−Kronig constrained variational analysis[36], differential reflection (or transmission) spectroscopy[37], scattering-type scanning near-field optical microscopy (s-SNOM)[38, 39], and spectroscopic ellipsometry (SE)[4042]. Among these, SE has emerged as a particularly effective method for extracting the optical constants of 2D materials, as it allows for the accurate determination of both the refractive index and extinction coefficient without requiring additional assumptions or functions. SE works by detecting the changes in the polarization state of light before and after interaction with the sample, providing a detailed and reliable analysis of the material's optical properties.

    In this study, we investigate the layer-dependent dielectric and optical properties of CdSe CQWs with monolayer numbers ranging from 2 to 7. Using a combination of thermal injection and atomic layer deposition, we successfully synthesized high-quality CdSe CQWs with precise control over their thickness. We utilized spectroscopic ellipsometry to systematically study the evolution of the complex refractive index, dielectric function, and bandgap as a function of monolayer thickness. The optical bandgap of the CdSe CQWs was calculated using absorption spectra and the Tauc formula, revealing a decrease in the bandgap from 3.1 eV at 2 monolayers to 2.0 eV at 7 monolayers, consistent with quantum confinement effects. Additionally, we performed first-principles calculations to further explore the relationship between monolayer thickness and the material's optical properties.

    To capture the excitonic behavior in CdSe CQWs, we employed an approach that accounts for exciton localization and asymmetric broadening in the absorption spectra, allowing for an accurate determination of exciton binding energy as a function of layer thickness. This analysis provides valuable insights into the correlation between monolayer number and excitonic properties, which are crucial for optimizing the performance of optoelectronic devices based on CQWs. Our work presents a comprehensive investigation of the layer-dependent dielectric and optical properties of CdSe CQWs, offering a detailed understanding of how monolayer thickness influences key material characteristics. Our findings provide critical guidance for the design and optimization of CQW-based optoelectronic devices, where precise control over material properties is essential for achieving high performance.

    Result and discussion

    High-quality CdSe CQWs with 2 to 7 monolayers (ML) were synthesized using thermal injection and c-ALD methods. Cadmium myristate and selenium were used as precursors, and CQWs ranging from 2 to 5 ML were grown in a three-neck flask by gradually increasing the reaction temperature. After the nucleation and growth of 4 and 5 ML CdSe CQWs, a mixture of cadmium acetate dihydrate (Cd(OAc)₂·2H₂O) in N-methylformamide (NMF) was injected into the reaction medium. The reaction mixture was then heated to 230 °C, promoting lateral growth and increasing the thickness of the initially formed 4 and 5 ML CQWs by steps of 1 ML. The experimental details are provided in the supplementary material.

    Fig. 1(a) displays the luminescence images of 2 to 7 ML CdSe CQWs under ultraviolet irradiation, showing a gradual spectral shift from blue to red as the monolayer number increases. The CQWs were spin-coated onto quartz substrates for further characterization. The atomic planes of cadmium and selenium alternate along the CQWs’ short axis, aligned with the (001) direction of the cubic structure, as illustrated in Fig. 1(d). In a cubic zinc blende structure, the six (100) directions are equivalent, but in CdSe CQWs, the (001) direction typically defines the thickness. Surface ligands cause stress and a resulting tetragonal distortion, making the lattice parameters in the thickness direction unequal to those in the in-plane directions.

    (Color online) (a) Photographs of 2 to 7 ML CdSe CQWs under ultraviolet light, showing the variation in photoluminescence with increasing layer thickness. (b) AFM surface topography of a 4 ML CdSe CQW film on a quartz substrate, illustrating the smoothness of the film. Scale bar: 3 μm. (c) Scratch height profile of the 4 ML CdSe CQW film, showing the step height between the film and the substrate. (d) Visualization of the (220) crystal plane in the CdSe CQW model. (e) Static X-ray diffraction (XRD) patterns of 2 to 7 ML CdSe CQWs, indicating structural properties and phase purity. (f) Gaussian fitting of the (220) plane in the 4 ML CdSe CQWs. (g) Enlarged view of the (111) diffraction peaks for 2 to 7 ML CdSe CQWs.

    Figure 1.(Color online) (a) Photographs of 2 to 7 ML CdSe CQWs under ultraviolet light, showing the variation in photoluminescence with increasing layer thickness. (b) AFM surface topography of a 4 ML CdSe CQW film on a quartz substrate, illustrating the smoothness of the film. Scale bar: 3 μm. (c) Scratch height profile of the 4 ML CdSe CQW film, showing the step height between the film and the substrate. (d) Visualization of the (220) crystal plane in the CdSe CQW model. (e) Static X-ray diffraction (XRD) patterns of 2 to 7 ML CdSe CQWs, indicating structural properties and phase purity. (f) Gaussian fitting of the (220) plane in the 4 ML CdSe CQWs. (g) Enlarged view of the (111) diffraction peaks for 2 to 7 ML CdSe CQWs.

    To investigate the lattice structure, X-ray diffraction (XRD) was used on CQWs with different monolayer numbers, as shown in Fig. 1(e). The XRD patterns revealed characteristic diffraction peaks associated with the zinc blende phase of CdSe, which belongs to the F−43m space group. As the number of monolayers increased, the reaction temperature also rose, but this did not affect the zinc blende phase, indicating the purity and structural integrity of the synthesized samples.

    The (111) peaks were smoothed and fitted using a Gaussian function with OMNIC software to calculate the lattice constants. As the monolayer number decreased, the lattice constants deviated further from those of bulk CdSe (JCPDS database #19−0191; space group F−43m; a = 6.077 Å), as presented in Table S1. This deviation is due to the smaller number of single crystal cells and the presence of surface ligands, which induce in-plane lattice expansion and a reduction in the vertical lattice constants. These variations indicate that the lattice parameters are no longer equivalent in all three directions, lifting the degeneracy seen in the zinc blende diffraction pattern.

    As the crystal size decreases, Scherrer broadening causes the (400) diffraction peak to appear primarily in the plane along the longest dimension of the CQWs [(100) axis or x-direction)], although it may become too weak or broad to be clearly observed. The (220) peak of 4 ML CdSe CQWs was also fitted, as shown in Fig. 1(f). The peaks in the (220), (202), and (022) families showed varying widths and intensities, due to the differing repetition of single crystal cells along these directions. Among these, the (022) peak exhibited the strongest intensity. The analysis of the (220) peak family provides insights into both in-plane and out-of-plane dynamics in the CQWs.

    The surface morphology and thickness of the CdSe CQW films were examined using atomic force microscopy (AFM) in non-contact mode. As shown in Fig. 1(b), the films exhibit a specific degree of surface roughness. The RMS roughness values are summarized in Table S2. Knowing the approximate film thickness was crucial for the accurate fitting process during spectroscopic ellipsometry analysis. The thickness was measured by lightly scratching the film with tweezers and analyzing the height difference between the film and the substrate using AFM. The resulting film thicknesses—134, 132, 130, 125, 129, and 127 nm—correspond well with the measurements obtained from spectroscopic ellipsometry (see Table S2).

    To investigate the optical properties and layer-dependent bandgap of CdSe CQWs, we measured the absorption and photoluminescence (PL) spectra of CdSe films with varying monolayer numbers, covering an energy range of 1.8 to 3.9 eV (310 to 690 nm), as shown in Fig. 2(a) and Fig. S2. In the absorption spectra, CdSe CQWs exhibit three distinct bleaching features corresponding to heavy-hole (HH), light-hole (LH), and spin−orbit (SO) hole transitions to the first electronic shelf (E1). As we analyzed the series of CdSe CQWs, we observed that the absorption peaks associated with HH excitons shifted from 394 to 607 nm with increasing monolayer numbers. This shift occurs because quantum confinement is primarily restricted to the thickness of the CQWs, as the lateral dimensions are larger than the exciton Bohr radius. Consequently, CQWs have narrower optical characteristics compared to spherical nanocrystals (NCs), which are more isotropic in their quantum confinement. For example, the PL spectrum of 4 ML CdSe CQWs displays a sharp peak at 517 nm with a full width at half maximum (FWHM) of 12 nm (see supplementary material for details). The quantum confinement effect limits the electron energy levels spatially, resulting in a bandgap increase as the monolayer number decreases, which is reflected in the blue shift of the PL spectra.

    (Color online) (a) Absorption spectra of 2 to 7 ML CdSe CQWs, illustrating the layer-dependent shift in absorption peaks. HH and LH represent the heavy-hole and light-hole transitions in the absorption spectra. (b) Tauc plots for 2 to 7 ML CdSe CQWs, with dotted lines indicating the intercept method used to estimate the optical bandgap. (c) Relationship between the optical bandgap as derived from the Tauc plots (red line) and density functional theory (DFT) calculations (blue line) as a function of monolayer thickness. Simulated band structure obtained through first-principles calculations for (d) 2 ML CdSe CQWs, (e) 4 ML CdSe CQWs, and (f) 7 ML CdSe CQWs, highlighting the changes in electronic structure with increasing monolayers.

    Figure 2.(Color online) (a) Absorption spectra of 2 to 7 ML CdSe CQWs, illustrating the layer-dependent shift in absorption peaks. HH and LH represent the heavy-hole and light-hole transitions in the absorption spectra. (b) Tauc plots for 2 to 7 ML CdSe CQWs, with dotted lines indicating the intercept method used to estimate the optical bandgap. (c) Relationship between the optical bandgap as derived from the Tauc plots (red line) and density functional theory (DFT) calculations (blue line) as a function of monolayer thickness. Simulated band structure obtained through first-principles calculations for (d) 2 ML CdSe CQWs, (e) 4 ML CdSe CQWs, and (f) 7 ML CdSe CQWs, highlighting the changes in electronic structure with increasing monolayers.

    In layered 2D materials like CdSe CQWs, the absorption spectra combined with the Tauc formula are commonly used to calculate the optical bandgap. The Tauc relation is given by[40]:(αhv)ρ=C×(hvEg),where hv is the photon energy, Eg is the optical bandgap, and C is a constant. The parameter ρ represents the nature of the electronic transition and is typically 1/2 for indirect bandgap materials and 2 for direct bandgap materials. Since CdSe is a direct bandgap semiconductor, ρ is set to 2. By plotting (αhv)² against photon energy, the optical bandgap of CdSe CQWs can be estimated from the intercept, as illustrated in Fig. 2(b). The results show that the optical bandgap increases as the number of monolayers decreases, depicted in Fig. 2(c).

    To understand the reason behind the increasing bandgap with decreasing monolayer numbers, we calculated the band structure of CdSe CQWs using first-principles density functional theory (DFT). The theoretical simulations for 2, 4, and 7 ML CdSe CQWs are shown in Fig. 2(d)−2(f) (additional simulations for 3, 5, and 6 ML are provided in the supplementary material). The calculated bandgaps are generally smaller than experimental values, which can be attributed to the limitations of DFT in accounting for long-range multibody interactions. Nonetheless, the calculations confirm that CdSe is a direct bandgap semiconductor, with the bandgap increasing as the number of monolayers decreases. The calculated bandgaps for CdSe are 2.8, 2.3, 2.1, 1.9, 1.7, and 1.6 eV for 2, 3, 4, 5, 6, and 7 ML, respectively.

    The observed bandgap variation is driven by interlayer coupling, which increases with monolayer numbers and leads to orbital hybridization. This phenomenon is not unique to CdSe; it has also been observed in other 2D materials such as PdSe₂[43] and MoS₂[44], where the bandgap decreases as the number of layers increases. However, CdSe exhibits stronger orbital hybridization due to the higher number of valence electrons in cadmium (12) compared to palladium (10) and molybdenum (6), resulting in a more pronounced change in bandgap with varying monolayer numbers.

    To determine the optical constants of CdSe CQWs and systematically examine their variation with different monolayer numbers, we employed SE to analyze CdSe CQW films with thicknesses of CQWs ranging from 2 to 7 ML. Based on AFM measurements, the surface of the CdSe CQW films exhibited similar roughness. To account for this and avoid compromising the SE analysis, we developed a four-phase optical model (air/rough layer/CdSe CQW film/quartz substrate), as shown in Fig. 3(a). The optical constants of the quartz substrate were previously measured using SE and verified against standard reference values.

    (Color online) (a) Structural model of the CdSe CQW film on a quartz substrate. (b) Experimental spectroscopic ellipsometry spectra (solid lines) and best-fit curves (dots) for 4 ML CdSe CQWs at a 65° incidence angle. (c) Real part of the dielectric function for 2 to 7 ML CdSe CQWs, showing the variation with increasing monolayer thickness. (d) Imaginary part of the dielectric function for 2 to 7 ML CdSe CQWs, illustrating the evolution of optical transitions with layer thickness.

    Figure 3.(Color online) (a) Structural model of the CdSe CQW film on a quartz substrate. (b) Experimental spectroscopic ellipsometry spectra (solid lines) and best-fit curves (dots) for 4 ML CdSe CQWs at a 65° incidence angle. (c) Real part of the dielectric function for 2 to 7 ML CdSe CQWs, showing the variation with increasing monolayer thickness. (d) Imaginary part of the dielectric function for 2 to 7 ML CdSe CQWs, illustrating the evolution of optical transitions with layer thickness.

    In the SE analysis, the dispersion model describes the absorption characteristics of the sample by fitting the complex dielectric function. We employed the Tauc−Lorentz oscillator model, which captures the essential features of dispersive excitons. In this model, excitons are not excited at low energies, leading to a zero value for the imaginary part of the dielectric function at those energies. Additionally, phonon coupling between bright and dark excitons causes the absorption peak to adopt an asymmetric shape. Given the strong excitonic effects present in CdSe CQWs, we used a four-oscillator Tauc−Lorentz model to accurately describe the optical properties of CdSe CQWs from 2 to 7 ML. Details of this model are provided in the supplementary material.

    Fig. 3(b) and Fig. S3 show the experimental ellipsometry spectra (solid lines) and the best-fit spectra (dots) for the elliptical polarization parameters (Psi and Delta) across the range of CdSe films. The close agreement between the experimental and fitted curves demonstrates the accuracy of the model. Following the fitting procedure, the measured thickness and roughness values for the 2 to 7 ML CdSe films, as detailed in Table S2, closely matched the AFM data. During the fitting process, the mean squared error (MSE) was used to quantify the deviation between the experimental and fitted values, with MSE values for all samples remaining below 2, further validating the accuracy of the SE analysis.

    By fitting the SE spectra, we obtained both the real (ε1) and imaginary (ε2) parts of the dielectric function for the 2 to 7 ML CdSe CQWs, as shown in Fig. 3(c) and 3(d). Across the examined spectral range, both ε1 and ε2 show a clear layer-dependent behavior. The imaginary part, ε2, is associated with electronic transitions between the valence and conduction bands of CdSe CQWs. As the number of monolayers increases, the peaks of ε2 shift towards lower energies (redshift), indicating a reduction in the electron transition energy between the valence and conduction bands. This trend aligns with the observed decrease in the bandgap, as measured by the absorption spectra, as the monolayer thickness increases.

    The dielectric function and complex refractive index of a material are interrelated, and their relationship is expressed as follows[45]:

    ϵ=N2=ϵ1iϵ2=(n2k2)i2nk,where ε is the dielectric function of the material, ε1 and ε2 are the real and imaginary components of the dielectric function, respectively. N represents the complex refractive index, and n and k correspond to the refractive index and extinction coefficient. By using this relationship, the refractive index and extinction coefficient for the 2 to 7 ML CdSe CQW films were calculated and plotted as functions of energy, as shown in Fig. 4(a) and 4(b). Both parameters show a clear dependence on energy and layer thickness. For the refractive index, it increases with increasing energy, demonstrating normal dispersion. Additionally, as the number of monolayers increases, the refractive index also rises, which can be explained by the higher filling fraction of inorganic components in thicker films. This results in a greater dielectric contribution from atomic and molecular oscillations under photo-excitation.

    (Color online) (a) Refractive index and (b) extinction coefficient of 2 to 7 ML CdSe CQW films, demonstrating their layer-dependent optical properties. HH and LH represent the two prominent absorption peaks of CdSe CQW films. (c) Variation of the refractive index peak values from (a) with monolayer numbers. The inset illustrates the change in extinction coefficient peak values from (b) with monolayer numbers. (d) Variation of the normalized refractive index and normalized extinction coefficient (the inset) with monolayer number. (e) (Top) Deconvolution of the absorption spectrum for 4 ML CdSe CQWs, showing contributions from light-hole (LH) and heavy-hole (HH) excitons as well as free carriers. The exciton binding energy, estimated as 171 meV, is indicated. (Bottom) Absorption coefficient of 4 ML CdSe CQWs, with the positions of the LH and HH exciton lines highlighted. (f) Relationship between exciton binding energy and optical bandgap as a function of monolayer thickness in CdSe CQWs.

    Figure 4.(Color online) (a) Refractive index and (b) extinction coefficient of 2 to 7 ML CdSe CQW films, demonstrating their layer-dependent optical properties. HH and LH represent the two prominent absorption peaks of CdSe CQW films. (c) Variation of the refractive index peak values from (a) with monolayer numbers. The inset illustrates the change in extinction coefficient peak values from (b) with monolayer numbers. (d) Variation of the normalized refractive index and normalized extinction coefficient (the inset) with monolayer number. (e) (Top) Deconvolution of the absorption spectrum for 4 ML CdSe CQWs, showing contributions from light-hole (LH) and heavy-hole (HH) excitons as well as free carriers. The exciton binding energy, estimated as 171 meV, is indicated. (Bottom) Absorption coefficient of 4 ML CdSe CQWs, with the positions of the LH and HH exciton lines highlighted. (f) Relationship between exciton binding energy and optical bandgap as a function of monolayer thickness in CdSe CQWs.

    For the extinction coefficient, the transition peaks observed in the ellipsometry analysis correspond to the HH and LH transitions in the absorption spectra, supporting the consistency of the measurement results. As shown in Fig. 4(b), the extinction coefficient is also dependent on the number of monolayers. HH and LH represent the two prominent absorption peaks of CdSe CQW films. Specifically, there is a coupling effect between adjacent layers in the multilayer CdSe CQWs. With increasing monolayer numbers, the interlayer coupling becomes stronger, leading to more complex electron and hole behavior and a corresponding increase in the extinction coefficient across multiple energy bands.

    To clarify the relationships between the refractive index, extinction coefficient, and monolayer thickness, we extracted the peak values of these parameters from Fig. 4(a) and 4(b) and plotted them against the number of monolayers, as shown in Fig. 4(c). In thicker CdSe CQWs, a greater amount of material interacts with light, resulting in increased photon absorption and a corresponding rise in the extinction coefficient. However, beyond a certain thickness, the absorption levels off, even as the number of layers continues to increase. Fig. 4(d) presents the refractive index and extinction coefficient normalized by monolayer number. In contrast to the absolute values, thinner CQWs (with fewer monolayers) demonstrate higher normalized refractive indices and extinction coefficients, indicating stronger light−matter interactions per unit volume of material. We attribute this enhanced light−matter interaction in thinner CQWs to greater exciton binding energy, which arises from increased quantum confinement effects and reduced Coulombic screening in these structures.

    We calculated the absorption coefficients for the 2 to 7 ML CdSe CQWs using α = 4πk/λ[46]. Notably, the long-wavelength tail typical of light scattering in stacked CQWs is absent, leaving only the characteristic absorption peaks corresponding to HH and LH excitons [as shown in Fig. 4(e) for 4 ML CdSe CQWs, with spectra for other CQWs provided in the supplementary material]. To fit the absorption spectra for the 2 to 7 ML CdSe CQWs, we used a model that accounts for exciton localization by incorporating asymmetric broadening. The absorption coefficient is decomposed into a sum of exciton lines arising from HH and LH states, as well as transitions from free carriers. The detailed fitting process is described in the supplementary material. As illustrated in Fig. 4(e), the exciton binding energy Eb is the difference between the exciton energy Ex and the energy at which the free carrier transition line reaches half its high-energy limit. In 2D materials, the quantum confinement effect decreases with increasing monolayer thickness, resulting in a reduction in exciton binding energy, as shown in Fig. 4(f). The detailed fitting results are provided in Table S3. Compared with CdSe quantum wells, the exciton binding energies of CdSe quantum dots and CdSe nanorods exhibit significant differences, primarily attributed to their quantum confinement effect and dimensional differences. In quantum dots, the strong confinement in the three spatial dimensions localizes the electron and hole wave functions, thereby enhancing the coulomb interaction. This pronounced quantum confinement effect leads to a maximum exciton binding energy. In contrast, the one-dimensional extension in nanorods and the two-dimensional extension in quantum wells gradually weaken the coulomb interaction between electrons and holes, leading to a reduction in exciton binding energy. In summary, the differences in exciton binding energy among these three low-dimensional CdSe nanostructures are primarily due to the quantum confinement effect and coulomb interaction, and this variation has important significance for the design of optoelectronic devices and optical applications based on exciton properties.

    Conclusion

    This work provides a detailed investigation of the layer-dependent optical and dielectric properties of CdSe semiconductor colloidal quantum wells, shedding light on how key characteristics like bandgap, refractive index, and extinction coefficient evolve with monolayer thickness. As the monolayer number increases, we observe a pronounced decrease in bandgap, primarily driven by quantum confinement effects, alongside predictable shifts in refractive index and extinction coefficient. Through a rigorous analysis of the absorption spectra, considering exciton localization effects, we achieved a highly accurate representation of excitonic behavior in CdSe CQWs. This research contributes to the fundamental understanding of these materials and paves the way for their integration into a range of high-performance optoelectronic devices. The ability to tailor optical properties through precise control of monolayer thickness offers promising opportunities for applications such as quantum emitters, solar cells, and photodetectors, marking an important step forward in the practical use of 2D semiconductor colloidal quantum wells.

    References

    [1] B T Diroll, B Guzelturk, H Po et al. 2D Ⅱ–Ⅵ semiconductor nanoplatelets: From material synthesis to optoelectronic integration. Chem Rev, 123, 3543(2023).

    [2] D F Macias-Pinilla, J Planelles, I Mora-Seró et al. Comparison between trion and exciton electronic properties in CdSe and PbS nanoplatelets. J Phys Chem C, 125, 15614(2021).

    [3] R Scott, J Heckmann, A V Prudnikau et al. Directed emission of CdSe nanoplatelets originating from strongly anisotropic 2D electronic structure. Nat Nanotechnol, 12, 1155(2017).

    [4] I Tanghe, M Samoli, I Wagner et al. Optical gain and lasing from bulk cadmium sulfide nanocrystals through bandgap renormalization. Nat Nanotechnol, 18, 1423(2023).

    [5] Z T Zhang, Y T Thung, X X Chen et al. Study of complex optical constants of neat cadmium selenide nanoplatelets thin films by spectroscopic ellipsometry. J Phys Chem Lett, 12, 191(2021).

    [6] B T Diroll, R D Schaller. Reexamination of the giant oscillator strength effect in CdSe nanoplatelets. J Phys Chem C, 127, 4601(2023).

    [7] M A Ibrahem, M Waris, M R Miah et al. Orientation-dependent photoconductivity of quasi-2D nanocrystal self-assemblies: Face-down, edge-up versus randomly oriented quantum wells. Small, 20, 2401423(2024).

    [8] S Das, A Tripathi, A Parida et al. Visible light photodetectors based on hydrothermally synthesized Cd-Se-Te nanostructures. ACS Appl Electron Mater, 6, 6522(2024).

    [9] Y Zhang, W B Xiang, R Wang et al. Study of the mechanisms of the phonon bottleneck effect in CdSe/CdS core/shell quantum dots and nanoplatelets and their application in hot carrier multi-junction solar cells. Nanoscale Adv, 5, 5594(2023).

    [10] S X Liao, Z Z Yang, J D Lin et al. A hierarchical structure perovskite quantum dots film for laser-driven projection display. Adv Funct Mater, 33, 2210558(2023).

    [11] J Kim, J Roh, M Park et al. Recent advances and challenges of colloidal quantum dot light-emitting diodes for display applications. Adv Mater, 36, 2212220(2024).

    [12] K D Park, M A May, H X Leng et al. Tip-enhanced strong coupling spectroscopy, imaging, and control of a single quantum emitter. Sci Adv, 5, eaav5931(2019).

    [13] K Jo, E Marino, J Lynch et al. Direct nano-imaging of light-matter interactions in nanoscale excitonic emitters. Nat Commun, 14, 2649(2023).

    [14] E Ugur, A A Said, P Dally et al. Enhanced cation interaction in perovskites for efficient tandem solar cells with silicon. Science, 385, 533(2024).

    [15] D Kumar, H R Li, U K Das et al. Flexible solution-processable black-phosphorus-based optoelectronic memristive synapses for neuromorphic computing and artificial visual perception applications. Adv Mater, 35, 2300446(2023).

    [16] A Rani, A Verma, B C Yadav. Advancements in transition metal dichalcogenides (TMDCs) for self-powered photodetectors: Challenges, properties, and functionalization strategies. Mater Adv, 5, 3535(2024).

    [17] Y G Yao, Y K Zhu, A Hu et al. Temperature-regulated in-plane exciton dynamics in CdSe/CdSeS colloidal quantum well heterostructures. ACS Photonics, 10, 4052(2023).

    [18] A A Marder, J Cassidy, D Harankahage et al. CdS/CdSe/CdS spherical quantum wells with near-unity biexciton quantum yield for light-emitting-device applications. ACS Materials Lett, 5, 1411(2023).

    [19] G H Ba, Y M Yang, F Huang et al. Gradient alloy shell enabling colloidal quantum wells light-emitting diodes with efficiency exceeding 22%. Nano Lett, 24, 4454(2024).

    [20] N Taghipour, S Delikanli, S Shendre et al. Sub-single exciton optical gain threshold in colloidal semiconductor quantum wells with gradient alloy shelling. Nat Commun, 11, 3305(2020).

    [21] S Delikanli, G N Yu, A Yeltik et al. Ultrathin highly luminescent two-monolayer colloidal CdSe nanoplatelets. Adv Funct Mater, 29, 1901028(2019).

    [22] S Jana, R Martins, E Fortunato. Stacking-dependent electrical transport in a colloidal CdSe nanoplatelet thin-film transistor. Nano Lett, 22, 2780(2022).

    [23] A Sharma, M Sharma, K Gungor et al. Near-infrared-emitting five-monolayer thick copper-doped CdSe nanoplatelets. Adv Optical Mater, 7, 1900831(2019).

    [24] A H Khan, V Pinchetti, I Tanghe et al. Tunable and efficient red to near-infrared photoluminescence by synergistic exploitation of core and surface silver doping of CdSe nanoplatelets. Chem Mater, 31, 1450(2019).

    [25] T K Kormilina, S A Cherevkov, A V Fedorov et al. Cadmium chalcogenide nano-heteroplatelets: Creating advanced nanostructured materials by shell growth, substitution, and attachment. Small, 13, 1702300(2017).

    [26] Y L Li, L F Wang, D M Xiang et al. Dielectric and wavefunction engineering of electron spin lifetime in colloidal nanoplatelet heterostructures. Adv Sci, 11, 2306518(2024).

    [27] Y Altintas, K Gungor, Y Gao et al. Giant alloyed hot injection shells enable ultralow optical gain threshold in colloidal quantum wells. ACS Nano, 13, 10662(2019).

    [28] S Christodoulou, J I Climente, J Planelles et al. Chloride-induced thickness control in CdSe nanoplatelets. Nano Lett, 18, 6248(2018).

    [29] B T Diroll. Colloidal quantum wells for optoelectronic devices. J Mater Chem C, 8, 10628(2020).

    [30] M Ashry, S Fares. Radiation effect on the optical and electrical properties of CdSe(In)/p-Si heterojunction photovoltaic solar cells. J Semicond, 33, 102001(2012).

    [31] Y T Bao, X R Wang, S J Xu. Sub-bandgap refractive indexes and optical properties of Si-doped β-Ga2O3 semiconductor thin films. J Semicond, 43, 062802(2022).

    [32] C X She, I Fedin, D S Dolzhnikov et al. Low-threshold stimulated emission using colloidal quantum wells. Nano Lett, 14, 2772(2014).

    [33] Y Gao, M J Li, S Delikanli et al. Low-threshold lasing from colloidal CdSe/CdSeTe core/alloyed-crown type-II heteronanoplatelets. Nanoscale, 10, 9466(2018).

    [34] T Galle, D Spittel, N Weiß et al. Simultaneous ligand and cation exchange of colloidal CdSe nanoplatelets toward PbSe nanoplatelets for application in photodetectors. J Phys Chem Lett, 12, 5214(2021).

    [35] B T Diroll, R D Schaller. Intersubband relaxation in CdSe colloidal quantum wells. ACS Nano, 14, 12082(2020).

    [36] Y L Li. Measurement of the optical dielectric function of monolayer transition metal dichalcogenides: MoS2, MoSe2, WS2, and WSe2. Phys Rev B, 90, 205422(2014).

    [37] Y V Morozov, M Kuno. Optical constants and dynamic conductivities of single layer MoS2, MoSe2, and WSe2. Appl Phys Lett, 107, 083103(2015).

    [38] D B Hu, X X Yang, C Li et al. Probing optical anisotropy of nanometer-thin van der waals microcrystals by near-field imaging. Nat Commun, 8, 1471(2017).

    [39] X Z Chen, D B Hu, R Mescall et al. Modern scattering-type scanning near-field optical microscopy for advanced material research. Adv Mater, 31, 1804774(2019).

    [40] C L Wang, M J Ying, J Lian et al. Structural, optical and dielectric properties of (Co and Sm) co-implanting O-polar ZnO films on sapphire substrate. J Alloys Compd, 876, 160017(2021).

    [41] C L Wang, M J Ying, J Lian et al. Structural, optical and half-metallic properties of Mn and As co-implanted ZnO thin films. Appl Surf Sci, 575, 151703(2022).

    [42] C Y Ye, W Lin, J Zhou et al. Optical properties of InN studied by spectroscopic ellipsometry. J Semicond, 37, 102002(2016).

    [43] M Y Wei, J Lian, Y Zhang et al. Layer-dependent optical and dielectric properties of centimeter-scale PdSe2 films grown by chemical vapor deposition. NPJ 2D Mater Appl, 6, 1(2022).

    [44] H G Gu, B K Song, M S Fang et al. Layer-dependent dielectric and optical properties of centimeter-scale 2D WSe2: Evolution from a single layer to few layers. Nanoscale, 11, 22762(2019).

    [45] M L Zhao, Y J Shi, J Dai et al. Ellipsometric study of the complex optical constants of a CsPbBr3 perovskite thin film. J Mater Chem C, 6, 10450(2018).

    [46] F Ströhl, D L Wolfson, I S Opstad et al. Label-free superior contrast with c-band ultra-violet extinction microscopy. Light Sci Appl, 12, 56(2023).

    Chenlin Wang, Haixiao Zhao, Xian Zhao, Baoqing Sun, Jie Lian, Yuan Gao. Layer-dependent optical and dielectric properties of CdSe semiconductor colloidal quantum wells characterized by spectroscopic ellipsometry[J]. Journal of Semiconductors, 2025, 46(4): 042102
    Download Citation