Brought to you by:
Paper

Optical methods for determining thicknesses of few-layer graphene flakes

, , , , and

Published 22 November 2013 © 2013 IOP Publishing Ltd
, , Citation Wengen Ouyang et al 2013 Nanotechnology 24 505701 DOI 10.1088/0957-4484/24/50/505701

0957-4484/24/50/505701

Abstract

Optical microscopy (OM) methods have been commonly used as a convenient means for locating and identifying few-layer graphene (FLG) on SiO2/Si substrates. However, it is less clear how reliably optical images of FLG could be used to determine the sample thickness. In this work, various OM methods based on color differences and color contrasts are presented and their reliabilities are evaluated. Our analysis shows that these color-based OM methods depend sensitively on certain parameters of the measuring system, particularly the light source and the reference substrate. These parameters have usually been overlooked and less controlled in routine experiments. From evaluating the performance of these OM methods with both virtual and real FLG samples, we propose some practical guidelines for minimizing the impact of these less-controlled experimental parameters and provide a user-friendly MATLAB script for facilitating the implementation.

Export citation and abstract BibTeX RIS

1. Introduction

Graphene, a one-atom-thick layer of carbon, has become a wonder material for both fundamental scientific studies and technological applications since its first isolation by mechanical exfoliation in 2004 [1]. Due to its special electronic band structure, graphene has many unique and amazing properties such as the quantum Hall effect [2] and high electron mobility at room temperature [1, 3]. In addition to monolayer graphene, few-layer graphene (FLG) samples are also intriguing because they possess many useful properties that are distinct from those of monolayer graphene [4]. Many devices based on monolayer and few-layer graphene have been proposed, such as spin qubits [5], transistors [6], transparent electrodes [7], and ultratough paper [8]. Currently mechanical cleavage is still one of the most convenient methods for producing high-quality graphene flakes [9]. For samples prepared with the mechanical exfoliation method, monolayer graphene flakes are usually very few and sparsely distributed on the substrate surface together with a large number of relatively thick graphite flakes. To facilitate the study and application of graphene, it is beneficial to find a method for identifying the thickness of these graphite flakes quickly and reliably. Atomic force microscopy (AFM) has been used to measure the thickness of these ultrathin materials [10]. However, the procedures are time-consuming and can potentially induce unexpected structural defects [9, 11]. Recently, Raman spectroscopy has emerged as a powerful technique for characterizing the number of layers of graphene samples [11, 12]. However, its application is limited by the fact that the distinction of Raman spectra of few-layer graphene flakes becomes nonobvious when the number of layers exceeds about 4 [9, 12]. Although an improved method has been proposed for determining thicker graphene samples [12], this method still needs a dedicated Raman system and generally only provides results at discrete points rather than a continuous field.

On the basis of the different reflectivities of graphene flakes upon incidence of monochromatic light, researchers have used the reflectivity contrast to identify the layer number of FLG [9, 13, 14]. However, as pointed out by Wang et al [15], these methods require some dedicated instruments to measure the reflectivity. Recently, the phenomenon of FLG exhibiting color variation with thickness has led to the development of more convenient OM methods based on the apparent color of the FLG samples. Gao et al [16] used the total color difference to identify the thickness of FLG on Al2O3/Si substrate; Wang et al [17] and Jung et al [18] used RGB color contrast for FLG on SiO2/Si and Si3N4/Si substrates. Since the apparent color exhibited by the samples is generally affected by the spectral power distribution of the incident light source, how the variation of light sources affects the thickness determination has become a critical issue for reliable application of these color-based OM methods.

The present work reviews the basic theory for calculating the apparent color of multilayer structures. Under a unified theoretical framework, the OM methods based on various color differences and color contrasts (defined in the international standard color spaces [19, 20]) are presented and their reliability is discussed. Our analysis shows that these color-based OM methods are generally very sensitive to the parameters of the measuring system, particularly the light source and the reference substrate. This is an aspect that has often been overlooked in previous studies. From evaluating the performance of these OM methods with both virtual and real FLG samples, we propose some practical guidelines for minimizing the impact of these less-controlled experimental parameters and provide a user-friendly MATLAB script for facilitating the implementation.

2. Theory for calculating the color of multilayer film structures

2.1. Reflectance of a multilayer structure

Figure 1 shows a schematic diagram of an optical path when an incident light passes through an m-layer film structure. We denote as R(λ,θi;n,d) the global reflectance of the m-layer structure and as T(λ,θi;n,d) its transmittance, both of which are functions of the incident light wavelength λ, the incident angle θi, the complex refractive index vector n = (ni,n1,n2,...,nm,ns)T, and the thickness vector d = (di,d1,d2,...,dm,ds)T of the structure. In a typical experimental setup, the incident medium is usually air with a refractive index of ni = 1.0003, and the thickness of the incident medium, di, and the substrate, ds, are usually assumed to be infinite.

Figure 1.

Figure 1. A schematic diagram of an m-layer thin film structure. nk,dk, θk are the complex refractive index, thickness and refractive angle in the kth layer. The media are labeled in sequence with index 'i' corresponding to the incident medium and 's' representing the substrate.

Standard image High-resolution image

By following the matrix method of multilayer theory, we calculated the reflectance and transmittance of the m-layer structure as follows [21, 22]:

Equation (1)

Equation (2)

where C0 = cosθi/cosθs for p-polarized light and C0 = 1 for s-polarized light, while ${E}_{k}^{+}$ is the amplitude of the electric field vector of the wave traveling in the direction of incidence in the kth layer and ${E}_{k}^{-}$ is that of the negative-going wavevector. Apparently, both reflectance and transmittance are functions of the wavelength and incident angle of the incident light for a given multilayer structure.

In a real optical system, the incident angle of light is not strictly zero, but has a certain distribution [16, 18, 23]. This distribution of incident range is usually described via the numerical aperture (NA) [24] of the objective lens. The maximum incident angle θmax can be described by the equation [16, 18] NA = nisinθmax, where ni is the refraction index of the medium between samples and the objective lens; ni = 1.0003 when the medium is air. Assuming that the intensity of light at different incident angles follows a Gaussian distribution, we can calculate the modified total reflectivity of light using the following integration [16]:

Equation (3)

where ϕ(θi,0,θmax/3) is the Gaussian distribution with the mean of 0 and standard deviation of θmax/3, where θmax = arcsin(NA/ni).

2.2. Apparent color of a multilayer structure

Once the reflectance of a multilayer structure is known, its color can be calculated by a standard procedure [25]. According to the international standards, color can be described in the International Commission on Illumination (usually abbreviated as CIE, for its French name) color space by using different sets of parameters (or tristimuli), e.g. XYZ parameters, RGB parameters, or L*a*b* parameters [25] (the corresponding color spaces are called CIE XYZ, CIE RGB and CIELAB color spaces respectively). For a given light source with a normalized spectral power distribution (SPD) P(λ) in the wavelength space, the CIE XYZ tristimulus components can be obtained as

Equation (4)

where R(λ,θi;n,d) is the reflectance of the multilayer structure, $\bar {x}(\lambda ),\bar {y}(\lambda )$ and $\bar {z}(\lambda )$ are the CIE color-matching functions, to account for the response of the eyes [20]. For a color to be captured and displayed, its XYZ tristimulus needs to be converted to a device-dependent color scheme (such as RGB), which is accomplished by multiplying the XYZ vector by a transformation matrix ${\mathbf{M}}_{\mathrm{d}}^{-1}$ as indicated in equation (5):

Equation (5)

Conversely, the RGB parameters can be converted to the XYZ components by multiplying by Md. Because the transformation matrix is generated using the phosphor chromaticity coordinates and the reference white point of the recording device (more details can be found in [2527]), Md is device dependent; thereby the RGB parameters are also device dependent. Correctly estimating Md requires obtaining detailed knowledge about the optical recording devices, which is usually very difficult if not infeasible. For a standard RGB (sRGB) system, the illuminant D65 is usually adopted as the virtual reference light source [28] and the Md can be calculated from a standard procedure [2527].

Because the CIELAB color space has several distinct advantages [16, 29], e.g. its gamut is relatively uniform and it has been extensively studied with several standardized definitions of color difference, the XYZ parameters are often expressed in the CIELAB color space [29, 30] using L*,a*,b* parameters as follows:

Equation (6)

where L* represents the lightness, a* and b* represent the chromaticity, X0,Y0 and Z0 are the CIE XYZ tristimulus values of the standard illuminant and Φ(t) is given by

On the basis of the CIE color spaces, we can now quantitatively define color differences/contrasts between two colors. In this work, we will consider two standard color differences in the CIELAB color space and two color contrasts in the CIE RGB color space, whose definitions are given in the following.

The CIEDE76 color difference [16, 29] is defined as

Equation (7)

where ΔL*,Δa* and Δb* represent the differences between the L*a*b* color parameters of the sample and those of the reference substrate. It is noted that the CIEDE76 color difference is essentially the same as the total color difference adopted by Gao et al in their OM method [16].

The CIEDE2000 color difference [31] is defined as

Equation (8)

where ΔL',ΔC' and ΔH' represent the modified differences between the L,C (chroma) and H (hue) parameters of the sample and those of the reference substrate. SL,SC and SH are compensation weight functions for lightness, chroma and hue respectively; RT is a rotation term for the hue and kL,kC and kH are parametric weighting factors (more details can be found in [31, 32]).

In addition to color differences defined in the CIELAB space, color contrasts defined in the CIE RGB space were also proposed in prior OM methods [17, 18]. For example, Jung et al [18] used the color contrast defined as

Equation (9)

where (R0,G0,B0) and (R,G,B) are the RGB components of the reference substrate and those of the sample respectively; while the method by Wang et al [17] was based on a color contrast involving the green (G) channel alone:

Equation (10)

From equations (3)–(10), it is evident that the color differences/contrasts generally depend not only on R(λ,θi;n,d), i.e. the multilayer structure, but also on P(λ) and Md, i.e. the characteristics of the light source and the optical recording system. In a typical experimental setup, the precise SPD of the incident light source and the characteristics of the optical recording system are usually not known. How color differences/contrasts change with the light source and optical system for an OM method will determine its robustness. In section 3, we will explore the influences of these uncertain factors on color-based OM methods by performing virtual and real experiments.

3. Color-based methods for thickness determination

As discussed in section 2, the color of a multilayer system depends on its structure. Therefore, by comparing the color between a graphene sample and a reference substrate, one can devise an optical method for deducing the sample structure (thickness). In this section, we will first present detailed procedures for implementing these optical methods. Then we will perform computational experiments on virtual images generated from known multilayer structures in order to evaluate the intrinsic performance of each method. Finally, these methods will be implemented for a real experiment and their sensitivities to various experimental parameters will be compared and discussed.

Figure 2 shows a flowchart highlighting the general procedures of a color-based optical method for identifying the thickness of a few-layer graphene sample from experimental images. Firstly, the RGB components of the color images from experiments are extracted. The experimental color contrasts/differences between the graphene sample regions and the reference substrate region are calculated from their RGB components (or L*a*b* components). In the meantime, if the light source, the characteristics of optical recording system and the structure of the reference substrate are known, the theoretical color differences/contrasts between a given few-layer graphene sample (with its thickness as a variable) and the reference substrate can be computed. Finally, the thickness of the experimental sample can be determined by iteratively varying the thickness variable until the theoretical color differences/contrasts match the experimental values. To facilitate the implementation, we have coded the above procedures into a user-friendly and publicly accessible MATLAB script (more details can be found in SI, available at stacks.iop.org/Nano/24/505701/mmedia).

Figure 2.

Figure 2. A flow chart highlighting the procedures of a color-based OM method for identifying the thickness of an FLG sample.

Standard image High-resolution image

3.1. Theoretical calculation: color differences/contrasts versus sample thickness

As indicated in figure 3, the most critical step in determining the sample thickness is theoretically computing the dependence of the color differences/contrasts on the sample thickness. Since SiO2/Si is the most common substrate for graphene applications [13, 14, 1618, 24], we will focus our discussions on graphene/SiO2/Si systems in this work; however, it should be noted that the findings of this work are generally applicable for other multilayer structures. For a graphene/SiO2/Si multilayer structure, the system consists of four components: air (incident medium), the few-layer graphene flake (sample), the SiO2 layer (middle layer), and Si (substrate). The complex refractive indices of Si and SiO2 are characterized by λ-dependent functions as suggested in [33]. The complex refractive index of FLG is more subtle and it is still an ongoing research subject [9, 13, 34]. Different refractive index values have been deduced indirectly for few-layer graphene samples [9, 13, 35]. In this work, we chose a constant value of the refractive index, as suggested by Blake et al [13]. As will be shown in a later attempt to calibrate the OM method, when we let the refractive index of FLG be a free variable to fit our experimental data, the optimal value of the refractive index is closer to that of bulk graphite, which is consistent with the results of Blake et al [13] for a similar system.

Figure 3.

Figure 3. Variation of calculated color contrasts/differences (a) ΔEG, (b) ΔERGB, (c) ΔE76, (d) ΔE00, with the number of graphene layers under illumination for five standard light sources (i.e. F4, F6, F12, E and D65) for a graphene/SiO2(285 nm)/Si structure. The thickness of each graphene layer was assumed to be 0.34 nm throughout this paper.

Standard image High-resolution image

For a graphene/SiO2(285 nm)/Si structure illuminated with five standard illuminants, the theoretical variations of the color differences/contrasts with the number of graphene layers are shown in figures 3(a)–(d). In order for a sample thickness to be determined unambiguously, the color differences/contrasts have to vary monotonically with the thickness. For example, if the thickness of an unknown graphene sample is less than ten layers, then its value can be determined unambiguously from all four color differences/contrasts, as suggested by figure 3. However, if the thickness of an unknown sample can be more than ten layers, one has to carefully specify the possible range of graphene thickness to ensure that only one thickness value can be determined ambiguously. Variations of color differences/contrasts with a larger range of graphene thickness can be found in the supplementary information (SI figure S1, available at stacks.iop.org/Nano/24/505701/mmedia).

As shown in figure 3, the variation of the color differences/contrasts changes when different incident light sources are considered. This dependence of the color differences/contrasts on the light source is relatively weak when the sample thickness is small (less than five layers) and it becomes more significant for thicker samples. For example, in figure 3(d), the color difference of 10 corresponds to a sample thickness of five layers if we choose light source F4; but the thickness will become ∼9 layers if we choose light source D65. Ideally, an exact light source should be chosen to ensure a reliable calculation when using color-based optical methods.

We also explored the dependence of color differences/contrasts on the NAs of the optical system (see SI figures 2 and 3, available at stacks.iop.org/Nano/24/505701/mmedia). Generally, the influence of the NAs is much smaller than that of the light sources. For simplicity, we will focus the following discussions on the cases where the NA is assumed to be zero.

In addition to the light source, color differences/contrasts should depend on the reference substrate (i.e. the thickness of the SiO2 layer). In order to investigate the influence of the reference substrate, we calculated the color differences/contrasts between one-, two-, three- and four-layer graphene and the reference substrate as a function of the SiO2 layer thickness. The results, assuming an incident light source of D65, are as shown in figure 4.

Figure 4.

Figure 4. Color contrasts/differences between the FLG and reference substrate as functions of the thickness of the SiO2 layer: (a) ΔERGB, (b) ΔEG, (c) ΔE76, (d) ΔE00. All light sources were assumed to be D65 illuminant. The red dots indicate the local peaks and the values are the corresponding SiO2 layer thicknesses.

Standard image High-resolution image

As clearly shown in figure 4, the color differences/contrasts are very sensitive to the thickness of the SiO2 layer. For example, the color contrast ΔEG between one-, two-, three- and four-layer graphene and the reference substrate only peaks when the SiO2 layer has certain thicknesses, and decays quickly becoming almost indistinguishable elsewhere. In order to achieve better optical visibility, one should choose an SiO2 substrate with a proper thickness so that the values of the color differences/contrasts are maximized. Previously, people have used the SiO2(285 nm)/Si substrate to enhance the reflectivity contrast of graphene for the given monochromatic light source [13, 14], which was similar to the present finding. The structure of the SiO2 layer affects not only the values of the color differences/contrasts but also their variations with the graphene thickness. For example, the variation trends shown in figure 3 would be drastically different if the SiO2 layer thickness (which was chosen to be 285 nm) was changed (see SI figure S4, available at stacks.iop.org/Nano/24/505701/mmedia). Because of the strong dependence of the color differences/contrasts on the thickness of the SiO2 layer, one should accurately measure the thickness of the SiO2 layer when implementing color-based optical methods.

The above theoretical discussions provide a few strategies for optimizing the performance of color-based optical methods: (a) choosing a color difference/contrast to ensure that its value varies monotonically with the sample thickness; (b) choosing a proper thickness of the SiO2 layer to enhance the values of the chosen color difference/contrast; (c) measuring the SPD of the light source (if possible) and the thickness of the SiO2 layer accurately.

3.2. Comparison of various optical methods: virtual experiments

To investigate the intrinsic performance of each optical method, we carried out a set of virtual experiments in this section. First, we generated a series of images (containing sRGB information) of virtual graphene samples using a standard illuminant, D65. These graphene samples were assumed to be on an SiO2(285 nm)/Si substrate with their thicknesses ranging from one layer to ten layers. After obtaining the virtual images, we used four optical methods based on the color differences/contrasts defined in section 2 to estimate the number of graphene layers. The light source and the thickness of the SiO2 layer were intentionally varied during the calculation; and their effects were explored by comparing the estimated thicknesses of the graphene samples with the initial 'true' values.

Figures 5(a)–(d) show the comparison between the calculated graphene thicknesses obtained from four optical methods and their initial 'true' thickness values. In each of these figures, five sets of data obtained assuming five different standard illuminants (i.e. F4, F6, F10, E and D65) are presented. From figure 5, one can easily find that when the original illuminant (D65) is assumed, all four methods could correctly calculate (recover) the graphene thickness. However, if other illuminants are assumed during the calculation, the estimated graphene thickness can deviate from the true values. This deviation depends on the choice of the color differences/contrasts and generally becomes more significant for thicker samples. According to the results shown in figures 5(a)–(d), the optical method based on color difference ΔE00 seems to be less affected by the choice of illuminants.

Figure 5.

Figure 5. Comparison of calculated and intrinsic pre-defined graphene thicknesses of virtual FLG samples, assuming five standard illuminants using OM methods based on (a), ΔEG, (b) ΔERGB, (c) ΔE76, (d) ΔE00. The thickness of SiO2 was set to be 285 nm.

Standard image High-resolution image

Like figures 5(a)–(d), 6(a)–(d) also illustrate the comparison between the calculated graphene thicknesses obtained from four optical methods and their initial 'true' thickness values. However, in each plot of figures 6(a)–(d), three sets of data obtained assuming three different thicknesses of the SiO2 layer (i.e. 270, 285 and 300 nm) are presented. When the original thickness of the SiO2 layer (285 nm) is considered, all four methods could correctly calculate (recover) the graphene thickness. However, if different SiO2 layer thicknesses (even only 5% off the original value) are assumed, the calculated graphene thickness can deviate substantially from the true values, as shown in figures 6(a)–(d). This high sensitivity of optical methods to the SiO2 layer thickness is consistent with the theoretical analysis given in figure 4. Therefore, the accurate measurement of the thickness of the SiO2 layer is a prerequisite for the implementation of color-based optical methods.

Figure 6.

Figure 6. Comparison of calculated and intrinsic pre-defined graphene thicknesses of virtual FLG samples, assuming three thicknesses of the SiO2 layer, using OM methods based on (a), ΔEG, (b) ΔERGB, (c) ΔE76, (d) ΔE00. The light source was set to be D65.

Standard image High-resolution image

3.3. Comparison of various optical methods: real experiments

In this section, we will use the optical methods to estimate the thickness of the graphene samples in a real experimental setup. The graphene samples were produced from Kish graphite by mechanical exfoliation [10] and then transferred to SiO2/Si substrate. The thickness of the SiO2 was found to be 299.3 ± 1.7 nm, as independently measured by ellipsometry and further confirmed by scanning electron microscopy. The optical images of the graphene samples were obtained with an Olympus BX-51 microscope under two different light sources (xenon light, Olympus AH2-RH, and halogen light, AX LH100).

Figures 7(a) and (b) are the optical images of FLG recorded under xenon light and halogen light respectively. We can see that the color varies slightly for different regions of graphene, indicating changes in sample thickness. Figure 7(c) shows a Raman G band intensity image (the brightness corresponds to the G peak intensity) obtained by Raman microscopy (NTEGRA Spectra, NT-MDT, excitation laser wavelength 532 nm) of the same selected area. From the intensity contrast and the detailed Raman spectra (figure 7(f)), the thicknesses of different regions of graphene were determined and labeled correspondingly in figures 7(a)–(c). Using the optical method based on the CIEDE2000 color difference and assuming the ideal sRGB scheme for the optical recording system, we calculated the thicknesses of different regions and compared them with the true values obtained by Raman microscopy. The results are shown in figures 7(d) and (e), based on the images collected under xenon light and halogen light respectively. Because the SPD of the real incident light was not known, we can use the known real thickness of SiO2 to find an effective light source (details can be found in SI, available at stacks.iop.org/Nano/24/505701/mmedia). Using the 'calibrated' or 'gauged' effective light source, the thicknesses of both the graphene and the SiO2 substrate can be correctly recovered, as shown by the filled circles in figures 7(d) and (e). Theoretically, this effective light source once calibrated should also work for other samples if the light source and the recording system are kept the same. To validate this hypothesis, we prepared another FLG sample and used the same effective light source to estimate the sample thickness. The results turned out to be accurate and consistent with independent Raman measurements (see SI for more details, available at stacks.iop.org/Nano/24/505701/mmedia).

Figure 7.

Figure 7. ((a), (b)) Raw optical images of the same FLG samples on an SiO2/Si substrate under illumination with a xenon light and a halogen light respectively; (c) Raman G band intensity image of (a) and (b); ((d), (e)) thicknesses of FLG samples identified by the OM method using the optical images shown in (a) and (b) respectively; (f) typical Raman spectra of monolayer to four-layer graphene samples; for the OM method, the thicknesses of both the SiO2 and the graphene were deduced using the CIEDE2000 color difference.

Standard image High-resolution image

4. Conclusions

Four color-based optical methods for identifying layer numbers of FLG samples were presented under a unified theoretical framework. The theoretical analysis suggested that these optical methods are generally sensitive to the experimental parameters, e.g. the spectral power distribution of the incident light, the characteristics of the optical recording system and the structure of the reference substrate. However, these parameters are usually less controlled in experiments and their effects have not been systematically studied. From evaluating the performance of these methods in both virtual and real experiments, we proposed the following practical guidelines for minimizing the impact of these commonly less-controlled experimental parameters: (a) an SiO2 layer with a proper thickness should be chosen to enhance the color differences/contrasts; (b) the thickness of the SiO2 layer can be estimated by the same optical method to offset the impact of different optical recording systems; (c) an FLG sample with a known structure can be used to find an effective illuminant for best approximating the real light source and minimizing the distortion from the recording system. Because the color differences/contrasts depend on these experimental parameters in a complicated manner, it is unlikely that one will find a specific solution (or a specific combination of experimental parameters) that works universally for all systems. Instead, we built a user-friendly script including all the optical methods discussed in order to provide a convenient and nondestructive means for whole-field thickness identification of two-dimensional materials.

Acknowledgments

This work was supported by the National Key Basic Research Program of China (2013CB934200 and 2013CB933003), the NSFC through Grant Nos 11272177, 10832005, and the Thousand Young Talents Program (20121770071). WGO would like to thank Drs Ze Liu and Tongbiao Wang and Professor Yao Cheng for discussions. QL would like to thank Professor R W Carpick for reading the manuscript and giving constructive comments. XZL would like to acknowledge the support from the National Science Foundation (NSF) under awards NSF/MRSEC (No. DMR-050020) & NSF/ENG (CMMI-08001541068741). Use of the University of Pennsylvania Nano/Bio Interface Center and the Carpick Group Nanotribology Lab instrumentation is acknowledged. Additionally, this research was partially supported by the Nano/Bio Interface Center through the National Science Foundation, NSEC DMR08-32802.

Please wait… references are loading.
10.1088/0957-4484/24/50/505701