Open Access
1 January 2010 Discrimination analysis of human lung cancer cells associated with histological type and malignancy using Raman spectroscopy
Yusuke Oshima, Hideyuki Shinzawa, Tatsuji Takenaka, Chie Furihata, Hidetoshi Sato
Author Affiliations +
Abstract
The Raman spectroscopic technique enables the observation of intracellular molecules without fixation or labeling procedures in situ. Raman spectroscopy is a promising technology for diagnosing cancers-especially lung cancer, one of the most common cancers in humans-and other diseases. The purpose of this study was to find an effective marker for the identification of cancer cells and their malignancy using Raman spectroscopy. We demonstrate a classification of cultured human lung cancer cells using Raman spectroscopy, principal component analysis (PCA), and linear discrimination analysis (LDA). Raman spectra of single, normal lung cells, along with four cancer cells with different pathological types, were successfully obtained with an excitation laser at 532 nm. The strong appearance of bands due to cytochrome c (cyt-c) indicates that spectra are resonant and enhanced via the Q-band near 550 nm with excitation light. The PCA loading plot suggests a large contribution of cyt-c in discriminating normal cells from cancer cells. The PCA results reflect the nature of the original cancer, such as its histological type and malignancy. The five cells were successfully discriminated by the LDA.

1.

Introduction

Lung cancer is the leading cause of cancer-related death worldwide. Early detection and appropriate treatment for cancerous lesions of patients can reduce the risk of death and improve the quality of life (QOL). Histopathological diagnosis of a lesion is necessary for physicians to decide on a therapeutic plan for a patient. Lung cancer is generally classified into two categories: small cell carcinoma (SC), and non-small cell carcinoma. This information reflects the characteristic of the tumor: SC is usually more aggressive with poorer differentiation potential than non-small cell carcinoma. Cancerous tissue of non-small lung cancer is further classified into three major types according to histological tissue pattern: adenocarcinoma (AD), squamous cell carcinoma (SQ), and large cell carcinoma (LC).1, 2, 3 Information on tissue morphology and cancer cell type is important for confirming a diagnosis. Raman spectroscopy is a type of vibrational spectroscopy that provides useful information for estimating the molecular composition of a sample. If the molecular composition changes in cancer tissues, its Raman spectrum must reflect those changes.

Raman spectroscopy is an especially powerful, noninvasive analysis tool for tissue and cells for two major reasons.4, 5, 6 One advantage is that the bands that form due to water do not overlap with the fingerprint region. This is a large advantage over infrared and near-infrared (NIR) spectroscopy, because biological samples are usually rich in water content. The other advantage is that Raman spectroscopy uses light in the visible and NIR regions, which permits the use of highly sensitive and useful optical devices in the development of Raman instruments, such as optical fiber and charge-coupled device (CCD) detectors. Haka demonstrated Raman spectroscopy in the diagnosis of breast cancer in vivo.7 They differentiated breast tissue into four pathological types—fibroadenoma, infiltrating carcinoma, fibrocystic change (not malignant), and normal—by using a model fitting the basic Raman spectra normalized to the sum total contribution of one. They also demonstrated Raman spectroscopy for intraoperative margin assessment of breast cancer in vivo.8 Huang measured bronchial tissue specimens, normal cells, squamous cell carcinoma, and adenocarcinoma. They found that differences between normal and tumor tissues seemed to arise from protein and lipid conformational changes and concentration of nucleic acid.9 Min demonstrated that 1064-nm -excitation NIR Raman spectroscopy could be employed to collect Raman spectra from fresh human lung tissue, which was found to emit strong fluorescence-interfering Raman signals under visible excitation.10 Yamazaki observed the amide I band on the Raman spectra, which is characteristic of cancerous tissue. They found that the amide I band was assignable to collagen-like material on the basis of their previous research, in which they used formalin-fixed tissue.11 Magee evaluated Raman spectroscopy in the diagnosis of lung cancer using a Raman mini-probe during bronchoscopy ex vivo.12 These reports have revealed that Raman spectroscopy has a high enough potential for the practical diagnosis of various types of cancer.

The identification of leukemia cells with Raman spectroscopy has been reported by Chan 13, 14 They measured the Raman spectra of T- and B-lymphocyte cells, and classified them into normal and cancerous groups by principal component analysis (PCA). The results indicated that differences in DNA and protein were responsible for discrimination via PCA. Changes in gene expression and enzyme activity in metabolic cycles are important signals for understanding the mechanism of cancer occurrence.1, 2, 3, 15, 16 Swain demonstrated cell cycle analysis with microscopic Raman spectroscopy.17 Notingher reported that, in Raman spectral dynamics of differentiation in embryonic stem (ES) cells, Raman bands of RNA contribute to distinguishing between differentiated and undifferentiated states.18

In the present study, Raman spectroscopy was applied to the analysis of differences among cultured cells of one normal cell and four lung cancer cell lines. Raman measurement was carried out directly on a culture dish with minimal treatment. Multivariable analyses were performed to discriminate among cells and to investigate the origin of differences. An excitation wavelength of 532nm was employed instead of 785 or 633nm . Although it has been reported that visible excitation has the disadvantages of photodegradation and fluorescence,19, 20 the 532-nm line provides resonance enhancement of cytochrome c (cyt- c ), which cannot be observed with excitation wavelengths of 785 and 633nm . Our final purpose was the diagnosis of cancer using Raman spectroscopy. Spectra for tissue included information about the cancer cells and the extracellular matrix. We assumed that the composition and property of the extracellular matrix depended on the individual history of the disease, which can make a clear diagnosis difficult with Raman spectroscopy. Diagnosis can be easier if it is possible to obtain information from the cells alone. Hence, the present study focuses on cancer cells and demonstrates the performance of Raman spectroscopy for the diagnosis and discrimination of the single cell.

2.

Materials and Methods

2.1.

Cell Cultures

All cell lines were purchased from the Japanese Collection of Research Bioresources (JCRB) cell bank. Cell lines included one normal lung cell line [MRC-5 (JCRB9008)] and four cancerous strains [RERF-LC-MS (JCRB0081), EBC-1 (JCRB0820), Lu-65 (JCRB0079), and RERF-LC-MA (JCRB0812)]. Each cancerous cell strain was of a different histological type (Table 1 ): AD for RERF-LC-MS, SQ for EBC-1, LC for Lu-65, and SC for RERF-LC-MA. The cell lines MRC-5, RERS-LC-MS, EBC-1, and RERS-LC-MA were seeded into Eagle’s minimum essential medium (51200-038; Invitrogen Corp., Carlsbad, California, USA) with 1.0% (v/v) nonessential amino acids (11140-050), 292mgL L-glutamine (21051-024), and 10% (v/v) heat-inactivated fetal bovine serum (10082-147). Lu-65 was seeded into the RPMI1640 medium (11835-030) with nonessential amino acids and fetal bovine serum in the same manner. The cells were maintained at 37°C with 5% CO2 within six days on a synapse fine-view dish (SF-S-D27; Fine Plus International Ltd., Japan ), the bottom of which was made of silica glass for spectroscopic measurement. For cell viability tests, a Trypan blue dye method [0.4% (w/v), 15250-061; Invitrogen] was used a few minutes after the Raman measurement.

Table 1

List of human lung cell lines used in Raman measurement.

Name of cell lineCancer/NormalHistological typeMalignancy
MRC-5Normal
RERF-LC-MSCancerAdenocarcinomaLow-Med
EBC-1CancerSquamous cell carcinomaLow-Med
Lu-65CancerLarge cell carcinomaHigh
RERF-LC-MACancerSmall cell carcinomaHigh

2.2.

Raman Measurement

Raman spectra were obtained using a laser Raman microscope (in Via Raman microscope; Renishaw plc, UK). An excitation laser (Nd: YAG SHG) of 532nm with power of 50mW was focused on a single cell via a 50× objective lens with numerical aperture (NA)=0.75 . Raman scattering light from the sample was collected using the same objective lens and detected by a Peltier-cooled CCD ( 0.5inch , 576×384pixels ) with a grating (1800linesmm) for the visible region and a 60-s exposure time. The spectral resolution was less than 2cm1 , and the diameter of the laser spot at the focus point was less than 1μm . The culture medium was removed, and the cells were washed twice with a buffer solution (Hanks’ balanced salt solution, 14175-095; Invitrogen, Carlsbad, California, USA) before the Raman measurement. Five Raman spectra were randomly taken in each dish. Background spectra were obtained by focusing on a vacant region on each dish where no cell existed. Raman spectra measurements with excitation wavelengths of 633nm and 785nm were obtained using the Nanofinder® (Tokyo Instruments, Inc., Japan) equipped with a 600linesmm holographic grating blazed at 750nm . For analysis of the concentration of cyt- c , Raman measurements of 20 cells each of MRC-5 and RERF-LC-MS were carried out with a 60× water-immersion objective lens (NA=1.20) , and exposure time was 30s . Purified cyt- c from bovine heart (c2037; Sigma, Japan) was solved into distilled water as an oxidized cyt- c solution. A reduced cyt- c solution was prepared by adding bisulfate mix (QIAGEN, Japan) into the cyt- c solution.

2.3.

Data Analysis

Raman spectral data analysis was performed with the commercially available GRAMS software (Galactic Industries Corp., Salem, New Hampshire, USA). The background spectrum was subtracted from the raw Raman spectrum, and the intensity of individual Raman spectra was normalized with a band of 1003cm1 of phenylalanine. To verify the concentration of cyt- c , we used the Raman spectra of 20 cells each of MRC-5 and RERF-LC-MS. The averaged intensity of the band at 1128cm1 relative to the band at 1003cm1 and its standard error in each cell strain was calculated, and its significance in difference was confirmed with Student’s t -test. For chemometrics analysis, a Raman spectral dataset was prepared for PCA and linear discrimination analysis (LDA). Sirius 7.1 software (Pattern Recognition Systems, Norway) was employed for PCA. Before analysis, the data were treated with mean-centering, and a classification model was constructed with two principal components (PCs). The third component in all models had minimal or no important information when judged by the loading weight of PC3. Two-dimensional plots showed scores for PC1 ( x axis) and PC2 ( y axis). LDA was calculated by the SPSS Statistics 17.0 software (SPSS Japan Inc., Japan). We used a predictive model composed of the discrimination function based on the linear combination of predictor variables that provide the best discrimination among cell strains. Independent variables used to build the discrimination model were selected according to the PCA loading weight. The LDA model was also built with the independent variables selected by the step-wise method (which obtains independent variables automatically) for optimal results. All predictive models were cross-validated by the leave-one-out method to evaluate its accuracy.

3.

Results and Discussion

3.1.

Raman Spectroscopy

Microscopic Raman spectroscopy has been performed for biochemical characterization of single living cells. In our study, with Raman spectroscopy, we succeeded in obtaining high signal-to-noise Raman spectra of cells without fixation. There was no interference of fluorescence from the dish, the buffer, or the cell itself. An excitation laser was focused within an adherent cell. Raman signals from the substrate of the culture dish (made of quartz) appeared on the raw Raman spectra (data not shown). The quartz had broad Raman bands near 795, 1060, and 1200cm1 . The background spectra, obtained from the vacant space of each cell culture dish, had Raman bands arising from the quartz window and a band at 1640cm1 due to water. Accordingly, the background spectrum was subtracted from the sample spectra to remove unwanted Raman bands.21 Figure 1 shows the Raman spectra of the following cell lines: (a) MRC-5 (normal cell line), (b) RERF-LC-MS (AD cell line), (c) EBC-1 (SQ cell line), (d) Lu-65 (LC cell line), and (e) RERF-LC-MA (SC cell line). The spectra in Fig. 1 included the mean spectra of five measurements in each culture dish to eliminate the effects of site dependency and cell cycles. The strong band at 1659cm1 seemed to consist of the amide I band of protein and the C=C stretching band of lipids. The features at 1449, 1257, 1003, and 936cm1 were assigned to the CH2 deformation, the amide III, the symmetric ring breathing bands of phenylalanine of the protein, and C-C stretching, respectively.6, 13, 17, 18, 19, 20, 21, 22, 23, 24 Aromatic amino acid residues, phenylalanine, tyrosine, and tryptophan were expected to have weak and sharp Raman bands at 620, 642, 852, 1175, 1209, 1605, and 1618cm1 . The bands at 720, 785, 830, 1086, 1340, 1421, and 1577cm1 were assigned to nucleic acids (DNA and RNA).6, 13, 17, 18, 19, 20, 21, 22, 23, 24 Strong bands at 747, 1127, and 1583cm1 were assigned to cyt- c .25, 26 Figure 2 compares the spectra of pure cyt- c , as reduced [Fig. 2], and oxidized [Fig. 2]. Bands observed in the cell spectra indicate that cyt- c in cells mostly exist in a reduced state. Raman spectra of EBC-1 were measured with excitations of 785nm [Fig. 2], and 633nm [Fig. 2]. Cyt- c bands were not observed in these spectra, strongly suggesting that bands due to cyt- c are resonant enhanced via Q-bands from 510to550nm , generally observed in the absorption spectrum of cyt- c , with a 532-nm excitation wavelength (Fig. 2).

Fig. 1

Average Raman spectra of five cells: (a) MRC-5 (normal cell line), (b) RERF-LC-MS (AD cell line), (c) EBC-1 (SQ cell line), (d) Lu-65 (LC cell line), and (e) RERF-LC-MA (SC cell line).

017009_1_057001jbo1.jpg

Fig. 2

Raman spectra of EBC-1 measured with excitation wavelengths of (a) 785, (b) 633, and (c) 532nm . Spectra of (d) reduced, and (e) oxidized states of cyt- c .

017009_1_057001jbo2.jpg

All the cell strains had the adherence property; this cell adherence indicated healthy growth on the culture dish.21 Raman measurement was carried out to the confluent cell sheet. After the Raman measurements, cell morphology and cell adherence did not change. The results of the Trypan blue dye staining examination were also negative.20, 22, 27 Hence, we concluded that the cells did not die by using Raman measurement with 532-nm laser irradiation ( 50mW , 60s ). Notingher reported that 488-nm and 514-nm irradiation induced damage to the cell at 5mW of laser power and 5to20min of exposure.20 Spectra of a particular cell were measured repeatedly to estimate the damage threshold for the 532-nm laser exposure at 50mW of laser power. This measurement was carried out for four cells. The bands due to cyt- c showed no change up to 210s (data not shown) in every cell. A band at 1003cm1 , assigned to phenylalanine, showed a reduction in the band intensity for measurements after 210s of exposure time in two cells, but the other two cells did not show spectral changes even after 210s . No spectral changes were observed in the other bands up to 210s of exposure time.

On initial observation, the Raman spectral features of the cells were similar, which suggests that there was no significant difference in the overall molecular composition of the lipids and proteins in the normal and cancer cells. The intensity ratio of the amide I band at 1660cm1 to the CH bending band at 1440cm1 (amide I/CH bend) was high in the spectra measured with excitation wavelengths of 532 and 633nm . This may be due to the water band at 1640cm1 overlapping with the amide I band. The measuring site of the culture dish was altered on every measurement. The laser spot size (which depended on the objective lens) was estimated to be 1μm in diameter at the focus point, which was rather small compared to the 10-μm cell size. It was assumed that each spectrum before averaging reflected the information of localized cytoplasmic components.

To analyze the small differences between normal and cancer cells in mean spectra, we calculated the subtracted spectra of normal and cancer cells. Figure 3 depicts differences between the spectra of four strains in relation to the MRC-5 (normal) strain: (a) RERF-LC-MS, (b) EBC-1, (c) Lu-65, and (d) RERF-LC-MA. The sharp band of phenylalanine at 1003cm1 was used as the standard for subtraction. Strong bands at 748, 1129, and 1586cm1 were assigned to cyt- c , as shown in Figs. 3, 3, 3. Cyt- c , when associated with electron transfer in oxidative phosphorylation, is usually localized at the mitochondrial inner membrane in the cytoplasm. The intensity of these Raman signals seems to be associated with the amount of the mitochondrial contents and the distribution of mitochondria in the cell as reported by Hamada 26 The strong appearance of these bands suggests that the cancer cells were rich in cyt- c relative to phenylalanine, and possibly rich in mitochondrial cyt- c or in mitochondria compared to MRC-5 (normal cells). Figure 4 compares the concentration of cyt- c in MRC-5 and RERF-LC-MS calculated from the band intensity of their Raman spectra. Data are presented as the average and standard deviation (SD) of the mean. Significance was determined using Student’s t -test on the basis of the 40 cells (20 cells each of MRC-5 and RERF-LC-MS), and the p -value was 0.00012. Values of p<0.01 were considered to be statistically significant. The Raman signal intensity of cyt- c of RERF-LC-MA was weak, as shown in Fig. 3, suggesting that there was no significant difference in the concentration of cyt- c between the normal cell and RERF-LC-MA.

Fig. 3

Difference spectra between the normal cell (MRC-5) and the following cancer cells: (a) RERF-LC-MS, (b) EBC-1, (c) Lu-65, and (d) RERF-LC-MA.

017009_1_057001jbo3.jpg

Fig. 4

Mean ratio of the cyt- c band intensity at 1129cm1to1003cm1 (I1129I1003) in RERF-LC-MS and MRC-5. Data are presented as the average ± the SD of the mean.

017009_1_057001jbo4.jpg

A broad negative band near 800cm1 , which appears in Figs. 3, 3, 3, seems to be a baseline fluctuation due to the quartz substrate of the culture dish. Positive bands at 1660, 1450, and 1260cm1 were observed in the difference spectra of RERF-LC-MS [Fig. 3] and RERF-LC-MA [Fig. 3]. Negative bands at 1660 and 1450cm1 were observed in the spectrum of Lu-65 [Fig. 3]. Judging from the band features and wave numbers, these bands were apparently assigned to proteins and/or lipids, and their composition in these cells differed from a normal composition. However, it was difficult to identify the types of proteins and lipids as increasing and decreasing in the cancer cells. Weak positive bands due to nucleic acids were barely observed at 789, 1086, 1265, and 1337cm1 in the subtraction spectra, which suggests that the chromatin ratio (nucleic acid/protein) was higher in the cancer cells than in the normal cell. However, the intensities of these bands were too weak to analyze the chromatin ratio quantitatively.

3.2.

Principal Component Analysis

PCA classification models with a dataset for the normal cell and each cancer cell. The two spectral regions from 760cm1to870cm1 , and from 1000cm1to1110cm1 , seem to have been affected by the Raman band of quartz; therefore, the corresponding variables were excluded from the dataset. Figure 5 shows the score plots of PC1 ( x axis) and PC2 ( y axis) calculated for MRC-5 versus (a) RERF-LC-MS, (b) EBC-1, (c) Lu-65, and (d) RERF-LC-MA. In the score plots, the cancer cells were obviously distinguishable from the normal cell.

Fig. 5

Score plots of PCA models built with PC1 ( x axis) and PC2 ( y axis). PCA models were employed for the following Raman spectral datasets: (a) MRC-5 and RERF-LC-MS, (b) MRC-5 and EBC-1, (c) MRC-5 and Lu-65, and (d) MRC-5 and RERF-LC-MA. The contribution ratio of the total information for each component is shown in each component’s axis label. ○ indicates normal cells; × indicates cancer cells.

017009_1_057001jbo5.jpg

PC1 was the main contributor to the discrimination of cancer cells (RERF-LC-MS, EBC-1, and Lu-65) from the normal cell (MRC-5) in the score plots of Figs. 5, 5, 5. For discrimination of the SC cell (RERF-LC-MA), both PC1 and PC2 were necessary. The contribution ratio of PC1 was typically more than 55% in all models. The loading plots of PC1 in these PCA models are depicted in Figs. 6, 6, 6, 6 . The features in the plot curves closely resemble the difference spectra in Fig. 3. Although PC2 was not necessary to discriminate cancer cells (RERF-LC-MS, EBC-1, and Lu-65) from the normal cell, the typical contribution ratio was near 20%, suggesting that the information extracted from PC2 reflected site dependency in the cytoplasm and in the cell cycle of each cell. In the PC1 loading plots of Fig. 6, the bands were assigned to cyt- c at 748, 1129, and 1586cm1 [Figs. 6, 6, 6]. This indicates that the cancer cells were rich in cyt- c , so these bands contributed to the discrimination of cancer cells (RERF-LC-MS, EBC-1, and Lu-65) from the normal cell (MRC-5).

Fig. 6

Loading weight of PC1 in the following PCA models: (a) MRC-5 and RERF-LC-MS, (b) MRC-5 and EBC-1, (c) MRC-5 and Lu-65, and (d) MRC-5 and RERF-LC-MA. Datasets were excluded from PCA analysis in the spectral ranges of 760870cm1 and 10001110cm1 because the spectra had a strong contribution of window material in the culture dish.

017009_1_057001jbo6.jpg

PC2 was necessary in addition to PC1 for discriminating RERF-LC-MA cancer cells from normal cells in spite of the low contribution rate of 16.4%. This feature of the PC1 loading plot curve is similar to the difference spectrum of Fig. 3. The bands at 1317, 1450, and 1661cm1 appear to have been assigned to proteins. Although lipids have bands close to these wave numbers, we estimated that the contribution of lipids was considerably weak because the C=O stretching band of ester was absent in the curve. bands due to cyt- c appear in the PC2 loading plot instead of the PC1 loading plot, as shown in Fig. 6. These observations suggest that the cancer cell RERF-LC-MA induced a conformational change and/or the production of a special protein.

Using a discrimination plot, it is possible to estimate the malignancy of these cancers in the Raman spectrum. Figure 7 depicts score plots of the PCA classification model of the four cancer cells: RERF-LC-MS (AD), EBC-1 (SQ), Lu-65 (LC), and RERF-LC-MA (SC). The four cancer cell types were almost successfully discriminated in the PCA model. In the score plot of PC1 and PC2, cells were classified into three groups—the SQ group, the LC group, and the group of AD plus SC—except for one SQ cell, which was classified in the LC group. In the score plot of PC1 and PC3, the AD and SC cells were well separated along with the PC3 axis [Fig. 7]. SC is a type of poorly differentiated cancer, and it is generally classified as a high-grade malignancy. LC is also known as a poorly differentiated cancer with high-grade malignancy. AD and SQ cancers are well-differentiated cancers, and their malignancy is relatively low.2 loading plots of PCs are depicted in Fig. 8 . PC1 [Fig. 8] shows bands at 1660cm1 and 1440cm1 , which were assigned to lipids and/or proteins; these differences could be related to cell morphology or to differences in protein and/or lipid production in these cells. PC2 [Fig. 8] shows strong bands originating from cyt- c and several small bands assignable to nucleic acids. PC3 shows strong bands at 1129cm1 and 749cm1 . Judging from their wave numbers, these bonds could be assigned to cyt- c , but PC3 did not produce bands at 1583cm1 and 1314cm1 . We could not determine the reason that this spectral change took place in the cancer cells. It may reflect a change in the ratio of cyt- c to other cytochromes or a change in the protein-cyt- c interaction in the cell.

Fig. 7

Score plots built with (a) PC1 ( x axis) and PC2 ( y axis), and (b) PC1 ( x axis) and PC3 ( y axis) for PCA models calculated from cancer cells RERF-LC-MS (AD), EBC-1 (SQ), RERF-LC-MA (SC), and Lu-65 (LC). The contribution ratios of PC1, PC2, and PC3 are shown in each axis label.

017009_1_057001jbo7.jpg

Fig. 8

Loading plots of (a) PC1, (b) PC2, and (c) PC3 for the PCA classification model in Fig. 7. Datasets were excluded from PCA analysis in the spectral ranges of 760870cm1 and 10001110cm1 because the spectra had a strong contribution of window material in the culture dish.

017009_1_057001jbo8.jpg

3.3.

Linear Discrimination Analysis

LDA was used to discriminate cancer cells from normal cells. To construct the discrimination model, the intensities of cyt- c bands at 747, 1127, and 1583cm1 (Table 2 ) were selected as the independent variables. Five normal cells and 20 cancerous cells, including AD, SQ, LC, and SC, were tested on the prediction model on the basis of the discrimination function. One SQ cell and four SC cells were classified into the normal group (false negative). The accuracy of the discrimination analysis was 80%. The results suggest that Raman bands of cyt- c are a potentially useful marker to distinguish normal lung cells from lung cancer cells except for SC cells. These results are consistent with those for PCA. For optimal classification, we employed the variable selection method of step-wise LDA. Thirteen variable datasets were automatically selected by the step-wise seeking method to maximize the discrimination efficiency of the cell groups. Five cell groups (normal, AD, SQ, LC, and SC) were successfully separated into five groups. The discrimination accuracy, determined by cross-validation, was 100%, which suggests that LDA used with the step-wise selection method is capable of extracting an effective classifier for cell histology. The 13 variables included comprehensive information on cyt- c , proteins, nucleic acids, and lipids. In this analysis, the SC group laid in the region closest to the normal cell group on the discrimination map (due to the fact that SC cells have a spectrum similar to normal cells).

Table 2

Results of LDA for cancer cells and normal cells.

Group of histological typeGroup of predictive model
NormalCancer
Normal50
Adenocarcinoma05
Squamous cell carcinoma114
Large cell carcinoma05
Small cell carcinoma411
Accuracy (normal or cancer)80%

aFalse-negative on the discrimination model

The pathological diagnosis to characterize the histological type of tumor tissue is generally based on morphological information and the nucleus and cytoplasm (N/C) ratio of biopsied tissues, as well as the size and depth of the cancer. 2, 7, 8 However, the present study strongly suggests that the lipids and/or proteins composition and the concentration of cyt- c can be reliable markers to discriminate the histological type and malignancy of the cancer cells. These markers are detectable in live single cells using Raman spectroscopy with the 532-nm excitation wavelength.

4.

Conclusions

In this study, microscopic Raman spectroscopy was applied for direct measurement of human lung cancer cells in a culture dish in situ. High-quality Raman spectra were obtained in 60scell , and no damage was observed in the cells following Raman measurement. The spectra had strong bands assigned to cyt- c due to the resonance effect with an excitation wavelength of 532nm . PCA was successfully applied to analyze spectral differences among the four cancer cells (RERF-LC-MS, EBC-1, Lu-65, and RERF-LC-MA) and one normal cell (MRC-5). The loading plot of the PCA model strongly suggests that resonance Raman bands of cyt- c are useful for distinguishing cancer cells from normal cells. The discrimination accuracy of the four cancer cells using PCA analysis was 80% in the current study. In the discrimination, Raman bands due to cyt- c contributed significantly to PC2 as well as to the bands due to lipids and/or proteins in PC1. These results suggest an advantage to using a 532-nm excitation wavelength, which gives resonance bands due to cyt- c . Step-wise LDA provided an effective model to classify the histological types of these cells by malignancy type and strain. The discrimination accuracy after cross-validation was 100%. The current results demonstrate the potential of 532-nm -excitation Raman spectroscopy for cell diagnosis.

Raman spectra of cancer cells were obtained without interference of fluorescence in the present study. This result suggests that fluorescence emitted from tissue samples is generated in the extracellular matrix. We are currently developing miniaturized Raman probes with several tens μm of spatial resolution and fluorescent instruments for nonstained tissue imaging.28, 29, 30, 31 These techniques will make it possible to focus the Raman measurement on the active cellular part of cancer tissues and exclude interference from the extracellular matrix. The current work provides important information and shows the viability of future application of Raman spectroscopy in cancer diagnosis.

Acknowledgments

This study was supported by Grants-in-Aid for Scientific Research from the Japan Society for the Promotion of Science (JSPS) and from SENTAN (Development of systems and technology for advanced measurement and analysis), Japan Science and Technology Agency (JST).

References

1. 

L. G. Collins, C. Haines, R. Perkel, and R. E. Enck, “Lung cancer: Diagnosis and management,” Am. Fam. Physician, 75 (1), 56 –63 (2007). 0002-838X Google Scholar

2. 

W. D. Travis, E. Brambilla, H. K. Müller-Hermelink, and C. C. Harris, “World Health Organization classification of tumours. Pathology and genetics of tumour of the lung, pleura, thymus and heart,” Tumours of the Lung, 9 –50 IARC Press, Lyon, NY (2004). Google Scholar

3. 

A. Risch and C. Plass, “Lung cancer epigenetics and genetics,” Int. J. Cancer, 123 1 –7 (2008). https://doi.org/10.1002/ijc.23605 0020-7136 Google Scholar

4. 

L.-P. Choo-Smith, H. G. Edwards, H. P. Endtz, J. M. Kros, F. Heule, H. Barr, J. S. Robinson Jr., H. A. Bruining, and G. J. Puppels, “Medical applications of Raman spectroscopy: from proof of principle to clinical implementation,” Biopolymers, 67 1 –9 (2002). https://doi.org/10.1002/bip.10064 0006-3525 Google Scholar

5. 

H. B. Hanlon, R. Manoharan, T.-W. Koo, K. E. Shafer, J. T. Motz, M. Fitzmaurice, J. R. Kramer, I. Itzkan, R. R. Dasari, and M. S. Feld, “Prospects for in vivo Raman spectroscopy,” Phys. Med. Biol., 45 R1 –R59 (2000). https://doi.org/10.1088/0031-9155/45/2/201 0031-9155 Google Scholar

6. 

Z. Movasaghi, S. Rehman, and I. Rehman, “Raman spectroscopy of biological tissues,” Appl. Spectrosc. Rev., 42 (5), 493 –541 (2007). https://doi.org/10.1080/05704920701551530 0570-4928 Google Scholar

7. 

A. S. Haka, K. E. Shafer-Peltier, M. Fitzmaurice, J. Crowe, R. R. Dasari, and M. S. Feld, “Diagnosing breast cancer by using Raman spectroscopy,” Proc. Natl. Acad. Sci. U.S.A., 102 (35), 12371 –12376 (2005). https://doi.org/10.1073/pnas.0501390102 0027-8424 Google Scholar

8. 

A. S. Haka, Z. Volynskaya, J. A. Gardecki, J. Nazemi, J. Lyons, D. Hicks, M. Fitzmaurice, R. R. Dasari, J. P. Crowe, and M. S. Feld, “In vivo margin assessment during partial mastectomy breast surgery using Raman spectroscopy,” Cancer Res., 66 (6), 3317 –3322 (2006). https://doi.org/10.1158/0008-5472.CAN-05-2815 0008-5472 Google Scholar

9. 

Z. Huang, A. McWilliams, H. Lui, D. I. McLean, S. Lam, and H. Zeng, “Near-infrared Raman spectroscopy for optical diagnosis of lung cancer,” Int. J. Cancer, 107 1047 –1052 (2003). https://doi.org/10.1002/ijc.11500 0020-7136 Google Scholar

10. 

Y. K. Min, T. Yamamoto, E. Kohda, T. Ito, and H. Hamaguchi, J. Raman Spectrosc., 36 73 –76 (2005). https://doi.org/10.1002/jrs.1280 0377-0486 Google Scholar

11. 

H. Yamazaki, S. Kaminaka, E. Kohda, M. Mukai, and H. Hamaguchi, “The diagnosis of lung cancer using 1064-nm excited near-infrared multichannel Raman spectroscopy,” Radiat. Med., 21 (1), 1 –6 (2003). 0288-2043 Google Scholar

12. 

N. D. Magee, J. S. Villaumie, E. T. Marple, M. Ennis, J. S. Elborn, and J. J. McGarvey, “Ex vivo diagnosis of lung cancer using a Raman miniprobe,” J. Phys. Chem. B, 113 (23), 8137 –8141 (2009). https://doi.org/10.1021/jp900379w 1089-5647 Google Scholar

13. 

J. W. Chan, D. S. Taylor, S. M. Lane, T. Zwerdling, J. Tuscano, and T. Huser, “Nondestructive identification of individual leukemia cells by laser trapping Raman spectroscopy,” Anal. Chem., 80 (6), 2180 –2187 (2008). https://doi.org/10.1021/ac7022348 0003-2700 Google Scholar

14. 

A. Y. Lau, L. P. Lee, and J. W. Chan, “An integrated optofluidic platform for Raman-activated cell sorting,” Lab Chip, 8 (7), 1116 –1120 (2008). https://doi.org/10.1039/b803598a 1473-0197 Google Scholar

15. 

M. Placzek, B. Przybilla, U. Kerkmann, S. Gaube, and K.-P. Gilbertz, “Effect of ultraviolet (UV) A, UVB or ionizing radiation on the cell cycle of human melanoma cells,” Br. J. Dermatol., 156 (5), 843 –847 (2007). https://doi.org/10.1111/j.1365-2133.2007.07795.x 0007-0963 Google Scholar

16. 

D. I. Pattison and M. J. Davies, In Cancer: Cell Structures, Carcinogens and Genomic Instability, 131 –157 (2006) Google Scholar

17. 

R. J. Swain, G. Jell, and M. M. Stevens, “Non-invasive analysis of cell cycle dynamics in single living cells with Raman micro-spectroscopy,” J. Cell. Biochem., 104 (4), 1427 –1438 (2008). https://doi.org/10.1002/jcb.21720 0730-2312 Google Scholar

18. 

I. Notingher, I. Bisson, A. E. Bishop, W. L. Randle, J. M. P. Polak, and L. L. Hench, “In situ spectral monitoring of mRNA translation in embryonic stem cells during differentiation in vitro,” Anal. Chem., 76 (11), 3185 –3193 (2004). https://doi.org/10.1021/ac0498720 0003-2700 Google Scholar

19. 

G. J. Puppels, J. H. F. Olminkhof, G. M. J. Segers-Nolten, C. Otto, F. F. M. de Mul, and J. Greve, “Laser irradiation and Raman spectroscopy of single living cells and chromosomes: Sample degradation occurs with 514.5nm but not with 660nm laser light,” Exp. Cell Res., 195 (2), 361 –367 (1991). https://doi.org/10.1016/0014-4827(91)90385-8 0014-4827 Google Scholar

20. 

I. Notingher, S. Verrier, H. Romanska, A. E. Bishop, J. M. Polak, and L. L. Hench, “In situ characterization of living cells by Raman spectroscopy,” Spectroscopy, 16 (2), 43 –51 (2002). 0887-6703 Google Scholar

21. 

F. Draux, P. Jeannesson, A. Beljebbar, A. Tfayli, N. Fourre, M. Manfait, J. Sulé-Suso, and G. D. Sockalingum, “Raman spectral imaging of single living cancer cells: A preliminary study,” Analyst (Cambridge, U.K.), 134 (3), 542 –548 (2009). 0003-2654 Google Scholar

22. 

I. Notingher, S. Verrier, S. Haque, J. M. Polak, and L. L. Hench, “Spectroscopic study of human lung epithelial cells (A549) in culture: Living cells versus dead cells,” Biopolymers, 72 (4), 230 –240 (2003). https://doi.org/10.1002/bip.10378 0006-3525 Google Scholar

23. 

A. D. Meade, F. M. Lyng, P. Knief, and H. J. Byrne, “Growth substrate induced functional changes elucidated by FTIR and Raman spectroscopy in in-vitro cultured human keratinocytes,” Anal. Bioanal. Chem., 387 (5), 1717 –1728 (2007). https://doi.org/10.1007/s00216-006-0876-5 1618-2642 Google Scholar

24. 

R. J. Swain and M. M. Stevens, “Raman microscopy for non-invasive biochemical analysis of single cells,” Biochem. Soc. Trans., 35 (3), 544 –549 (2007). https://doi.org/10.1042/BST0350544 0300-5127 Google Scholar

25. 

T. G. Spiro and T. C. Strekas, “Resonance Raman spectra of hemoglobin and cytochrome c: Inverse polarization and vibronic scattering,” Proc. Natl. Acad. Sci. U.S.A., 69 (9), 2622 –2626 (1972). https://doi.org/10.1073/pnas.69.9.2622 0027-8424 Google Scholar

26. 

K. Hamada, K. Fujita, H. I. Smith, M. Kobayashi, Y. Inouye, and S. Kawata, “Raman microscopy for dynamic molecular imaging of living cells,” J. Biomed. Opt., 13 (4), 044027 (2008). https://doi.org/10.1117/1.2952192 1083-3668 Google Scholar

27. 

D. C. Allison and P. Ridolpho, “Use of a trypan blue assay to measure the deoxyribonucleic acid content and radioactive labeling of viable cells,” J. Histochem. Cytochem., 28 (7), 700 –703 (1980). 0022-1554 Google Scholar

28. 

T. Katagiri, Y. S. Yamamoto, Y. Ozaki, Y. Matsuura, and H. Sato, “High axial resolution Raman probe made of a single hollow optical fiber,” Appl. Spectrosc., 63 (1), 103 –107 (2009). https://doi.org/10.1366/000370209787169650 0003-7028 Google Scholar

29. 

Y. S. Yamamoto, Y. Oshima, H. Shinzawa, T. Katagiri, Y. Matsuura, Y. Ozaki, and H. Sato, “Subsurface sensing of biomedical tissues using a miniaturized Raman probe: study of thin-layered model samples,” Anal. Chim. Acta, 619 (1), 8 –13 (2008). https://doi.org/10.1016/j.aca.2008.02.027 0003-2670 Google Scholar

30. 

Y. Oshima, C. Furihata, and H. Sato, “Light sheet direct Raman imaging technique for observation of mixing of solvent,” Appl. Spectrosc., 63 (10), 1115 –1120 (2009). https://doi.org/10.1366/000370209789553264 0003-7028 Google Scholar

31. 

Y. Komachi, H. Sato, K. Aizawa, and H. Tashiro, “Micro-optical fiber probe for use in an intravascular Raman endoscope,” Appl. Opt., 44 (22), 4722 –4732 (2005). https://doi.org/10.1364/AO.44.004722 0003-6935 Google Scholar
©(2010) Society of Photo-Optical Instrumentation Engineers (SPIE)
Yusuke Oshima, Hideyuki Shinzawa, Tatsuji Takenaka, Chie Furihata, and Hidetoshi Sato "Discrimination analysis of human lung cancer cells associated with histological type and malignancy using Raman spectroscopy," Journal of Biomedical Optics 15(1), 017009 (1 January 2010). https://doi.org/10.1117/1.3316296
Published: 1 January 2010
Lens.org Logo
CITATIONS
Cited by 112 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Raman spectroscopy

Cancer

Lung cancer

Principal component analysis

Proteins

Tissues

Tumor growth modeling

Back to Top