Open Access
1 March 2007 Early spectral changes of cellular malignant transformation using Fourier transform infrared microspectroscopy
E. Bogomolny, Mahmoud Huleihel, Y. Suproun, R. K. Sahu, Shaul Mordechai
Author Affiliations +
Abstract
Fourier transform infrared microspectroscopy (FTIR-MSP) is potentially a powerful analytical method for identifying the spectral properties of biological activity in cells. The goal of the present research is the implementation of FTIR-MSP to study early spectral changes accompanying malignant transformation of cells. As a model system, cells in culture are infected by the murine sarcoma virus (MuSV), which induces malignant transformation. The spectral measurements are taken at various postinfection time intervals. To follow up systematically the progress of the spectral changes at early stages of cell transformation, it is essential first to determine and validate consistent and significant spectral parameters (biomarkers), which can evidently discriminate between normal and cancerous cells. Early stages of cell transformation are classified by an array of spectral biomarkers utilizing cluster analysis and discriminant classification function techniques. The classifications indicate that the first spectral changes are detectable much earlier than the first morphological signs of cell transformation. Our results point out that the first spectral signs of malignant transformation are observed on the first and third day of postinfection (PI) (for NIH/3T3 and MEF cell cultures, respectively), while the first visible morphological alterations are observed only on the third and seventh day, respectively. These results strongly support the potential of developing FTIR microspectroscopy as a simple, reagent-free method for early detection of malignancy.

1.

Introduction

Fourier transform infrared (FTIR) microspectroscopy (MSP) has emerged as a powerful tool for chemical analysis because of its ability to provide detailed information on the spatial distribution of chemical composition at the molecular level.1 In applications requiring qualitative and quantitative analysis, the potential of IR spectroscopy to identify chemical components via fingerprinting analysis of their vibrational spectrum is unsurpassed. When this capability is coupled to an IR microscope, microspectroscopy of μm size samples and high contrast microscopy of 2-D samples based on chemical mapping become possible. Its applications cover a range of disciplines including material science, forensics, biochemistry, biomedical science, and geochemistry, comprising both basic and applied research goals.2, 3, 4, 5, 6

Apart from the conventional methods of cancer diagnosis, there is a need to develop new approaches that are safe, noninvasive, and effectively detect malignancy at earliest stages. Early detection of cancer is a guarantee in most cases of an effective treatment and in some cases for a complete cure. FTIR-MSP has shown encouraging trends in the field of cancer diagnosis in the last decade.7 The differences in the absorbance spectra in the mid-IR region between normal and abnormal tissues have been shown to be a possible criterion for detection and characterization of various types of cancers such as: colon,8, 9, 10 breast,11 leukemia,12, 13 cervical, 14, 15, 16, 17, 18, 19, 20 colorectal,21 skin,22, 23 brain,24 prostate,25, 26 and also neck and head tumors.27 Cell cultures are advantageous and more convenient for basic research,28, 29, 30, 31, 32, 33 compared to “real” tissues due to their homogeneity and the ability to control important culture parameters such as growth and malignant transformation rate. Moreover, malignant metamorphose can be monitored by light microscope, in parallel to the spectral measurements. Thus, cell cultures provide an ideal model for detecting early cellular changes during cell transformation.

In the present work, we used two different cell cultures: murine fibroblast cell line (NIH/3T3) and mouse embryonic fibroblast (MEF, primary cells) as a model system to study early spectral changes induced by cancerous transformation. For this purpose, we first validated consistent spectral biomarkers that were found in previous studies as good biomarkers for detection of malignancy34, 35, 36, 37 using completely transformed fibroblast cell lines. These biomarkers were then utilized for the follow up of malignant cell transformation progression as a function of postinfection time.

2.

Materials and Methods

2.1.

Cells and Viruses

Murine fibroblast cell lines (NIH/3T3, long-term in vitro) and mouse embryonic fibroblast cells (MEF, primary cells) were grown at 37°C in Roswell Park Memorial Institute (RPMI) medium supplemented with 10% newborn calf serum (NBCS) and the antibiotics penicillin, streptomycin, and neomycin.

Clone 124 of TB cells chronically releasing the Moloney murine sarcoma virus (MuSV-124) was used to prepare a virus stock that contained an approximately 30-fold excess of MuSV particles over Moloney murine leukemia virus (MuLV) particles.38 MuLV and MuSV used in this research were grown on NIH/3T3 cells. The virus concentration was determined by counting the number of foci (ffu-focus-forming units).

2.2.

Cell Infection and Determination of Malignant Transformation

Monolayers of NIH/3T3 and MEF cells were grown in 9-cm2 tissue culture plates and treated with 0.8μgml of polybrene (a cationic polymer required for neutralizing the negative charge of the cell membrane) for 24h before infection with the virus. Free polybrene was then removed, and both types of cells were incubated at 37°C for 2h with the infecting virus (MuSV-124) at various concentrations in RPMI medium containing 2% of NBCS. The unabsorbed virus particles were removed, fresh medium containing 2% NBCS was added, and the monolayers were incubated at 37°C . After various time intervals, the cell cultures were carefully examined for the appearance of malignant transformed cells by the following methods in parallel:

  • 1. morphological observations

  • 2. growth on soft agar37

  • 3. FTIR-MSP measurements.

2.3.

Sample Preparation for Fourier Transform Infrared Microscopy Measurements

Since ordinary glass slides exhibit strong absorption in the wavelength range of interest, zinc sellenide crystals, which are highly transparent to IR radiation, were used. Cell cultures were washed with a physiological saline solution and picked up from the tissue culture plates after treatment with trypsin (0.25%) for 1min . The cells were pelleted by centrifugation at 1000rpm for 5min . Each pellet was washed twice with saline and resuspended in 100μl of saline. The number of cells was counted with a hematocytometer, and all tested samples were pelleted again and resuspended in an appropriate volume of saline to give a concentration of 1000cellsμl . A drop of 1μl of each sample was placed on a certain area on the zinc sellenide crystal, air dried for 1h , and measured by FTIR microscopy. The radius of such 1-μl drop was about 1mm , producing a monolayer of cells with about 10μm . Figure 1 displays characteristic sites for measurements as observed by a light microscope for normal [Fig. 1a] and completely transformed [Fig. 1b] murine fibroblast cell line (NIH/3T3).

Fig. 1

Photomicrograph of (a) normal NIH/3T3 cell line and (b) completely transformed fibroblast cell line (NIH/MuSV).

024003_1_021702jbo1.jpg

2.4.

Fourier Transform Infrared Microspectroscopy and Data Acquisition

Measurements on cell cultures were performed using the FTIR microscope IR scope 2 with a liquid-nitrogen-cooled mercury-cadmium-telluride (MCT) detector, coupled to the FTIR spectrometer (Bruker Equinox model 55/S, OPUS software). To achieve high signal-to-noise ratio (SNR), 128 co-added scans were collected in each measurement in the wavenumber region 800to4000cm1 . The measurement site was circular with a diameter of 100μm and spectral resolution of 4cm1 was used. To reduce cell amount variation and guarantee proper comparison between different samples, the following procedures were adopted.

  • 1. Each sample was measured at least five times at different spots.

  • 2. Analog to digital converter rates were empirically chosen between 2000to3000countssec (which allows us to measure areas with similar cellular density).

  • 3. The obtained spectra were baseline corrected using the rubberband method with 64 consecutive points and normalized using min-max normalization in OPUS software.39

2.5.

Statistical Analysis

The obtained parameters (biomarkers) were classified using the cluster analysis according to Ward’s method40 and the discriminant classification function (DCF) method.41, 42 The differences were considered significant at P<0.05 .

3.

Results

3.1.

Spectral Differences between Normal and Malignant Cell Lines

The main objective of this research is to identify and study early changes during malignant transformation using FTIR-MSP. As a first step, it was important to find spectral biomarkers that can discriminate between normal and completely malignant cells.

We analyzed FTIR spectra of 50 different samples of both normal (NIH/3T3) and completely transformed murine fibroblast cell lines (NIH/3T3/MuSV). Two regions with significant and consistent differences were identified at 3000to2820cm1 and 1145to1000cm1 . For an effective comparison, these regions were cut from the entire spectra, baseline corrected and normalized.

The results in the region 3000to2820cm1 , presented in Fig. 2a , show four prominent absorbance bands: near 2852cm1 (due to the symmetric stretching of the methylene chains in membrane lipids); at 2923cm1 (due to the antisymmetric CH2 stretch); at 2958cm1 (due to antisymmetric stretching of the methyl groups of both lipids and proteins); and at 2871cm1 (arising from the symmetric CH3 stretching mode).1 The average absorption intensities of normal and transformed fibroblast cell lines are distinctive at 2852-cm1 and 2958-cm1 bands [Fig. 2a]. It was found that the best discriminating values were obtained by deriving the intensity ratio of these two vibrational modes (i.e., A2958A2852 or vas CH3vsCH2 ).

Fig. 2

FTIR spectra in the regions: (a) 2820to3000cm1 , (b) 1000to1145cm1 , and (c) the second derivative at 1000to1145cm1 of the normal murine fibroblast cell line (NIH/3T3) and of the completly transformed murine fibroblast cell line (NIH3T3/MuSV). Spectra are the average of 50 samples and five measurements of each sample after baseline correction and normalization.

024003_1_021702jbo2.jpg

The dimensionless ratio eliminates artifact, which may arise due to the baseline contribution underneath each band. Table 1 summarizes the statistical values of the previous ratio for the normal and malignant cell line. The t -value of the two groups is 11.25 (Table 1). Therefore, this ratio may be considered as a satisfactory biomarker to follow the progress of malignant transformation.

Table 1

Statistical analysis of the biomarkers derived from FTIR spectra of normal and transformed murine fibroblast cell lines.

A2958∕A2852 A1121∕A1015 Wavenumber shift due to νs PO2− (relative to 1082cm−1 )
NormalAbnormalNormalAbnormalNormalAbnormal
Mean1.091.271.694.261.753.3
SD0.050.060.370.610.210.75
T -value11.2516.28.87
P -value 11010 1.21013 1.9108
Max value1.181.372.45.532.134.5
Min value1.021.191.213.251.432.12

In the second region at 1145to1000cm1 [Fig. 2b], there are plenty of overlapping vibrational modes associated with absorbance of macromolecules such as proteins, nucleic acids, carbohydrates, and phospholipids. The bands at 1082 and 1056cm1 correspond to absorbance of the νs PO2 of phosphodiesters of nucleic acids1 and the O–H stretching coupled with C–O bending of C–OH groups of carbohydrates, respectively.39 Other bands at 1121 and 1015cm1 can be clearly seen in the second derivative spectra [Fig. 2c]. Previous works have shown that A1121 arises from RNA absorbance, whereas the 1015cm1 shoulder is due to DNA.43, 44, 45

From this region it is possible to derive two additional spectral biomarkers with outstanding statistical characteristics: A1121A1015 ratio (assigned as RNA/DNA ratio) and the wavenumber shift due to νs PO2 (relative to 1082cm1 ). Even though the variability of these biomarkers is high due to overlapping absorbance, the average values of normal are still significantly different compared to malignant cells (Table 1).

3.2.

Early Stages of Malignant Cell Transformation

Both primary cells (MEF) and murine fibroblast cell lines (NIH/3T3) were infected with MuSV (1ffucell) and examined at various postinfection times for morphological and spectral changes. Figure 3 shows the expanded spectra of both cell cultures in the two wavenumber regions. This figure clearly demonstrates the gradual spectral variations following cell infection. Dramatic changes are observable in the case of MEF transformation, where the band at 1056cm1 decreases gradually and the band at 1082cm1 is shifted systematically to higher wavenumbers versus infection time [Fig. 3d]. Thus, it is possible to determine the first spectral signs of malignancy according to the alterations in the calculated values of the previously discussed biomarkers. The observed spectral changes in malignant cells compared to control cells are summarized in Table 2 . As can be seen in Table 2, the first morphological changes confirmed by microscopical observations and growth on soft agar appear considerably later than the first spectral signs. For example, the first spectral identification is possible on the first day ( A1121A1015 biomarker), while morphologically it can be discerned on the third day in the case of NIH/3T3 cell transformations (Table 2). In the case of MEF primary cells, the spectral changes induced by cell transformation were even more significant compared to those induced in the NIH/3T3 transformation [Fig. 3b]. Also in MEF primary cells, the first spectral signs appeared significantly earlier than the morphological changes (on the third day compared to the seventh day).

Fig. 3

FTIR spectra of (a) NIH/3T3 and (b) MEF cells at various intervals of postinfection time in the region 2820to3000cm1 . FTIR spectra of (c) NIH/3T3 and (d) MEF cells in various intervals of postinfection time in the region 1000to1145cm1 . Spectra are the average of (a) and (c) 20 transformations of NIH/3T3 cell lines and (b) and (d) 12 transformations of MEF cells.

024003_1_021702jbo3.jpg

4.

Statistical Analysis

4.1.

Cluster Analysis

Cluster analysis was used to classify the infected cells at each postinfection day into cancerous or normal groups. For this classification we utilized a vector array of spectral biomarkers, which were set as follows:

(A2958A2852A1121A1015shiftofνsPO2).
The postinfection days of transformations were characterized using the average values of previously derived spectral biomarkers.

The results presented in Fig. 4 show that cluster analysis can indeed classify the infected cells into the cancerous group already at the first postinfection day in case of NIH/3T3 cells and at the fifth postinfection day in case of MEF transformation. We note that in both cases, the classifications were significantly earlier than the morphological identification of the malignant cells.

Fig. 4

Cluster analysis of (a) NIH/3T3 and (b) MEF cells at various intervals of postinfection. Cluster analysis is based on the average values of three biomarkers. Each postinfection day of transformation is represented using array of an average value of three biomarkers: ( A2958A2852 , A1121A1015 , and the shift of νs PO2 ).

024003_1_021702jbo4.jpg

4.2.

Discriminant Classification Function

Discriminant classification function (DCF) is a statistical tool that enables us to improve discrimination between malignant stages by representing an adequate quantitative follow up of transformations versus time. DCF generates a classification score for each postinfection day, which is a linear combination of previously derived array of biomarkers with weight coefficients,37, 38 as can be seen in the following equation:

S=c+w1x1+w2x2++wmxm+,
where wm is the weight coefficient, xm is biomarker value, and S denotes the resultant classification score.

The weight coefficients were determined empirically in such a way that they nullify the average classification score of NIH/3T3 array (normal cell score) and yield 100 score (cancerous score) for the average NIH/MuSV array. The same weights were applied also to the MEF transformation. From the DCF analysis, we obtained a classification of the transformations that correspond to a sigmoid fit (Fig. 5 ). The abnormality can be distinguished as early as the first day in the case of the NIH/3T3 transformation [while first morphological identification is possible on the third day, Fig. 5a]. Similarly, Fig. 5b shows that the abnormality in the case of MEF transformation is apparent on the third postinfection day (while the first morphological identification is possible on the seventh day). In both cases, the infected cells reach an upper level plateau after full transformation (score of 100 for NIH/3T3 and score of 137 for cancerous MEF, Fig. 5).

Fig. 5

Discriminant classification function of (a) NIH/3T3 and (b) MEF cells at various intervals of postinfection. Each postinfection day of transformation is represented using an array of average values of three biomarkers: ( A2958A2852 , A1121A1015 , and the shift of νs PO2 ).

024003_1_021702jbo5.jpg

Table 2

First signs of malignant transformation.

Type ofcells incultureFirst spectral detection of each biomarker (postinfection day andthe percentage changes compared to the control cells)Firstmorphologicalsigns
A2958∕A2852 A1121∕A1015 shift due to νs PO2−
NIH/3T3day 2/ 8.25±2.7% day 1 / 53.8±11% day 2/ 43.2±19% Day 3
MEFday 5/ 32.3±8% day 3/ 99.4±23% day 5/ 54.3±21% Day 7

5.

Discussion

In the present work, we implemented FTIR-MSP to study the spectral changes of cancerous transformation in vitro and focused mainly on early detection of malignancy. For this purpose we utilized an array of spectral biomarkers ( A2958A2852 , A1121A1015 , and the wavenumber shift due to νs PO2 ). The obtained results revealed that three FTIR spectral indicators consistently altered during the malignant transformation and can discern malignancy before morphological changes can be observed. Such spectral alterations were considerably higher and significant in the transformations of primary MEF cells compared to NIH/3T3 cell line transformations. Most of the biological characteristics of primary cells are completely different from those of cell lines; they replicate slowly in culture and are very sensitive to the environmental conditions compared to the cell lines. In fact, primary cells are very similar to the normal organism cells in most of their characteristics, while cell lines have some similarity to malignant cells. The gradual changes in these biomarkers during the transformation processes can arise from several cellular activities.

5.1.

A2958A2852 ( νas CH3νsCH2 )

The phospholipids/lipids/triglycerides and proteins absorb in the wavenumber regions from 2800cm1to3000cm1 .1, 43 Previous studies46, 47 with rat fibroblast cell lines showed that lipids have more predominant absorbance relative to other biomolecules, including proteins, at the 2800- to 3000- cm1 region. Also, changes in the absorbance due to νs CH2 and νas CH3 vibrational modes of lipids during carcinogenesis were found.48, 49 Our results showed remarkable increment in this ratio in malignant cells compared to normal cell cultures. Similar behavior of this biomarker was observed in leukemia,50 cervical, colon, and colorectal cancer,51 as well as in murine fibroblast cell lines and rabbit bone marrow primary cells transformed by MuSV or H-Ras.50 In addition, the CH3CH2 ratio was found to increase as a function of the progress in malignant lymphoma grade.52

Lipids are considered as important components of the cellular membrane, which significantly affect its permeability and metabolites transportation during carcinogenesis,53 and they also form an influential source of energy that might be essential for malignant metabolism. Moreover, the evidence that transformed cells differ in their average cellular volume compared to the normal cells54 (Fig. 1) may also contribute to the observed changes seen in the previous ratio.

5.2.

A1121A1015 (RNA/DNA) and Wavenumber Shift due to νs PO2

The region between 900to1200cm1 has many overlapping bands that correspond to the nucleic acid absorbance.43, 44, 45 Differences in DNA isolated from cancer and normal cells/tissues using FTIR spectroscopy have been the basis of a number of studies for diagnosis of cancer.55, 56 A statistical comparison of the FTIR spectra of DNA obtained from prostate cancer and from normal prostate tissues of healthy younger men revealed a broad array of differences in base structures (e.g., N–H and C–O) as well as in vertical base-stacking interactions and in the phosphodiester-deoxyribose backbone.55 Also, structural disorders in the pancreatic tumor DNA were detected in the phosphodiester-deoxyribose spectral region.56

Our results showed significant increment in the A1121A1015 ratio and νs PO2 peak shift to higher wavenumbers in malignant cells compared to normal cell cultures. The utilization of these spectral indicators was widely reported in previous studies. 7, 8, 9, 10, 12, 13, 19, 20, 34, 35, 36, 37, 57 The same tendency was observed for A1121A1015 ratio in melanoma,58 leukemia,12, 13 cervical,19, 20 and colon,7, 8, 9, 10, 57 cancers as well as murine fibroblast cell lines transformed by MuSV or H-Ras34, 35, 36 and lymphoma.51, 59 In the case of νs PO2 , it was found that the phosphodiester group shifted to higher wavenumbers in various malignant cell cultures such as human primary fibroblast, mouse primary fibroblast, murine fibroblast cell line (NIH/3T3), etc.37 This shift has been also seen in breast cancer,60 cancerous stomach tissues,61 and neoplastic human gastric cells.62

It was suggested that the pivotal role of these biomarkers stems from nucleic acid absorbance. The change in the absorbance and conformation of nucleic acids (which can cause a shift in the absorbance of the phosphodiesters group) during carcinogenesis arises from a sharp increment of the proliferation and metabolic activity in the transformed cells and from the high levels of retrovirus DNA and RNA production in the infected or transformed cells. Also, these changes could arise from a variation in the nuclear volume of the transformed cells as previously reported.63, 64 Cell transformation progression in time can be well described by a sigmoid function that was obtained using a discriminant classification function. In a short interval of postinfection time (meaningfully shorter than the cell cycle), there are no detectable spectral differences between infected and control cell cultures. Then the spectral values of the infected cell cultures gradually approach the spectral values of the fully transformed cells (as can be seen in Fig. 5). After obtaining first signs of morphological transformation (as confirmed by microscope and growth on soft agar), all spectral indicators showed identical values to those of the fully transformed cells.

The results presented in this study prove the superiority of FTIR spectroscopy over the conventional technique used for detection of malignant cells in culture. Thus, FTIR-MSP in tandem with proper statistical tools may offer a promising technique for the detection of early stages of malignancy and for monitoring their progression.

Acknowledgments

This research work was supported by the Israel Science Foundation (ISF grant number 788/01) and the Israel Cancer Association (ICA).

References

1. 

H. Mantsch and D. Chapman, Infrared Spectroscopy of Biomolecules, John Wiley, New York (1996). Google Scholar

2. 

M. Diem, S. Boyston-White, and L. Chiriboga, “Infrared spectroscopy of cells and tissues: Shining light on to a novel subject,” Appl. Spectrosc., 53 148A –161A (1999). https://doi.org/10.1366/0003702991946712 0003-7028 Google Scholar

3. 

D. Naumann, D. Helm, and H. Labischinski, “Microbiological characterizations by FT-IR spectroscopy,” Nature (London), 351 81 –82. (1991). https://doi.org/10.1038/351081a0 0028-0836 Google Scholar

4. 

H. Gremlich and B. Yang, Infrared and Raman Spectroscopy of Biological Materials, Marcel Dekker, New York (2001). Google Scholar

5. 

P. Lasch, M. Boese, A. Pacifico, and M. Diem, “FT-IR spectroscopic investigations of single cells on the subcellular level,” Vib. Spectrosc., 28 (1), 147 –157 (2002). 0924-2031 Google Scholar

6. 

S. Argov, R. K. Sahu, E. Bernshtain, A. Salman, G. Shohat1, U. Zelig, and S. Mordechai, “Inflammatory bowel diseases as an intermediate stage between normal and cancer: a FTIR-microspectroscopy approach,” Biopolymers, 75 (5), 384 –392 (2004). 0006-3525 Google Scholar

7. 

R. K. Sahu and S. Mordechai, “Fourier transform infrared spectroscopy in cancer detection,” Future Oncol., 1 (5), 635 –647 (2005). Google Scholar

8. 

S. Argov, J. Ramesh, A. Salman, I. Silenikov, J. Goldstein, H. Guterman, and S. Mordechai, “Diagnostic potential of fourier transformed infrared microspectroscopy and advanced computational methods in colon cancer patients,” J. Biomed. Opt., 7 (2), 248 –254 (2002). https://doi.org/10.1117/1.1463051 1083-3668 Google Scholar

9. 

A. Salman, S. Argov, J. Ramesh, J. Goldstein, I. Silenikov, H. Guterman, and S. Mordechai, “FTIR microscopic characterization of normal and malignant human colonic tissues,” Cell Mol. Biol. (Paris), 47 (22), 159 –166 (2001). 0145-5680 Google Scholar

10. 

R. K. Sahu, S. Argov, E. Bernshtain, A. Salman, S. Walfisch, J. Goldstein, and S. Mordechai, “Detection of abnormal proliferation in histologically ‘normal’ colonic biopsies using FTIR-microspectroscopy,” Scand. J. Gastroenterol., 39 (6), 557 –566 (2004). 0036-5521 Google Scholar

11. 

T. Gao, J. Feng, and Y. Ci, “Human breast carcinomal tissues display distinctive FTIR spectra: implication for the histological characterization of carcinomas,” Anal. Cell. Pathol., 18 87 –93 (1999). 0921-8912 Google Scholar

12. 

R. K. Sahu, U. Zelig, M. Huleihel, N. Brosh, M. Talyshinsky, M. Harosh, S. Mordechai, and J. Kapelushnik, “Continuous monitoring of WBC (biochemistry) in an adult leukemia patient using advanced FTIR-spectroscopy,” Leuk. Res., 30 (6), 687 –693 (2006). https://doi.org/10.1016/j.leukres.2005.10.011 0145-2126 Google Scholar

13. 

J. Ramesh, J. Kapelushnik, J. Mordechai, A. Moser, M. Huleihel, V. Erukhimovitch, C. Levi, and S. Mordechai, “Novel methodology for the follow-up of acute lymphoblastic leukemia using FTIR Microspectroscopy,” J. Biochem. Biophys. Methods, 51 (3), 251 –261 (2002). 0165-022X Google Scholar

14. 

L. Chiriboga, P. Xie, H. Yee, D. Zarou, D. Zakim, and M. Diem, “Infrared spectroscopy of human tissue. IV. Detection of human cervical tissue via infrared microscopy,” Cell Mol. Biol. (Paris), 44 (1), 219 –229 (1998). 0145-5680 Google Scholar

15. 

P. Podshyvalov, R. K. Sahu, S. Mark, K. Kantarovich, H. Guterman, J. Goldstein, J. Ramesh, S. Argov, and S. Mordechai, “Distinction of cervical cancer biopsies by use of infrared microspectroscopy and probabilistic neural networks,” Appl. Opt., 44 (18), 3725 –3734 (2005). https://doi.org/10.1364/AO.44.003725 0003-6935 Google Scholar

16. 

W. Steller, J. Einenkel, L. C. Horn, U. D. Braumann, H. Binder, R. Salzer, and C. Kraft, “Delimitation of squamous cell cervical carcinoma using infrared microspectroscopic imaging,” Anal. Bioanal. Chem., 384 (1), 145 –154 (2006). 1618-2642 Google Scholar

17. 

B. R. Wood, L. Chiriboga, D. McNaughton, and M. Diem, “Fourier transform infrared (FTIR) spectral mapping of the cervical transformation zone, and dysplastic squamous epithelium,” Gynecol. Oncol., 93 (1), 59 –68 (2004). 0090-8258 Google Scholar

18. 

J. I. Chang, Y. B. Huang, P. C. Wu, C. C. Chen, S. C. Huang, and Y. H. Tsai, “Characterization of human cervical precancerous tissue through the fourier transform infrared microscopy with mapping method,” Gynecol. Oncol., 91 (3), 577 –583 (2003). 0090-8258 Google Scholar

19. 

S. Mordechai, R. K. Sahu, Z. Hammody, S. Mark, K. Kantarovich, H. Guterman, P. Podshyvalov, J. Goldstein, and S. Argov, “Possible common biomarkers from FTIR microspectroscopy of cervical cancer and melanoma,” J. Microsc., 215 86 –91 (2004). 0022-2720 Google Scholar

20. 

S. Mark, R. K. Sahu, K. Kantarovich, P. Podshyvalov, H. Guterman, J. Goldstein, S. Argov, and S. Mordechai, “Fourier transform infrared microspectroscopy as a quantitative diagnostic tool for assignment of premalignancy grading in cervical neoplasia,” J. Biomed. Opt., 9 (3), 558 –567 (2004). https://doi.org/10.1117/1.1699041 1083-3668 Google Scholar

21. 

P. Lasch, W. Haensch, D. Naumann, and Max Diem, “Imaging of colorectal adenocarcinoma using FT-IR microspectroscopy and cluster analysis,” Biochim. Biophys. Acta, 1668 (2), 176 –186 (2004). 0006-3002 Google Scholar

22. 

A. Tfayli, O. Piot, A. Durlach, A. Bernard, and M. Manfait, “Discriminating nevus and melanoma on paraffin-embedded skin biopsies using FTIR microspectroscopy,” Biochim. Biophys. Acta, 1724 (3), 262 –269 (2005). 0006-3002 Google Scholar

23. 

N. Eikje, K. Aizawa, and Y. Ozaki, “Vibrational spectroscopy for molecular characterization and diagnosis of benign, premalignant and malignant skin tumours,” Biotechnol. Annu. Rev., 11 191 –225 (2005). Google Scholar

24. 

C. Krafft, L. Shapoval, S. B. Sobottka, K. D. Geiger, G. Schackert, and R. Salzer, “Identification of primary tumors of brain metastases by SIMCA classification of IR spectroscopic images,” Biochim. Biophys. Acta, 1758 (7), 883 –891 (2006). 0006-3002 Google Scholar

25. 

E. Gazi, M. Baker, J. Dwyer, N. P. Lockyer, P. Gardner, J. H. Shanks, R. S. Reeve, C. A. Hart, N. W. Clarke, and M. D. Brown, “A correlation of FTIR spectra derived from prostate cancer biopsies with gleason grade and tumour stage,” Eur. Urol., 50 (4), 750 –760 (2006). 0302-2838 Google Scholar

26. 

D. Fernandez, R. Bhargava, S. M. Hewitt, and I. W. Levin, “Infrared spectroscopic imaging for histopathologic recognition,” Nat. Biotechnol., 23 (4), 469 –474 (2005). 1087-0156 Google Scholar

27. 

P. Bruni, C. Conti, E. Giorgini, M. Pisani, C. Rubini, and G. Tosi, “Histological and microscopy FT-IR imaging study on the proliferative activity and angiogenesis in head and neck tumours,” Faraday Discuss., 126 19 –26 (2004). 0301-7249 Google Scholar

28. 

M. German, A. Hammiche, N. Ragavan, M. J. Tobin, J. L. Cooper, S. S. Matanhelia, A. C. Hindley, C. M. Nicholson, N. J. Fullwood, H. M. Pollock, and F. L. Martin, “Infrared spectroscopy with multivariate analysis potentially facilitates the segregation of different types of prostate cell,” Biophys. J., 90 (10), 3783 –3795 (2006). 0006-3495 Google Scholar

29. 

Y. Yang, J. Sule-Suso, G. D. Sockalingum, G. Kegelaer, M. Manfait, and A. El Haj, “Study of tumor cell invasion by Fourier transform infrared Microspectroscopy,” Biopolymers, 78 (6), 311 –317 (2005). 0006-3525 Google Scholar

30. 

L. Zhang, G. W. Small, A. S. Haka, L. H. Kidder, and E. N. Lewis, “Classification of Fourier transform infrared microscopic imaging data of human breast cells by cluster analysis and artificial neural networks,” Appl. Spectrosc., 57 (1), 14 –22 (2003). https://doi.org/10.1366/000370203321165151 0003-7028 Google Scholar

31. 

J. Sule-Suso, A. Forster, V. Zholobenko, N. Stone, and A. El Haj, “Effects of CaCl2 and MgCl2 on Fourier transform infrared spectra of lung cancer cells,” Appl. Spectrosc., 58 (1), 61 –67 (2004). https://doi.org/10.1366/000370204322729487 0003-7028 Google Scholar

32. 

C. M. Krishna, G. Kegelaer, I. Adt, S. Rubin, V. B. Kartha, M. Manfait, and G. D. Sockalingum, “Combined Fourier transform infrared and Raman spectroscopic approach for identification of multidrug resistance phenotype in cancer cell lines,” Biopolymers, 82 (5), 462 –470 (2006). 0006-3525 Google Scholar

33. 

C. M. Krishna, G. D. Sockalingum, R. A. Bhat, L. Venteo, P. Kushtagi, M. Pluot, and M. Manfait, “FTIR and Raman microspectroscopy of normal, benign, and malignant formalin-fixed ovarian tissues,” Anal. Bioanal. Chem., 387 (5), 1649 –1656 (2007). 1618-2642 Google Scholar

34. 

M. Huleihel, A. Salman, V. Erukhimovitch, J. Ramesh, Z. Hammody, and S. Mordechai, “Novel spectral method for the study of viral carcinogenesis in vitro,” J. Biochem. Biophys. Methods, 50 (2–3), 111 –121 (2002). https://doi.org/10.1016/S0165-022X(01)00177-4 0165-022X Google Scholar

35. 

J. Ramesh, A. Salman, Z. Hammody, B. Cohen, J. Gopas, N. Grossman, and S. Mordechai, “Application of FTIR microscopy for the characterization of malignancy: H-ras transfected murine fibroblasts as an example,” J. Biochem. Biophys. Methods, 50 (1), 33 –42 (2001). 0165-022X Google Scholar

36. 

A. Salman, J. Ramesh, V. Erukhimovitch, M. Talyshinsky, S. Mordechai, and M. Huleihel, “FTIR microspectroscopy of malignant fibroblasts transformed by mouse sarcoma virus,” J. Biochem. Biophys. Methods, 55 (2), 141 –153 (2003). 0165-022X Google Scholar

37. 

V. Erukhimovitch, M. Talyshinsky, Y. Souprun, and M. Huleihel, “Spectroscopic characterization of human and mouse primary cells, cell lines and malignant cells,” Photochem. Photobiol., 76 446 –451 (2002). https://doi.org/10.1562/0031-8655(2002)076<0446:SCOHAM>2.0.CO;2 0031-8655 Google Scholar

38. 

J. Ball, J. McCarter, and S. Sunderland, “Evidence for helper independent murine sarcoma virus. I. Segregation of replication defective and transformation defective viruses,” Virology, 56 268 –284 (1973). 0042-6822 Google Scholar

39. 

S. Wartewig, IR and Raman Spectroscopy, 75 –122 John Wiley and Sons, New York (2003). Google Scholar

40. 

B. Everitt, Cluster Analysis, John Wiley and Sons, New York (1980). Google Scholar

41. 

C. Huberty, Applied Discriminant Analysis, John Wiley and Sons, New York (1994). Google Scholar

42. 

R. A. Fisher, “The use of multiple measures in taxonomic problems,” Ann. Eugenics, 7 179 –188 (1936). Google Scholar

43. 

F. S. Parker, Application of Infrared Spectroscopy in Biochemistry, Biology and Medicine, Plenum, New York (1971). Google Scholar

44. 

D. Naumann, “FT-infrared and FT-Raman spectroscopy in bio-medical research,” Appl. Spectrosc. Rev., 36 239 –298 (2001). 0570-4928 Google Scholar

45. 

E. Benedetti, E. Bramanti, F. Papineschi, I. Rossi, and E. Benedetti, “Determination of the relative amount of nucleic acids and proteins in leukemic and normal lymphocytes by means of Fourier transform infrared microspectroscopy,” Appl. Spectrosc., 51 792 –797 (1997). https://doi.org/10.1366/0003702971941304 0003-7028 Google Scholar

46. 

J. Mourant, K. W. Short, S. Carpenter, N. Kunapareddy, L. Coburn, T. M. Powers, and J. P. Freyer, “Biochemical differences in tumorigenic and nontumorigenic cells measured by Raman and infrared spectroscopy,” J. Biomed. Opt., 10 (3), 031106 (2005). https://doi.org/10.1117/1.1928050 1083-3668 Google Scholar

47. 

K. W. Short, S. Carpenter, J. P. Freyer, and J. Mourant, “Raman spectroscopy detects biochemical changes due to proliferation in mammalian cell cultures,” Biophys. J., 88 4274 –4288 (2005). https://doi.org/10.1529/biophysj.103.038604 0006-3495 Google Scholar

48. 

L. M. McIntosh, M. Jackson, H. H. Mantsch, M. F. Stranc, D. Pilavdzic, and A. N. Crowson, “Infrared spectra of basal cell carcinomas are distinct from non-tumor-bearing skin components,” J. Invest. Dermatol., 112 951 –956 (1999). https://doi.org/10.1046/j.1523-1747.1999.00612.x 0022-202X Google Scholar

49. 

T. Gao and Y. Ci, “Fourier transform infrared spectroscopic features of human breast benign and malignant tissues,” Internet J. Vibrational Spectrosc., 3 (5), (1999). Google Scholar

50. 

R. K. Sahu, S. Argov, A. Salman, U. Zelig, M. Huleihel, N. Grossman, J. Gopas, J. Kapelushnik, and S. Mordechai, “Can vibrational spectroscopy at higher wavenumbers (midIR) shed light on universal biomarkers for carcinogenesis in tissues?,” J. Biomed. Opt., 10 (5), 054017 (2005). https://doi.org/10.1117/1.2080368 1083-3668 Google Scholar

51. 

B. Rigas, S. Morgello, I. S. Goldman, and P. T. Wong, “Human colorectal cancers display abnormal Fourier-transform infrared spectra,” Proc. Natl. Acad. Sci. U.S.A., 87 8140 –8144 (1990). 0027-8424 Google Scholar

52. 

P. G. Andrus, “Cancer monitoring by FTIR spectroscopy,” Technol. Cancer Res. Treat., 5 (2), 157 –167 (2006). 1533-0346 Google Scholar

53. 

C. Dombi, J. Banoczy, M. Kramer, W. P. Wertz, and A. C. Squier, “Study of the permeability of oral leukoplakia,” Fogorv. Sz., 92 (5), 137 –142 (1999). 0015-5314 Google Scholar

54. 

F. Lang, M. Ritter, E. Woll, H. Weiss, D. Haussinger, J. Hoflacher, K. Maly, and H. Grunicke, “Altered cell volume regulation in ras oncogene expressing NIH fibroblasts,” Pfluegers Arch., 420 (5–6), 424 –427 (1992). 0031-6768 Google Scholar

55. 

D. C. Malins, P. M. Johnson, E. A. Barker, N. L. Polissar, T. M. Wheeler, and K. M. Anderson, “Cancer-related changes in prostate DNA as men age and early identification of metastasis in primary prostate tumors,” Proc. Natl. Acad. Sci. U.S.A., 100 (9), 5401 –5406 (2003). https://doi.org/10.1073/pnas.0931396100 0027-8424 Google Scholar

56. 

V. R. Kondepati, M. Keese, H. M. Heise, and J. Backhaus, “Detection of structural disorders in pancreatic tumour DNA with Fourier-transform infrared spectroscopy,” Vib. Spectrosc., 40 33 –39 (2006). 0924-2031 Google Scholar

57. 

R. K. Sahu, S. Argov, A. Salman, M. Huleihel, N. Grossman, Z. Hammody, J. Kapelushnik, and S. Mordechai, “Characteristic absorbance of nucleic acids in the mid-IR region as possible common biomarkers for diagnosis of malignancy,” Technol. Cancer Res. Treat., 3 (6), 629 –638 (2004). 1533-0346 Google Scholar

58. 

Z. Hammody, R. K. Sahu, S. Mordechai, E. Cagnano, and S. Argov, “Characterization of malignant melanoma using vibrational spectroscopy,” Scientific World J., 5 173 –182 (2005). Google Scholar

59. 

P. G. Andrus and R. D. Strickland, “Cancer grading by Fourier transform infrared spectroscopy,” Biospectroscopy, 4 (1), 37 –46 (1998). https://doi.org/10.1002/(SICI)1520-6343(1998)4:1<37::AID-BSPY4>3.0.CO;2-P 1075-4261 Google Scholar

60. 

C. Liu, Y. Zhang, X. Yan, X. Zhang, C. Li, W. Yang, and D. Shi, “Infrared absorption of human breast tissues in vitro,” J. Lumin., 119–120 132 –136 (2006). 0022-2313 Google Scholar

61. 

Y. Tong and Y. Lin, “FTIR study on the normal and cancerous stomach tissues,” Guangpuxue Yu Guangpu Fenxi, 21 (3), 324 –327 (2001). 1000-0593 Google Scholar

62. 

N. Fujioka, Y. Morimoto, T. Arai, K. Takeuchi, M. Yoshioka, and M. Kikuchi, “Differences between infrared spectra of normal and neoplastic human gastric cells,” Spectroscopy (Amsterdam), 18 59 –66 (2004). 0712-4813 Google Scholar

63. 

D. Mihailovic, B. Dordevic, and V. Mihailovic, “Nuclear volume in type I gastric intestinal metaplasia,” Anal Quant Cytol. Histol., 21 (2), 143 –144 (1999). 0884-6812 Google Scholar

64. 

R. Drezek, M. Guillaud, T. Collier, I. Boiko, A. Malpica, C. Macaulay, M. Follen, and R. Richards-Kortum, “Light scattering from cervical cells throughout neoplastic progression: influence of nuclear morphology, DNA content, and chromatin texture,” J. Biomed. Opt., 8 (1), 7 –16 (2003). https://doi.org/10.1117/1.1528950 1083-3668 Google Scholar
©(2007) Society of Photo-Optical Instrumentation Engineers (SPIE)
E. Bogomolny, Mahmoud Huleihel, Y. Suproun, R. K. Sahu, and Shaul Mordechai "Early spectral changes of cellular malignant transformation using Fourier transform infrared microspectroscopy," Journal of Biomedical Optics 12(2), 024003 (1 March 2007). https://doi.org/10.1117/1.2717186
Published: 1 March 2007
Lens.org Logo
CITATIONS
Cited by 52 scholarly publications and 6 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
FT-IR spectroscopy

Absorbance

Imaging spectroscopy

Statistical analysis

Cancer

Tissues

Biological research

Back to Top