Open Access
6 February 2012 Identification of fungal phytopathogens using Fourier transform infrared-attenuated total reflection spectroscopy and advanced statistical methods
Ahmad Salman, Elad Shufan, Itshak Lapidot, Ami Pomerantz, Mahmoud Huleihel, Leah Tsror, Raymond Moreh, Shaul Mordechai
Author Affiliations +
Abstract
The early diagnosis of phytopathogens is of a great importance; it could save large economical losses due to crops damaged by fungal diseases, and prevent unnecessary soil fumigation or the use of fungicides and bactericides and thus prevent considerable environmental pollution. In this study, 18 isolates of three different fungi genera were investigated; six isolates of Colletotrichum coccodes, six isolates of Verticillium dahliae and six isolates of Fusarium oxysporum. Our main goal was to differentiate these fungi samples on the level of isolates, based on their infrared absorption spectra obtained using the Fourier transform infrared-attenuated total reflection (FTIR-ATR) sampling technique. Advanced statistical and mathematical methods: principal component analysis (PCA), linear discriminant analysis (LDA), and k-means were applied to the spectra after manipulation. Our results showed significant spectral differences between the various fungi genera examined. The use of k-means enabled classification between the genera with a 94.5% accuracy, whereas the use of PCA [3 principal components (PCs)] and LDA has achieved a 99.7% success rate. However, on the level of isolates, the best differentiation results were obtained using PCA (9 PCs) and LDA for the lower wavenumber region (800-1775  cm−1), with identification success rates of 87%, 85.5%, and 94.5% for Colletotrichum, Fusarium, and Verticillium strains, respectively.

1.

Introduction

The fungi investigated in this study belong to the same phylum (divisions) named “Ascomycota,” a group whose members are also known as Sac fungi, and have evolved from one common ancestor. Living organisms are biologically classified into different units according to their similarity. A species is often defined as a group of organisms capable of interbreeding and producing fertile offspring. More precise measures are often used, such as similarity of DNA, morphology, or ecological niche. Species that are believed to have the same ancestors are grouped together, and this group is called a genus. Each species may include different isolates (strains) which usually result from one or more mutations. In the hierarchy of the binomial classification system, species is above isolate and below genus.

As far as the genus is concerned, these fungi are grouped into three classes: Fusarium, Colletotrichum, and Verticillium. These fungi are pathogens which attack crops, resulting in fungal diseases, which lead to large economic losses.1,2 For example, V. ahlia causes a wilt disease in hundreds of species of eudicot plants. Many economically important plants are susceptible, including cotton, tomatoes, potatoes, eggplants, and peppers. Solanaceous crop may be infected at any age by fungi causing the Fusarium wilt and Verticillium wilt, with similar symptoms. The diseases therefore cannot be distinguished based on symptoms alone.

Figures 1(a)1(c)3,4 show plants infected by the various fungi genera studied here. Figure 1(a) shows potato tubers infected with F. oxysporum which causes a wilt symptom characterized by browning of the vascular ring and the stem end. In Fig. 1(b) Verticillium wilt symptoms on tomato leaves can be seen. Infected plants usually survive, but both the yields and the fruits may become small, depending on the severity of attack. Figure 1(c) shows a tomato infected with C. coccodes which usually causes serious damage to the fruit.

Fig. 1

Different fungi genera affecting crops. (a) Potato tuber discoloration of the vascular ring, caused by the Fusarium wilt, (b) v-shaped lesions on tomato leaves, associated with Verticillium wilt, and (c) C. coccodes causes anthracnose on tomatoes and black dot disease on potatoes.

JBO_17_1_017002_f001.png

Early detection of phytopathogens is very important for both successful protection and effective treatment,5 thus enhancing the chances of recovery and simultaneously preventing environmental pollution. By standard physiological methods, it is very difficult to differentiate between closely related species and strains6,7 based on cultivation. In addition, fungi identification by visual and microscopic observations are usually time consuming and not always very specific.8

Serological, molecular biology methods are based on differences in the nucleic acid genome of the tested fungi, whereas the immunological method is based on differences in the proteins constructing the fungi and on the availability of specific monoclonal antibodies against the tested fungi. These methods are very sensitive and rapid relative to physiological methods for the identification of pathogens; however, their use is still limited due to their availability for only a small number of fungal pathogens.911 Developing such methods for different fungi strains is complicated, costly, and not always possible. Therefore, the use of these methods is expected to be limited and not available for the screening of large numbers of samples of different fungal strains.

Many studies have indicated the potential of FTIR spectroscopy methods for the detection and identification of microorganisms, especially in food products.12,13 The vast information already achieved about spectral bands obtained from FTIR spectra of living cells14 combined with features of infrared spectroscopy such as sensitivity, rapidity, low expense, and simplicity15 enabled us to differentiate the fungi even on the levels of species16,17 and strains.1823 The use of Fourier transform infrared-attenuated total reflection (FTIR-ATR) spectroscopy followed by principal component analysis (PCA) and linear discriminant analysis (LDA) for biological classification has gained momentum in the last decades (see Martin et al.).2426

Linker and Tsror19 showed good results in differentiating between fungi genera and species using FTIR-ATR and applying statistical analysis techniques: PCA, and cluster and canonical variate analysis (CVA). In a different study, A. Naumann18 classified 26 fungi strains belonging to 24 different species using FTIR-ATR and applying cluster and artificial neural network (ANN) statistical techniques. Fischer et al.22 tried to develop a method to reproducibly differentiate Aspergillus and Penicillium species on the generic, species, and strain levels. In their work they used nine different species and only one strain from each species. Shapaval et al.23 showed that it was possible to differentiate 11 species of five different fungal genera using FTIR spectroscopy. Our former study21 took the method further into a more challenging stage; by applying the PCA and LDA techniques, the ability of the FTIR-ATR method to classify six different strains of F. Oxysporum was examined. The results showed that it is possible to classify and differentiate between the strains with a 81.4% success rate.

The main goal of this study was to test the feasibility of the FTIR-ATR methodology in differentiating 18 isolates from three different fungi genera: Colletotrichum, Verticillium, and Fusarium. From each of these genera, we selected six isolates originating in a single species, namely six isolates of C. coccodes, six isolates of V. dahliae and six isolates of F. oxysporum. The classification procedure was done in two phases. In the first phase, the fungal samples were classified on the genus level, and in the second phase, the samples were classified on the isolates level. The classification procedures were based on infrared absorption spectra of the samples, and used advanced mathematical and statistical techniques; PCA followed by k-means and LDA.

The uniqueness of our study lies in the analysis of a larger number of isolates belonging to the same species. Isolates are very similar; their spectra are blended and overlap, which makes it difficult to differentiate between them, especially relative to the species and genera levels as previously mentioned. Moreover, for pattern recognition methods, increasing the class numbers is a real challenge, especially when the classes are strains of the same species as in this study.

2.

Materials and Methods

2.1.

Fungi

Eighteen isolates of three different fungi genera, namely C. coccodes, V. dahliae, and F. oxysporum, were examined, six isolates of each. The samples were obtained from the Department of Plant Pathology at the Gilat Experiment Station, ARO, Israel.

Fungi samples were isolated from different infected crops obtained directly from the plant stem, root, or tuber. Samples were scratched from the infected areas of the crops, mixed well in 1 ml of a potato dextrose medium and cultivated for several days on potato dextrose agar (PDA, Difco Laboratories, Becton, Dickison and Company, Sparks, MD) at 27°C. Several cultures (5 to 10) of single fungi colonies, obtained by micromanipulation, of each strain were cultivated in different batches. These cultures were grown for 3 to 10 days at 27°C in continuous shaking. Samples of the growing fungi were identified by visual and microscopic observations.

As a first step, the cultivated fungi strains were identified using classical microbiological techniques.27,28 Next, samples of the fungi were separated and purified by centrifuging about 1.5 mL of the mixture at 13,200 rpm for 4 min, rinsing the extract 4 times with distilled water, and suspending the pellet in an appropriate volume of distilled water (1mL) for spectroscopic measurements.

Pure samples enable us to control as many experimental parameters (such as growth conditions, amount of sample examined, duration of growth, etc.) as possible, and to verify that each absorption band in the IR spectrum was due to the specific sample.

2.2.

Sample Preparation

Due to the complicated structure of the fungi, it is difficult to achieve a homogeneous suspension of the fungi in water or spread them on the ZnSe crystal surface of the attenuated total reflection (ATR) accessory. Therefore, the fungi were mashed into small pieces and mixed as evenly as possible into the distilled water to obtain a suspension. About 500 µL of each fungal suspension sample was spread as homogenously as possible on the surface of the ATR ZnSe crystal to cover the entire crystal surface. These samples were air dried for about 30 min until all water had evaporated, and then measured by ATR spectroscopy. The ATR crystal, of a trapezoid shape 80 mm long, 10 mm wide, and 4 mm thick was obtained from PIKE technologies.

2.3.

FTIR Measurements

We used a Bruker Tensor 27 spectrometer in the ATR mode with DTGS detector for our measurements. After drying them, the samples were scanned 128 times in the range of 600 to 4000cm1, with a 4cm1 spectral resolution. OPUS software was used for spectral manipulation such as baseline correction, bisecting, and normalization. Our measurements were carried out over several weeks.

2.4.

Spectral Analysis

Two-hundred fifty six spectra were measured from six different C. coccodes isolates, 131 spectra were measured from six different F. oxysporum isolates, and 105 spectra were measured from six different V. dahliae isolates. All spectra were corrected with OPUS software according to the wavelength dependence of the penetration depth. The spectra were then bisected into two regions (800775cm1) and (28002990cm1) to exclude the water absorption band and the “dead” region between amide I and the lipid bands. The spectra in each region were baseline corrected using the rubber band method, normalized separately using the vector normalization method, and then offset corrected using OPUS software.

2.5.

Statistical Analysis

2.5.1.

PCA

PCA is an unsupervised29,30 multivariate analysis tool for dimensionality reduction31,32 which is widely used in pattern recognition. The common assumption is that the most separable dimensions are those with the highest variance; it frequently is so, but not necessarily.

Here, X={x1,,xN} is the data set, where xnRd×1, and Σ is the covariance matrix of X. Be Λ=[λ1λN] the vector of eigenvalues, so that λ1λ2λN with corresponding eigenvectors matrix: U=[u1ud]; uiRd×1. Then, the data after PCA is:

Y=UqTX=[u1uq]TX;qd,
where T is the transform operator.

Basically, PCA is a mathematical operator that projects the high dimensional data onto a subspace of low dimension which captures the orthogonal directions with the highest variability, i.e., instead of using many variables the variability in the data is described using only a few principal components (PCs).33

The first projection coefficient is called the first principal component (PC1). It contains most of the variance. PC2 contains most of the residual variance and is perpendicular to the first. The other PCs obey the same rules. This method allows the reduction of our spectra to three variables in the first phase of classification (genus level) and nine variables in the second phase of classification (isolate level), in the lower wavenumber region that accounts for almost 100% of the variance.34

2.5.2.

LDA

Following PCA, we applied the LDA,35 where the separation is based on the assumption that different classes sharing the same covariance matrix have different mean vectors for classification. The separation procedure is done by maximizing the probability of a class under these assumptions. The formulation of LDA is as follows36: given data which was derived from k classes, with a Gaussian probability density function (PDF) for each class:

Eq. (1)

fk(x)=1(2π)d/2|Σk|1/2exp{12(xμk)TΣk1(xμk)},
where μk is the mean vector and the covariance matrix Σk. Assuming that each class’s prior probability is πk, and all the classes have the same covariance matrix, Σk=Σ. The classification of the new input data xRd×1 is performed as follows:

Eq. (2)

k^=argmaxk{1,,K}{xTΣ1μk12μkTΣ1μk+log(πk)}.
Training and test sets were selected randomly from the database.

The examination of the results performed using two variants of k-fold cross-validation is applied frequently in pattern recognition. The first run was 5-fold, i.e., 20% to 80% when 80% of the data was used for training and 20% for testing. Each time, another 20% (1 of 5) was used for testing and the rest (4 parts) for training. We ran this approach 20 times, each time with a random partition of the data into 5 groups. The other variant “leave-one-out,”31,32 is when k=N, where N is the number of data points. It is usually applied when the amount of data is relatively small.

2.5.3.

K-Means

K-means is widely used for clustering and vector quantization.37 The aim is to partition the space into K cells, each represented by a vector, named a centroid or code-word. All code-words compose the codebook of the quantization process {ck}k=1K. The goal is to minimize some predefined average distance:

Eq. (3)

{ck*}k=1K=min{ck}k=1K{1Nn=1ND(cln,xn)}ln=argminl1,,KD(cl,xn)
The most common distance is the square Euclidian distance, D(x,y)=(xy)T(xy). The partitioning was performed using an iterative algorithm which converges to a local minimum.

3.

Results

Figure 2 shows the infrared absorption spectra of the three fungi genera investigated in this study, i.e., Colletotrichum, Verticillium, and Fusarium, each with a different color. The spectra for each fungus were measured from six different isolates. The main spectral features in the high wavenumber region (27002950cm1) are the bands detected at 2853 and 2922cm1. These bands derive mainly from phospholipid absorbance.38 Water absorbance bands in this region were excluded from the spectra as part of the analysis procedure. The main features in the low wavenumber region (8001775cm1), were the amide I and amide II with centroids at 1650cm1 and 1553cm1, respectively. There is a clear peak which arises from lipids absorbance39 at 1743cm1. A large absorbance band at 1076cm1 was mainly attributed to carbohydrate and nucleic acid vibrations. The centroid of the amid III band was detected at 1252cm1. The chitin band, which is specific to fungi, was detected at 1151cm1 and 1078cm1 due to its C−O and C−C stretching vibrations, and the glycogen C−O stretching vibration was detected at 1024cm1.

Fig. 2

Infrared absorption spectra of C. coccodes, V. dahliae, and F. oxysporum in the ranges 890 to 1770cm1 and 2800 to 2990cm1.

JBO_17_1_017002_f002.png

As previously mentioned, there are clear differences between the fungi mainly in the lower wavenumber region, namely, in the amide I and II, chitin, and glycogen peaks. These differences enabled us to differentiate the three genera using unsupervised methods, such as k-means, achieving an accuracy of 94.5%.

As shown in Fig. 2, the Colletotrichum fungus has low intensities in the amide I and II regions which may indicate lower protein content relative to Fusarim and Verticillium fungi.

Utilizing the PCA and LDA (20 to 80) methods, we classified the sample into three groups with great success. When using one PC we obtained an accuracy of 61.2%, using two PCs a 96.9% accuracy was achieved, which increased to 99.7% with three PCs.

Figure 3 shows the results for the three fungal groups obtained by LDA based on three PCs of PCA at the lower wavenumber region.

Fig. 3

Complete date sets of Colletotrichum, Verticillium, and Fusarium presented in a 3-D PCs domain. This figure summarizes the first phase (genus) classification results.

JBO_17_1_017002_f003.png

In this figure each spectrum was calculated as a superposition of three loadings (PCs) derived using the PCA algorithm. Each spectrum was identified by three numbers, which are the coefficients of the loadings. As shown in Fig. 3 the data is divided into three groups, red squares (Colletotrichum), blue crosses (Fusarium), and green open triangles (Verticillium). There is an excellent separation between the three groups. To find out which feature contributed to the classification procedure, we plotted two-dimensional figures for PC1 versus PC2, depicted in Fig. 4(a), which provided the best classification between the three studied genera. The projection of the data on the PC1 axis provided a good classification of Fusarium. The projection of the Colletotrichum and Verticillium data on the PC2 axis provided a complete classification between them. The first two PCs are plotted in Fig. 4(b) where the main peaks are labeled. These are: for PC1, the C=O stretching mode of lipids (1743cm1), the amide I and II bands (1641cm1 and 1544cm1, respectively), the asymmetric and symmetric CH3 bending modes of the protein methyl groups40 (1453cm1 and 1397cm1), collagen, the asymmetric stretching of phosphate groups of phosphodiester linkages in DNA and RNA41 (1224cm1), carbohydrates42 (1148cm1), the symmetric phosphate PO2 stretching43 (1076cm1), and OCH3 [polysaccharides-cellulose44 (989cm1)]. All the peaks labeled with dashed lines are also common for PC2. The other main peaks of PC2 are: C−N stretching and C−N−H bending of N−H45 (1478cm1), collagen46 (1345cm1), amide III, collagen47 (1280cm1), phosphate, oligosaccharides,48 and carbohydrates42 (1146cm1), stretching C−O ribose49 (1068cm1), C−C [Ref. 49(994cm1)], and the phosphodiester stretching bands region [for absorbance due to collagen and glycogen50 (926cm1)].

Fig. 4

(a) Two-dimensional plot of Colletotrichum, Verticillium, and Fusarium included in this study. (b) This figure presents the first 2 PCs, derived from PCA, in the lower range. The main peaks of each PC are labeled.

JBO_17_1_017002_f004.png

In the second phase we tried to classify the six different isolates within each group (genus). In contrast to the genus level where there were clear differences between the spectra, here the spectra of the different strains were blended and overlap [Figs. 5(a)5(c)], and in each spectrum all major absorption bands described earlier were detected. However, some differences were found, such as different absorbance intensities in different absorbance bands in the two regions.

Fig. 5

Infrared absorption spectra of six isolates of C. coccodes (a), six isolates of V. dahliae (b), and six isolates of F. oxysporum (c) in the ranges 800 to 1770cm1 and 2800 to 2990cm1.

JBO_17_1_017002_f005.png

The analysis was performed for different regions of the spectrum, and the best results were achieved when the lower wavenumber region (8001775cm1) was used. The k-means achieved poor results, thus we adopted the PCA and LDA methods. In the second phase where the number of measurements for each strain was low, the “leave-one-out” method was preferred, whereas in the first phase where the statistics were good the 20 to 80 algorithm was used.

Figure 6 shows the identification errors (percentage) derived using LDA together with the “leave-one-out” and 20 to 80 algorithms, as a function of the PC number used in the analysis of Colletotrichum. Due to the similarity between the strains of each of the three groups, more PCs had to be used than in the first (genus) phase.21 Nine PCs were used in the strain phase, which achieved a good classification and simultaneously kept the highest loading (PC) meaningful and noiseless.

Fig. 6

Identification error vs. PC No. in the lower wavenumber region.

JBO_17_1_017002_f006.png

The LDA results using the “leave-one-out” algorithm and nine PCs showed that it was possible to differentiate between the Colletotrichum isolates with a 86.8% success rate, whereas the classification success rates of the Fusarium and Verticillium strains were 85.5% and 94.6%, respectively. The LDA results using the 20 to 80 algorithm and 9 PCs showed that it was possible to differentiate between the Colletotrichum isolates with a 84.6% success rate, whereas the classification success rates of the Fusarium and Verticillium strains were 80.7% and 94.3%, respectively.

Table 1 shows the results of the classification of the studied genera on the isolate level using the “leave-one-out” method.

Table 1

Successful identification of (a) Colletotrichum (Coll), (b) Fusarium (Fus) and (c) Verticillium (Vert) isolates obtained using LDA calculations and the “leave-one-out” algorithm in the low wavenumber region (900–1775  cm−1).

Coll 1Coll 2Coll 3Coll 4Coll 4Coll 5
(a)Coll 11704000
Coll 20151020
Coll 30118100
Coll 40052001
Coll 50000191
Coll 60010023
Fus 1Fus 2Fus 3Fus 4Fus 4Fus 5
(b)Fus1800110
Fus 2071110
Fus 30010010
Fus 4020910
Fus 50001130
Fus 60000012
Vert 1Vert 2Vert 3Vert 4Vert 4Vert 5
(c)Vert 11310000
Vert 21190000
Vert 30020000
Vert 41001000
Vert 5030090
Vert 60000028

4.

Discussion

In principle, the ATR sampling technique is similar to measurements done using remote fiberoptic probes. Thus, evaluation of the ATR potential for fungal classification on large numbers of strains on the levels of genus, species, and isolates is an important step toward in vivo measurements using infrared fibers.

Many studies have shown that it is possible to use FTIR-ATR for detection and identification of fungi.18,19,21 In their study, Linker et al.19 managed to classify different fungi genera species and to differentiate between two isolates of Colletotrichum based on the measured FTIR-ATR spectra and advanced statistical methods: CVA and PCA. Based on the ATR sampling technique, Naumann18 succeeded to differentiate between 26 fungal isolates that belong to 24 different species using cluster analysis and ANN analysis.18

In their work, Fischer et al.22 used nine different species and only one strain from each species. Shapaval et al.23 used 11 different species of five different fungal genera. In our previous work, we used the FTIR-ATR to differentiate between six different isolates of F. oxysporum.21

By enlarging the number of strains, we come closer to reality, where tens of isolates from each species exist. All previous studies focused on just a few species belonging to the same genus, and a few strains belonging to the same species. Thus, success in the classification of large numbers of isolates and species provides a solid base for the future large scale in vivo examination of phytofungal pathogens using infrared spectroscopy.

In this study we investigated six different isolates from each of three fungi genera Colletotrichum, Verticillium and Fusarium. Our main objective was to test and evaluate the potential of FTIR-ATR spectroscopy to differentiate between these isolates. This was done in two phases of examinations of fungi samples, the first on the genus level and the second on the isolate level. The fungal isolates to be tested were collected, isolated, and purified to obtain pure samples of the desired strain.

Our analysis was based on the fungal IR evanescent wave absorption spectra. We used a statistical approach to analyze the large database of the obtained IR spectra.

Using the currently practiced classic physiological methods, it is very difficult to accurately differentiate between the strains because of their high morphological similarity. Molecular and serological methods are also not available for all these strains.

Due to the variance in the penetration depth occurring in the ATR mode measurements, absorption band intensities and positions are different than those obtained using the transmission mode.5153 Thus, all our spectra have undergone the same penetration depth correction using OPUS software. On the genus level, there were clear differences that have enabled the differentiation between the samples by unsupervised clustering analysis using K-means, with a 94.5% success rate. In-house developed software codes based on MATLAB software54 were used. After PCA calculation, a supervised classifier LDA was applied, which enabled the classification of the data with an accuracy of 99.7%. In this phase of the work (genus level), 3 PCs were sufficient to obtain such high accuracy, whereas in the second phase (isolate level) nine PCs were required. PCA enabled capturing the variability of the fungi using a small number of PCs. In the first phase the examined fungi genera were successfully differentiated with a high accuracy using two PCs, and using three PCs obtained almost a 100% success rate. The second phase, in which the samples were isolates of the same species, the differences between their spectra were minor, and therefore a good differentiation was achieved using a larger number of PCs. It is not guaranteed that the first PCs are the most important for separability; however, they carry most of the variance of the data.

The dependence on the PC number is task-dependent, i.e., number of classes to classify, level of classification (genus, species, or isolates), and also the type of samples.

From Fig. 4(a) we can see that the new directions defined by PC1 and PC2 give a good classification. Using PCA calculation, we obtained a new basis which defines new directions.55 Figure 4(b) shows PC1 and PC2 vectors in the standard basis. The peaks which are found in PC2 and not in PC1 (labeled by dashed lines) are suspected to be more dominant with respect to the classification. The comparison procedure of the peaks is relatively easy when a small number (2 to 3) of the PCs have to be compared.

Smaller numbers of PCs make it easier to relate the changes of the PCs shape to biology, whereas this is much harder when a larger number of PCs is used. In this study, the goal was classification, and thus we should choose the number of PCs which provided the best classification providing that the highest PC was still meaningful for the biological system investigated. This choice of PC number is still not objective in our field of spectroscopy and should be investigated more carefully.

As shown in Fig. 4(a), a good classification of Fusarium was achieved. This is probably due to some peaks in PC1 which are absent in PC2 [Fig. 4(b)]. These peaks which contribute mainly to the classification of Fusarium are 1453cm1 (methyl groups of proteins),40 1224cm1 (collagen, asymmetric, stretching of phosphate groups of phosphodiester linkages in DNA and RNA),41 1148cm1 (chitin, carbohydrates)42 and 1076cm1 (chitin, symmetric phosphate PO2 stretching).43

PC2 gives a good classification between Colletotrichum and Verticillium. It can be seen in Fig. 4(b) that some peaks in PC2 do not exist in PC1, and this is the main contribution to the classification between Colletotrichum and Verticillium. These peaks are 1478cm1 (C−N stretching and C−N−H bending N−H),45 1345cm1 (collagen),46 1280cm1 (amide III, collagen),47 1146cm1 (phosphate, oligosaccharides,48 and carbohydrates)42 and 1068cm1 (C−O ribose),49 994cm1 (C−O ribose, C−C),49 and 926cm1 [phosphodiester stretching bands region (for absorbance due to collagen and glycogen)].50

The LDA calculation was designed applying the two methods (“leave-one-out” and 20 to 80), using different sets for prediction each time, so that the test sets would be statistically independent. The “leave-one-out” method is a common method of cross-validation, extensively explored in machine learning, used to estimate the error in a small sized populations. Using it ensures the validation of results.56,57

The 20 to 80 method is used when large numbers of data exist, thus it was used in the first phase of the study. In the second phase of identification the “leave-one-out” was preferred because it achieved better results, but the 20 to 80 was used as well to validate the “leave-one-out” algorithm (Fig. 6).

Our results with a successful classification of the different strains of each genus using the “leave-one-out” method, are presented in Table 1(a)–(c).

In the second phase, nine PCs were used, thus the discrimination power is the overall contribution of the whole set of PCs. It is difficult to outline which PC contributes more to the discrimination between the different isolates (strains).

It is very difficult to discuss the relevance of spectral differences of the various species or isolates, due to the high similarity in the molecular and structural composition of these related isolates, which belong to the same species. According to the FTIR-ATR spectra, the differences are spread over the entire spectrum and not specific to a certain band. One limitation of FTIR is that the spectra obtained from biological samples are complex, and difficult to interpret because it is hard to know which bands arise from which biomolecule. For example, changes in the membrane are reflected in the absorption bands of lipids; however, spectral differences in the lipid absorption bands may not arise solely from membranes, but may be attributed to lipids that are not located in the cell membrane.

The applied method is objective and computerized. Enlarging the database and the strain number will improve the statistics and bring the method closer to reality where large numbers of strains exist. The method shows a great potential for the identification and study of phytofungal pathogens, as far as the level of strain identification.

This method may be useful for studying biological aspects of the different genera as shown in Fig. 2. It shows significantly lower intensities in the amide I and II regions for Colletotrichum compared with Fusarim and Verticillium fungi. These differences may reflect different protein contents between these genera. Certainly, small differences in their structural components (such as lipids, proteins, and sugars) do exist, but unfortunately the origins of these differences are yet unknown and it is very difficult to define the exact biological origin of the observed spectral differences. However, these difficulties do not affect the main objective of this study, which was to examine the potential of spectroscopic techniques for reliable detection and discrimination between different fungi strains.

5.

Conclusions

FTIR-ATR spectroscopy in tandem with PCA methods, followed by LDA calculations, enabled the differentiation between the fungal samples studied not only on the genus level but on the level of strains as well (where spectral differences are minute) with a good confidence level.

Enlarging the database to include more species and strains, could improve the statistics and bring the method one step forward toward real conditions, where tens of isolates from the same species exist.

Acknowledgment

Financial support by SCE internal research funding is gratefully acknowledged.

References

1. 

G. N. Agrios, Plant Pathology, Academic Press, New York (1978). Google Scholar

2. 

J. Katan, “Principles in plant pathology,” Plant Diseases in Israel, Volcani Center, Bet-Dagan, Israel (1998). Google Scholar

5. 

G. V. Doernet al., “Clinical impact of rapid in vitro susceptibility testing and bacterial identification,” J. Clin. Microbiol., 32 (7), 1757 –1762 (1994). JCMIDW 1070-633X Google Scholar

6. 

G. H. Kimet al., “Ophiostomatoid fungi isolated from Pinus radiata logs imported from New Zealand to Korea,” Can. J. Bot., 83 (3), 272 –278 (2005). http://dx.doi.org/10.1139/b04-170 CJBOAW 1480-3305 Google Scholar

7. 

U. MorethO. Schmidt, “Investigations on ribosomal DNA of indoor wood decay fungi for their characterization and identification,” Holzforschung, 59 (1), 90 –93 (2005). http://dx.doi.org/10.1515/HF.2005.014 HOLZAZ 1437-434X Google Scholar

8. 

G. L. SchumannC. J. DArcy, Essential Plant Pathology, The American Phytopathological Society, St. Paul, MI (2006). Google Scholar

9. 

S. NikkariD. A. Relman, “ Molecular approaches for identification of infectious agents in Wegener's granulomatosis and other vasculitides,” Curr. Opin. Rheumatol., 11 (1), 11 –16 (1999). http://dx.doi.org/10.1097/00002281-199901000-00003 CORHES 1040-8711 Google Scholar

10. 

N. C. Clarket al., “Detection of a streptomycin/spectinomycin adenylyltransferase gene (aadA) in Enterococcus faecalis,” in Antimicrob. Agents. Chemother., 157 –160 (1999). Google Scholar

11. 

M. VaneechoutteJ. Eldere Van, “The possibilities and limitation of nucleic acid amplification technology in diagnostic microbiology,” J. Med. Microbiol., 46 (3), 188 –194 (1997). http://dx.doi.org/10.1099/00222615-46-3-188 JMMIAV 0022-2615 Google Scholar

12. 

M. J. Guptaet al., “Differentiation of food pathogens using FTIR and artificial neural networks,” Trans. ASAE., 48 (5), 1889 –1892 (2005). TARSBB 2151-0032 Google Scholar

13. 

H. Lamprellet al., “ Discrimination of staphylococcus aureus strains from different species of staphylococcus using Fourier transform infrared (FTIR) spectroscopy,” Int. J. Food Microbiol., 108 (1), 125 –129 (2006). http://dx.doi.org/10.1016/j.ijfoodmicro.2005.10.015 IJFMDD 0168-1605 Google Scholar

14. 

D. NaumannD. HelmH. Labischinski, “Microbiological characterizations by FT-IR spectroscopy,” Nature, 351 (6321), 81 –82 (1991). http://dx.doi.org/10.1038/351081a0 NATUAS 0028-0836 Google Scholar

15. 

M. DiemS. Boydston-WhiteL. Chiriboga, “ Infrared spectroscopy of cells and tissues: shinning light onto a novel subject,” Appl. Spectrosc., 53 (4), 148 –161 (1999). http://dx.doi.org/10.1366/0003702991946712 APSPA4 0003-7028 Google Scholar

16. 

S. H. Beattieet al., “Discrimination among Bacillus cereus, B. mycoides and B. thuringiensis and some other species of the genus Bacillus by Fourier transform infrared spectroscopy,” FEMS Microbiol. Lett., 164 (1), 201 –206 (1998). http://dx.doi.org/10.1111/fml.1998.164.issue-1 FMLED7 0378-1097 Google Scholar

17. 

D. Lefieret al., “Effect of sampling procedure and strain variation in Listeria monocgtogenes on the discrimination of species in the genus Listeria by Fourier transform infrared spectroscopy and canonical variates analysis,” FEMS Microbiol. Lett., 147 (1), 45 –50 (1997). http://dx.doi.org/10.1111/j.1574-6968.1997.tb10218.x FMLED7 0378-1097 Google Scholar

18. 

A. Naumann, “A novel procedure for strain classification of fungal mycelium by cluster and artificial neural network analysis of Fourier transform infrared (FTIR) spectra,” Analyst, 134 (6), 1215 –1223 (2009). http://dx.doi.org/10.1039/b821286d ANLYAG 0365-4885 Google Scholar

19. 

R. LinkerL. (Lahkim) Tsror, “Discrimination of soil-borne fungi using Fourier transform infrared attenuated total reflection spectroscopy,” Appl. Spectrosc., 62 (3), 302 –305 (2008). http://dx.doi.org/10.1366/000370208783759678 APSPA4 0003-7028 Google Scholar

20. 

T. UdelhovenD. NaumannJ. Schmitt, “Development of a hierarchical classification system with artificial neural networks and FT-IR spectra for the identification of bacteria,” Appl. Spectrosc., 54 (10), 1471 –1479 (2000). http://dx.doi.org/10.1366/0003702001948619 APSPA4 0003-7028 Google Scholar

21. 

A. Salmanet al., “Distinction of Fusarium oxysporum fungal isolates (strains) using FTIR-ATR spectroscopy and advanced statistical methods,” Analyst, 136 (5), 988 –95 (2011). http://dx.doi.org/10.1039/c0an00801j ANLYAG 0365-4885 Google Scholar

22. 

G. Fischeret al., “FT-IR spectroscopy as a tool for rapid identification and intra-species characterization of airborne filamentous fungi,” J. Microbiol. Methods, 64 (1), 63 –77 (2006). http://dx.doi.org/10.1016/j.mimet.2005.04.005 JMIMDQ 0167-7012 Google Scholar

23. 

V. Shapavalet al., “A high-throughput microcultivation protocol for FTIR spectroscopic characterization and identification of fungi,” J. Biophotonics, 3 (8–9), 512 –521 (2010). http://dx.doi.org/10.1002/jbio.201000014 JBOIBX 1864-063X Google Scholar

24. 

F. L. Martinet al., “Distinguishing cell types or populations based on the computational analysis of their infrared spectra,” Nat. Protoc., 5 (11), 1748 –60 (2010). http://dx.doi.org/10.1038/nprot.2010.133 NPARDW 1750-2799 Google Scholar

25. 

J. Trevisanet al., “Syrian hamster embryo (SHE) assay (pH 6.7) coupled with infrared spectroscopy and chemometrics towards toxicological assessment,” Analyst, 135 (12), 3266 –3272 (2010). http://dx.doi.org/10.1039/c0an00586j ANLYAG 0365-4885 Google Scholar

26. 

M. J. Walshet al., “ATR microspectroscopy with multivariate analysis segregates grades of exfoliative cervical cytology,” Biochem. Biophys. Res. Commun., 352 (1), 213 –219 (2007). http://dx.doi.org/10.1016/j.bbrc.2006.11.005 BBRCA9 0006-291X Google Scholar

27. 

L. (Lahkim) Tsroret al., “Aggressiveness of Verticillium dahliae isolates from different vegetative compatibility groups to potato and tomato,” Plant Pathol., 50 (4), 477 –482 (2001). http://dx.doi.org/10.1046/j.1365-3059.2001.00587.x PLPAAD 1365-3059 Google Scholar

28. 

L. (Lahkim) TsrorM. Hazanovsky, “Effect of co-inoculation by Verticillium dahliae and Colletotrichum coccodes on disease symptoms and fungal colonization in four potato cultivars,” Plant Pathol., 50 (4), 483 –488 (2001). https://doi.org/doi://dx.doi.org/10.1046/j.1365-3059.2001.00585.x PLPAAD 1365-3059 Google Scholar

29. 

A. M. C. DaviesT. Fearn, “Back to basics: the principles of principal component analysis,” Spectrosc. Europe, 4 (17), 20 –23 (2004). SPEUEF 0966-0941 Google Scholar

30. 

C. M. Bishop, Pattern Recognition and Machine Learning, Springer, New York (2006). Google Scholar

31. 

F. CamastraA. Vinciarelli, Machine Learning for Audio, Image and Video Analysis, Springer, London (2008). Google Scholar

32. 

R. O. DudaP. E. HartD. G. Stork, Pattern Classification, 2nd ed.John Wiley & Sons, New York (2001). Google Scholar

33. 

A. Zwiellyet al., “Discrimination between drug-resistant and non-resistant human melanoma cell lines by FTIR spectroscopy,” Analyst, 134 (2), 294 –300 (2009). http://dx.doi.org/10.1039/b805223a ANLYAG 0365-4885 Google Scholar

34. 

M. DiemP. GriffithJ. Chalmers, Vibrational Spectroscopy for Medical Diagnosis, John Wiley & Sons, New York (2008). Google Scholar

35. 

R. A. Fisher, “The use of multiple measurements in taxonomic problems,” Ann. Hum. Genet., 7 (2), 179 –188 (1936). ANHGAA 0003-4800 Google Scholar

36. 

G. M. JamesT. J. Hasti, “ Functional linear discriminant analysis for irregularly sampled curves,” J. R. Stat. Soc. Series B Stat. Methodol., 63 (3), 533 –550 (2001). http://dx.doi.org/10.1111/rssb.2001.63.issue-3 1369-7412 Google Scholar

37. 

A. GershoR. M. Gray, Vector Quantization and Signal Compression, Kluwer Academic Press, Dordrecht, the Netherlands (1992). Google Scholar

38. 

Z. MovasaghiS. RehmanRehman, “Fourier transform infrared (FTIR) spectroscopy of biological tissues,” Appl. Spectrosc. Rev., 43 (2), 134 –179 (2008). http://dx.doi.org/10.1080/05704920701829043 APSRBB 0570-4928 Google Scholar

39. 

R. K. Dukor, Handbook of Vibrational Spectroscopy, John Wiley and Sons, Chichester UK (2001). Google Scholar

40. 

H. P. WangH. C. WangY. J. Huang, “Microscopic FTIR studies of lung cancer cells in pleural fluid,” Sci. Total Environ., 204 (3), 283 –287 (1997). http://dx.doi.org/10.1016/S0048-9697(97)00180-0 STENDL 0048-9697 Google Scholar

41. 

H. Fabianet al., “A comparative infrared spectroscopic study of human breast tumors and breast tumor cell xenografts,” Biospectroscopy, 1 (1), 37 –45 (1995). http://dx.doi.org/10.1002/(ISSN)1520-6343 BIOSFS 1075-4261 Google Scholar

42. 

M. F. K. Funget al., “Pressure-tuning Fourier transform infrared spectroscopic study of carcinogenesis in human endometrium,” Biospectroscopy, 2 (3), 155 –165 (1996). http://dx.doi.org/10.1002/(SICI)1520-6343(1996)2:3<155::AID-BSPY2>3.3.CO;2-F BIOSFS 1075-4261 Google Scholar

43. 

B. RigasP. T. T. Wong, “Human colon adenocarcinoma cell lines display infrared spectroscopic features of malignant colon tissues,” Canc. Res., 52 (1), 84 –88 (1992). CNREA8 0008-5472 Google Scholar

44. 

G. Shettyet al., “Raman spectroscopy: evaluation of biochemical changes in carcinogenesis of oesophagus,” Br. J. Cancer, 94 (10), 1460 –1464 (2006). http://dx.doi.org/10.1038/sj.bjc.6603102 BJCAAI 0007-0920 Google Scholar

45. 

R. Eckelet al., “Characteristic infrared spectroscopic patterns in the protein bands of human breast cancer tissue,” Vib. Spectrosc., 27 (2), 165 –173 (2001). http://dx.doi.org/10.1016/S0924-2031(01)00134-5 VISPEK 0924-2031 Google Scholar

46. 

N. Fujiokaet al., “Discrimination between normal and malignant human gastric tissues by Fourier transform infrared spectroscopy,” Cancer Detect. Prev., 28 (1), 32 –36 (2004). http://dx.doi.org/10.1016/j.cdp.2003.11.004 CDPRD4 0361-090X Google Scholar

47. 

L. Chiribogaet al., “Infrared spectroscopy of human tissue. I. Differentiation and maturation of epithelial cells in the human cervix,” Biospectroscopy, 4 (1), 47 –53 (1998). BIOSFS 1075-4261 Google Scholar

48. 

S. Yoshidaet al., “Fourier transform infrared spectroscopic analysis of rat brain microsomal membranes modified by dietary fatty acids: possible correlation with altered learning behavior,” Biospectroscopy, 3 (4), 281 –290 (1997). http://dx.doi.org/10.1002/(ISSN)1520-6343 BIOSFS 1075-4261 Google Scholar

49. 

G. I. Dovbeshkoet al., “FTIR spectroscopy studies of nucleic acid damage,” Talanta, 53 (1), 233 –246 (1997). http://dx.doi.org/10.1016/S0039-9140(00)00462-8 TLNTA2 0039-9140 Google Scholar

50. 

P. G. L. AndrusR. D. Strickland, “Cancer grading by Fourier transform infrared spectroscopy,” Biospectroscopy, 4 (1), 37 –46 (1998). http://dx.doi.org/10.1002/(ISSN)1520-6343 BIOSFS 1075-4261 Google Scholar

51. 

N. J. Harrick, “Principles of internal reflection spectroscopy,” Internal Reflection Spectroscopy, 13 –66 (1979). Google Scholar

52. 

C. L. Curlet al., “Refractive index measurement in viable cells using quantitative phase-amplitude microscopy and confocal microscopy,” Cytometry, 65A (1), 88 –92 (2005). http://dx.doi.org/10.1002/(ISSN)1552-4930 CYTODQ 0196-4763 Google Scholar

53. 

E. Bogomolnyet al., “Attenuated total reflectance spectroscopy: a promising technique for early detection of premalignancy,” Analyst, 135 (8), 1934 –1940 (2010). http://dx.doi.org/10.1039/b920591h ANLYAG 0365-4885 Google Scholar

54. 

MATLAB, Natick, MA (2007). Google Scholar

55. 

M. J. Germanet al., “Infrared spectroscopy with multivariate analysis potentially facilitates the segregation of different types of prostate cell,” Biophys. J., 90 (10), 3783 –3795 (2006). http://dx.doi.org/10.1529/biophysj.105.077255 BIOJAU 0006-3495 Google Scholar

56. 

T. EvgeiouM. PontilA. Elisseeff, “Leave-one-out error, stability, and generalization of voting combination of classifiers,” Mach. Learn., 55 (1), 71 –97 (2004). http://dx.doi.org/10.1023/B:MACH.0000019805.88351.60 MALEEZ 0885-6125 Google Scholar

57. 

A. ElisseeffM. Pontil, “Leave-one-out error and stability of learning algorithms with applications,” Advances in Learning Theory: Methods, Models and Applications, 190 111 –124 IOS Press, Amsterdam, the Netherlands (2003). Google Scholar
© 2012 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2012/$25.00 © 2012 SPIE
Ahmad Salman, Elad Shufan, Itshak Lapidot, Ami Pomerantz, Mahmoud Huleihel, Leah Tsror, Raymond Moreh, and Shaul Mordechai "Identification of fungal phytopathogens using Fourier transform infrared-attenuated total reflection spectroscopy and advanced statistical methods," Journal of Biomedical Optics 17(1), 017002 (6 February 2012). https://doi.org/10.1117/1.JBO.17.1.017002
Published: 6 February 2012
Lens.org Logo
CITATIONS
Cited by 40 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Fungi

Principal component analysis

Infrared radiation

Fourier transforms

Spectroscopy

Statistical analysis

Absorption

Back to Top