Anda di halaman 1dari 8

Food Chemistry 194 (2016) 873880

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Analytical Methods

NIR detection of honey adulteration reveals differences in water spectral


pattern
Gyrgy Bzr a,b,, Rbert Romvri a, Andrs Szab a, Tams Somogyi a, Viktria les a,
Roumiana Tsenkova b,
a
Institute of Food and Agricultural Product Qualification, Faculty of Agricultural and Environmental Sciences, Kaposvr University, 40 Guba Sndor Str, Kaposvr 7400, Hungary
b
Biomeasurement Technology Laboratory, Graduate School of Agricultural Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan

a r t i c l e i n f o a b s t r a c t

Article history: High fructose corn syrup (HFCS) was mixed with four artisanal Robinia honeys at various ratios (040%)
Received 24 July 2014 and near infrared (NIR) spectra were recorded with a fiber optic immersion probe. Levels of HFCS adul-
Received in revised form 17 August 2015 teration could be detected accurately using leave-one-honey-out cross-validation (RMSECV = 1.48;
Accepted 22 August 2015
R2CV = 0.987), partial least squares regression and the 13001800 nm spectral interval containing absorption
Available online 28 August 2015
bands related to both water and carbohydrates. Aquaphotomics-based evaluations showed that unifloral
honeys contained more highly organized water than the industrial sugar syrup, supposedly because of
Keywords:
the greater variety of molecules dissolved in the multi-component honeys. Adulteration with HFCS caused
Near infrared spectroscopy
Aquaphotomics
a gradual reduction of water molecular structures, especially water trimers, which facilitate interaction with
Honey other molecules. Quick, non-destructive NIR spectroscopy combined with aquaphotomics could be used to
Fructose describe water molecular structures in honey and to detect a rather common form of adulteration.
Water 2015 Elsevier Ltd. All rights reserved.
Aliment
Adulteration
Fraud

1. Introduction to it. However, since there is a great demand for high quality honey
and consumers intend to pay high prices for these products, valu-
Honey is an important commodity in the food market and is able honeys have become a target of adulteration with cheap sweet
acknowledged as a healthy natural sweetener, which is consumed foreign components worldwide. Adulterated honeys are often
alone or as an ingredient in a variety of sweet products. The worlds labeled as natural and priced the same as pure honey, which is
average annual honey production was 1.5 billion kg between 2005 both fraudulent and unfair to consumers (Chen et al., 2011). In gen-
and 2010 (FAOSTAT, 2013). Since consumption of honey in Hun- eral, the adulteration of honey does not pose a health risk. How-
gary is very low (0.24 kg/capita/year), the country has become ever, it does influence market growth negatively by damaging
the second biggest honey exporter in the European Union, deliver- consumer confidence (Mehryar, 2011). Consequently, identifica-
ing 19 million kg per year to the market, covering more than 18% of tion and authentication of unadulterated honey is important for
the total EU export of honey. China and many European countries, processors, retailers and consumers, as well as regulatory authori-
particularly in the East, produce considerable amounts of False ties (Chen et al., 2011).
Acacia (Robinia pseudoacacia L.) unifloral honey, which is one of Large-scale honey adulteration appeared on the world market
the most valuable honey products on the European market. How- in the 1970s with the introduction of high-fructose corn syrup
ever, the characteristics of Robinia honey (i.e. liquid due to the high (HFCS) by the industry (Mehryar, 2011). Based on the pioneering
fructose content glucose:fructose = 1:1.41.7, very light colored investigations of Doner and White (1977), HFCS content in natural
and flavored) can be easily counterfeited (Persano Oddo & Piro, honey samples was described using mass spectroscopy (Doner and
2004). White, 1978), gasliquid chromatography (Doner, White, &
The European Commission has defined honey (European Phillips, 1979) and high pressure liquid chromatography (White,
Commission, 2002) and stipulated that nothing should be added 1980). Although these methods and their further developed appli-
cations can accurately detect honey adulteration, they have a num-
Corresponding authors at: Institute of Food and Agricultural Product Qualifi- ber of disadvantages that discourage practical application in
cation, Faculty of Agricultural and Environmental Sciences, Kaposvr University, 40 screening systems, i.e., the need for highly skilled labor,
Guba Sndor Str, Kaposvr 7400, Hungary (G. Bzr). time-consuming, destructive and ultrapure preparation and large
E-mail addresses: bazar@agrilab.hu (G. Bzr), rtsen@kobe-u.ac.jp (R. Tsenkova).

http://dx.doi.org/10.1016/j.foodchem.2015.08.092
0308-8146/ 2015 Elsevier Ltd. All rights reserved.
874 G. Bzr et al. / Food Chemistry 194 (2016) 873880

instrumental analysis. Thus, there is an obvious demand for devel- flowering in 2012. Liquid, cleaned, sterile, ion exchanged and
oping easy to use, rapid, non-destructive, and low-cost analytical filtered isoglucose syrup (HFCS) was produced in a high-
methods to detect and quantify HFCS adulteration in commercial temperature closed process, contained 40% fructose and 33%
honey lots. glucose at 18.7% of moisture. All samples were stored in dark glass
Near infrared (NIR) spectroscopy is a rapid, highly accurate, ver- vials at room temperature (25 C) for approximately 5 months,
satile, multi-analytical technique based on the electromagnetic until NIR analysis. Because of the lack of refrigeration slight chem-
absorption of organic compounds in the short wavelength infrared ical changes may have occurred in the samples. However, from a
range (7802500 nm). It can be applied to qualitative or quantita- practical point of view, this represents the real-life situation as
tive analyses of multiple components of a sample by a single mea- regards commercial honeys (Kelly et al., 2006). One day prior to
surement. Moreover, no reagents are required and no hazardous NIR scanning, individual honey samples were diluted randomly
waste is produced. The NIR technique has been applied in a wide with HFCS (n = 40, range of honey concentration = 10060%;
range of fields in the last decades and several of these applications mean SD = 80.79 12.89). Samples (about 50 g) were produced
are currently in use as routine analyses, even in on-line monitoring for each mixture by measuring the mass of the constituents (one
systems, both in agriculture and the food industry (Downey, 1996; of the Robinia honeys plus HFCS) into glass vials on a laboratory
Huang, Yu, Xu, & Ying, 2008; Prieto, Roeho, Lavn, Batten, & Andrs, scale and blending with a glass rod. HoneyHFCS mixtures were
2009). covered and stored at room temperature for 24 h in order to
The use of the NIR technique for food authentication was remove any bubbles produced during blending, since these could
reviewed by Reid, ODonnell, and Downey (2006). In regards to interfere with NIR scanning through light scattering. Four pure
honey, Cho and Hong (1998) and Ha, Koo, and Ok (1998) developed honeys, the HFCS sample, and randomly selected (n = 19) honey
NIR prediction method for conventional quality parameters of HFCS mixtures were tested for dry matter content after NIR scan-
Korean honeys using transflectance cuvettes with gold and ning, by means of conventional drying at 105 C for 4 h, in three
aluminum reflector back. Garca-Alvarez, Huidobro, Hermida, and replicates, and mean values were applied for NIR calibration. The
Rodrguez-Otero (2000) analyzed fructose, glucose and moisture standard error of laboratory (SEL) was determined for the reference
content of honeys, and reported proper results in calibration and data according to Workman and Mark (2006).
independent validation for all three components when the
calibration was generated using transflectance spectra. As an 2.2. Near infrared spectra
authentication tool, NIR spectroscopy has been applied success-
fully for rapid determination of the floral origin of honeys (Chen Pure and artificially adulterated honey samples in glass holders
et al., 2012; Escuredo, Gonzlez-Martn, Rodrguez-Flores, & Seijo, were scanned at room temperature (25 C) using a NIRSystems
2015). Transflectance NIR spectra of Irish artisanal honeys 6500 spectrometer (FOSS NIRSystems, Inc., Laurel, MD, USA)
adulterated with fructose and glucose were acquired by Downey, equipped with an OptiProbe fiber optic immersion sampling unit
Fouratier, and Kelly (2003) using a gold back reflector with with 2 mm layer thickness. Transflectance spectra (1100
0.1 mm sample thickness, and accurate classification models were 2500 nm) were recorded with a spectral step of 2 nm as an average
generated to reveal adulteration. Kelly, Petisco, and Downey (2006) of 32 scans. The acquisition of logR 1 spectra was performed with
investigated the detection of honey adulteration with sugar-beet VISION 2.51 software (FOSS NIRSystems, Inc., Laurel, MD, USA). The
invert syrup and HFCS using both qualitative and quantitative noise level of model 6500 monochromator was certified for
NIR models. Applying a fiber optic probe, successful classification 50 lAbs unit. Slightly increased noise level was experienced during
was performed by Chen et al. (2011) using FT-NIR spectra of performance tests because of the application of the fiber optic
unadulterated and adulterated Chinese honeys collected from probe (noise level of the 13001800 nm region <80 lAbs unit).
apiaries and purchased in local groceries. The sample set was scanned in two runs, on two consecutive days.
Water is one of the most studied compounds in physics and The samples were presented for scanning in random order in both
chemistry, the properties of which are explained by several differ- repeated series. Six consecutive spectra were recorded for each
ent structural models based on the number, nature and strength of sample at each time. Thus, one sample was represented by 12 spec-
hydrogen bonds (Segtnan, Sasic, Isaksson, & Ozaki, 2001). tra in total, from two repeated measurement days. The total num-
Aquaphotomics, as a new approach in NIR spectroscopy, considers ber of scanned samples (pure honeys, HFCS and mixtures) was
water as a multi-element system that can be described by its n = 41, and the total number of stored spectra was n = 492. During
multi-dimensional NIR spectra (Tsenkova, 2009). Since water- scanning, the OptiProbe was immersed into a sample and glass
related H-bonds are present in most natural samples, this analyti- holders were rotated between repeated spectral acquisitions in
cal approach, using perturbed water in different environments as a order to flush the measuring slit with fresh sample. The OptiProbe
mirror for the rest of the molecules in the sample, can be applied was washed with hot water and dried after each sample.
effectively in various fields (Kinoshita et al., 2012; Tsenkova,
Meilina, Kuroki, & Burns, 2009). 2.3. Spectral analysis
The objective of the present study was to develop an applicable
NIR model for screening unifloral Robinia honey using a fiber-optic Recorded spectra were exported from VISION in NSAS file for-
probe. Recent findings of aquaphotomics were implemented to mat and The Unscrambler 9.7 (CAMO Software AS, Oslo, Norway)
interpret the chemometric calibration models for measuring the was applied for data processing and analysis. Low energy light at
level of HFCS adulteration, and to extract important information longer wavelengths could not pass through the samples, meaning
on functionality of the honey related to its water structures. absorbance spectra above 1850 nm showed stray light. In the
shorter wavelength region (11001200 nm) undefined noise was
detected during basic spectral evaluation. Considering these
2. Materials and methods effects, the 13001800 nm range was selected for evaluation of
NIR data. Moving average smoothing was applied with 15 spectral
2.1. Honey and HFCS samples points before standard normal variate (SNV) transformation
(Barnes, Dhanoa, & Lister, 1989) in order to correct scatter effect.
Pure Robinia honey samples were collected from four Derivative spectra were calculated using the gap-segment method
geographic regions of Hungary, during periods of False Acacia (Norris, 2001) with various gap and segment sizes. Principal
G. Bzr et al. / Food Chemistry 194 (2016) 873880 875

component analysis (PCA) was used to describe the basic multidi- 0.11. Based on the almost 5% difference in moisture of the test
mensional characteristics of the NIR data matrix (Cowe & McNicol, materials, a strong negative correlation (r = 0.99) was found
1985). Calibration for HFCS or dry matter content was conducted between dry matter and HFCS content of the mixtures of the four
using principal component regression (PCR) and partial least individual honeys. The correlation was weaker (r = 0.81) when
squares regression (PLSR), and models were tested by cross- it was calculated for all the tested honeyHFCS mixtures together.
validation (Ns, Isaksson, Fearn, & Davies, 2002). In case of PCR,
the principal components are generated solely from the spectral 3.2. Basic spectral differences
data and are optimized to describe the highest proportion of vari-
ance of the spectral variables. In contrast, PLSR applies also the Line plots in Fig. 1 show the NIR spectra of the samples investi-
Y-variables of the predictable constituents and optimizes the latent gated. The 2nd derivative spectra were calculated with 5-point gap
variables of the model to describe as much amount of the variance and 5-point segment. The negative peaks of the 2nd derivative
of X-(spectral data) and Y-variables (reference data) as possible, at graphs expose the underlying peaks appearing overlapped in the
the same time. Leave-one-honey-out cross-validation was applied, raw spectrum.
where spectra for one of the four honeys and all of its mixtures The average spectrum of the four pure honeys and that of the
were omitted from the calibration, and the model was tested on HFCS sample was calculated. Raw average spectra and their 2nd
the remaining samples. The procedure was repeated iteratively derivatives are shown in Fig. 2. The mean spectrum of HFCS was
until all groups were used for testing. The precision and accuracy subtracted from the honeys and the difference is plotted to empha-
of the chemometric models were evaluated by the determination size spectral regions showing the biggest variation in the two test
coefficient (R2, R2CV) and the root mean square error (RMSE, RMSECV) materials. In this case, the averaging of spectra decreased the over-
of calibration and cross-validation, respectively (Ns et al., 2002). all noise level and, thus, a more sensitive setting for derivative cal-
Regression vectors of the calibration models were analyzed in culation was applied: 5-point smoothing, then 2nd derivatives
order to assign the characteristic water bands, water matrix coor- with 5-point gap and 1-point segment. This setting could find
dinates (WAMACs), showing considerable changes in honey mix- and visualize narrow bands within the spectra.
tures according to HFCS content. The variation of WAMACs As for band assignment it must be noted that the spectral step
describes the water spectral pattern (WASP), which can be visual- and bandpass of the applied spectrometer was 2 nm and 10 nm,
ized in aquagrams (Kinoshita et al., 2012). The star-chart applied respectively. Because of the application of the fiber optic probe,
for the aquagram displays average normalized absorbance values increased noise level (<80 lAbs) was experienced and, thus, it
from different sample groups at several water bands on the axes was necessary to involve smoothing procedures with several
originating from the center of the graph. Absorbance values at points in the chemometric models. Based on all these effects, small
the WAMACs were placed on the respective radial axes. Aquagrams shifts of the bands can be expected during assignment and com-
were prepared in Microsoft Office Excel 2010 (Microsoft Co., Red- parison with literature data.
mond, WA, USA). The peak of the derivative graphs in Figs. 1 and 2 are consistent,
despite being calculated differently, and accentuate the following
bands: 1356, 14341436, 1476, 1502, and 1580 nm, representing
3. Results and discussion water with less or more Hbonds (Rambla, Garrigues, & de la
Guardia, 1997; Segtnan et al., 2001; Xantheas, 1995) as well as
3.1. Dry matter based evaluation 1690, 17281730, and 1780 nm, which are CH stretching in sug-
ars (Golic, Walsh, & Lawson, 2003; Kelly et al., 2006; Workman,
The dry matter content of pure honey samples and HFCS was 2001). The signal at 1690 nm is characteristic for fructose
87.2 0.65 (mean SD) and 81.3%, respectively, while that of the (Williams and Norris, 2001) and indicates the marked difference
honeyHFCS mixtures (n = 19) was 86.4 0.71 (mean SD). Stan- in glucose:fructose ratio of the investigated HFCS and honeys (glu-
dard error of laboratory (SEL) for the reference measurement was cose:fructose = 1:1.21.3 and 1:1.41.7 in HFCS or Robinia honey,

Fig. 1. (a) Raw and (b) 2nd derivative spectra of four honeys, HFCS, and honeyHFCS mixtures (n = 492; blue spectra: HFCS, reddish spectra: pure and adulterated honeys).
876 G. Bzr et al. / Food Chemistry 194 (2016) 873880

Fig. 2. (a) Raw and (b) 2nd derivative average spectra of pure honeys and HFCS, and their subtracted spectra.

respectively). Application of this single point of the spectrum, how-


ever, is problematic since the fructose ratio of artisanal honeys
may vary greatly. The dominant peak of the subtracted spectrum
at 1418 nm, in Fig. 2, is related to the ascending part of the water
first overtone absorption peak (1st overtone of OH stretching)
(Osborne, Fearn, & Hindle, 1993). The peak position of the sub-
tracted spectrum is not exactly in the position of the characteristic
peak but on its side, showing that differences in this region origi-
nate from the fact that HFCS provides a water peak shifted to the
shorter wavelength regions, while this absorption band in honey
appears at longer wavelengths, without great change in the absor-
bance level. This reflects that spectra contain information mainly
on the water structure. Subtracted peaks of Fig. 2 show spectral
variation between HFCS and honey at the 1st overtone of CH
bonds of carbohydrates around 1690, 1728, and 1780 nm (Golic
et al., 2003; Kelly et al., 2006; Workman, 2001).
PCA score plot generated on all recorded spectra and the line
plot of the first principal component (PC) are shown in Fig. 3. As
the score plot depicts, the basic spectral variation of the samples
resembles the adulteration level. HFCS samples form a separate
group, while honeyHFCS mixtures are plotted in the order of
the mixing ratio, between pure honeys and HFCS. The first PC
describing 97% of the total variance of the NIR data is dominated
by the difference in the spectral properties of the test materials.
The subtracted spectrum of non-derivative spectra shown in
Fig. 2 is nearly identical with the first PC (Fig. 3b).

3.3. Calibration models on dry matter and purity

PCR and PLSR were applied for calibration on dry matter and
purity (pure honey content against HFCS) of pure and adulterated
honeys. Spectra of pure HFCS were not applied for any of the mod- Fig. 3. (a) PCA score plot and (b) loadings of the 1st PC calculated on the entire
sample set (n = 492) after 15-point smoothing and SNV transformation of the NIR
els. Analyses were run using different spectral intervals, i.e. the
spectra (13001800 nm).
13001800 nm region was divided in two parts: 13001600 nm
representing the first overtone region of water stretching vibra-
tions (OH bands), and 16001800 nm, predominantly represent-
ing the first overtone region of carbohydrates (CH bands). formed before or after 2nd derivation. There were no differences
Second derivatives as spectral pretreatment were applied on in the appearance of the spectra when SNV was applied prior or
SNV transformed spectra. Since the order of the applied spectral post derivation, and only slight differences were observed in the
treatments may affect the results (Fearn, 2008), SNV was per- results, while regression vectors showed the same bands in both
G. Bzr et al. / Food Chemistry 194 (2016) 873880 877

Fig. 4. (a, c) Regression vectors of PLSR models with (b, d) the results of calibrations and cross-validations on (a, b) dry matter content and (c, d) purity (honey percentage
against HFCS) of adulterated honey samples using the 2nd derivative of the smoothed (15 points) and SNV transformed 13001800 nm spectral interval (nr. of applied
factors = 5).

cases. Thus, only results achieved with SNV made prior to the revealed signals from the carbohydrate region (16001800 nm) and
calculation of derivatives are shown here. the error (RMSECV) decreased to 1.5%. It is remarkable, however, that
Both PCR and PLSR models were developed using the aforemen- the addition of water regions and application of the whole range
tioned three spectral ranges, applying three different spectral (13001800 nm), generally, increased the accuracy and precision of
treatments: (a) smoothing with 15 points and SNV, (b) 2nd deriva- the models. This shows that water bands hold additional information
tive with 5 point gap and 15 point segment, (c) smoothing with 15 on the constituents of the mixtures.
points, SNV, and 2nd derivative with 5 point gap and 15 point seg- Kelly et al. (2006) investigated the detection of honey adulter-
ment. Only results of PLSR are summarized here, but the PCR mod- ation with HFCS using transflectance spectra of 0.1 mm thick layer
els and their regression coefficient vectors were also evaluated in of samples recorded with a camlock cell and gold reflector. The
order to assign the important spectral bands. applied sample set comprised 83 authentic honeys, of which 10
Fig. 4 shows the calibration and cross-validation results of the were adulterated with HFCS at the inclusion level of 10%, 30%,
PLSR models achieved with the 2nd derivatives of the smoothed 50% and 70%. The total number of samples in HFCS calibration
and SNV transformed spectra in the 13001800 nm interval. The was 123. PLS regression was used to predict the adulteration level,
regression vectors shown in Fig. 4 accentuate the bands having and the best coefficient of determination was 0.72 for HFCS con-
dominance in the calibrations. Results from the PLSR calibration tent, with an RMSECV of 11.9, achieved with the 2nd derivative
models with different spectral intervals are summarized in Table 1. spectra of 11002498 nm region. Compared to the results of
RMSECV results on dry matter content are comparable with the Kelly et al. (2006), the precision and accuracy of the models
SEL value of the reference method, which shows the strength of the reported in the present study are more robust, although spectra
applied NIR models. Calibration models provided results at high were recorded with a fiber optic probe, but there was a great differ-
regression and low error level on purity. Results achieved with ence in sample number and the range and variance of the adulter-
smoothed and SNV transformed non-derivative spectra were the ation level and, thus, results are not strictly comparable.
most accurate when the 13001800 nm spectral interval was Averaging repeated scans is a useful method to reduce errors
applied: error of cross-validation (RMSECV) was less than 1.5%. originating from sampling, environmental or instrumental noise
Using these spectral treatments, characteristic region of carbohy- (Norris, 2002). In this case, averaging was applied to determine
drates (16001800 nm) led to weaker results on purity than that the influence of noise on the results. The six consecutive spectra
of water (13001600 nm). Applying only the 1st overtone region of each sample were averaged in pairs: average of 1st and 6th,
of water, the error was 4% (R2CV = 0.91), still less than 1/3 of the stan- 2nd and 5th, 3rd and 4th was calculated. Applying this procedure,
dard deviation of the reference values. The application of derivatives the possible distorting effect of consecutive scanning was also
878 G. Bzr et al. / Food Chemistry 194 (2016) 873880

Table 1 in Robinia honey. These bands are the WAMACs for this complex
Results of PLSR calibrations and cross-validations on dry matter content and purity of system, and the variation in absorbance values at these bands
adulterated honey mixtures applying different spectral regions and various spectral
treatments.
describe the WASP of the sample sets. In our application, the aqua-
gram shows a relative fingerprint of a sample, as it contains nor-
Range Y var.a Nr.b RSQc RMSEd RSQCVe RMSECVf malized absorbance values at specific water bands, and the graph
(a) Smoothing (15 points), SNV of a group depends on the dominance and characteristics of the
13001800 nm DMg 5 0.981 0.10 0.981 0.12 other samples. The similarity of honeys and their unlikeness from
Purityh 5 0.997 0.72 0.988 1.49
13001600 nm DM 2 0.937 0.19 0.924 0.23
HFCS can be seen very clearly using this visualization method.
Purity 4 0.969 2.25 0.910 4.01 Aquagram in Fig. 5a shows the pure honey samples against the
16001800 nm DM 4 0.979 0.11 0.958 0.17 HFCS sample, while Fig. 5b demonstrates the aquagram of adulter-
Purity 2 0.946 2.98 0.892 4.49 ated mixtures of one honey (H1), plotted together with the mean
(b) 2nd derivative (gap = 5 points, segment = 15 points) graph of pure honeys and HFCS.
13001800 nm DM 2 0.975 0.12 0.945 0.20 The bands accentuated by the regression vectors of the devel-
Purity 6 0.996 0.81 0.985 1.62
oped models, and applied for aquagrams, are consistent with the
13001600 nm DM 3 0.971 0.13 0.958 0.17
Purity 4 0.934 3.31 0.889 4.49 assigned peaks of the derivative spectra (Figs. 1 and 2), although
16001800 nm DM 5 0.978 0.11 0.955 0.18 the calculation method of derivatives was not the same during
Purity 4 0.996 0.86 0.989 1.52 the different procedures this demonstrates the stability of the
(c) Smoothing (15 points), SNV, 2nd derivative (gap = 5 points, segment = 15 applied methodology. Basically, the 13221326 nm and 1358
points) 1366 nm regions represent non or less H-bonded water, where sig-
13001800 nm DM 5 0.981 0.10 0.984 0.11 nals of free OH vibrations dominate (Xantheas, 1995), but this
Purity 5 0.995 0.90 0.987 1.48
13001600 nm DM 2 0.949 0.17 0.943 0.20
region is also related to highly organized water in water solvation
Purity 4 0.917 3.72 0.816 5.74 shell as it corresponds to the 1st overtone of those bands described
16001800 nm DM 2 0.976 0.12 0.960 0.17 in relation with aqueous protons (Headrick et al., 2005), and free
Purity 4 0.993 1.00 0.987 1.54 OH in water clusters (Mizuse & Fujii, 2012). Temperature studies of
a
Reference variable of calibration model. liquid water demonstrated that the signals of less H-bonded water
b
Number of latent variables applied in the calibration model. appear around 1412 nm, and water having more H-bonds provides
c
Determination coefficient of calibration. signals at 1462 and 1490 nm (Collins, 1925; Segtnan et al., 2001).
d
Root mean square error of calibration.
e
Segtnan et al. (2001) also assigned an intermediate stage of water
Determination coefficient of cross-validation.
f
Root mean square error of cross-validation.
at 1438 nm. Bands above 1500 nm are related to the 1st overtone
g
Dry matter, in percentage (mean SD = 86.57 0.75; SEL = 0.11; nr. of of ice-like clusters of water, having highly organized structure of
samples = 23; nr. of spectra = 276). molecules (Headrick et al., 2005; Mizuse and Fujii, 2012), and such
h
Pure honey content of sample against HFCS, in percentage structure of water is expected around the hydrated macro-
(mean SD = 80.79 12.89; nr. of samples = 40; nr. of spectra = 480).
molecules. Rambla et al. (1997) assigned 15801590 nm region
for aqueous solutions of fructose (1583 nm), sucrose (1584 nm)
reduced. The PCR and PLSR models described above were per- and glucose (1587 nm), which are the main sugars of honey.
formed on the averaged data set. Results were slightly weaker than Different honeys in Fig. 5a show relatively large variation at
those obtained with the original spectra, and regression vector 14641472 nm and 15001504 nm bands. It can be assumed that
peaks appeared at the same positions as without averaging. This honeys functionality is highly related to these bands representing
showed that the calculation of models was not greatly influenced altering number of H-bonds in dynamic water structure that is the
by random noise or sample-light interactions, and the assigned essential form of water in biological systems (Chandler, 2002). It is
bands contained reliable information. also notable how these bands changed with thinning (Fig. 5b).
The peak positions in the regression coefficient vectors of the Especially the 14641472 nm band of the active water trimers
related models were collected to summarize the bands having (Tsenkova, 2009) decreased when honeys were adulterated with
the biggest weights in calibration on dry matter or HFCS content. industrial fructose syrup and, thus, structures which characterize
The accentuated regions were very similar in case of PLSR and functional water facilitating interactions with other molecules
PCR models. Bands having the highest importance in the calibra- reduced gradually during adulteration with HFCS.
tion on the adulteration level were not analogous with those of The aquagrams of the present study demonstrate evidence on
the calibration model on dry matter content (Fig. 4). This difference the differences in the water structure of the investigated unifloral
was observed in both PLSR and PCR models, and may reflect that Robinia honeys and the applied HFCS. The orientation of the graphs
although dry matter content (or moisture) can be predicted per- suggest that honeys contain larger amount of highly organized
fectly by NIR models, and it highly correlates with the adulteration water than the industrial fructose syrup, supposedly because of
level, NIR spectra contain more information on the level of adulter- the larger variety of molecules (proteins, fibers, vitamins, minerals)
ation than solely the moisture content of the samples. This dissolved in the multicomponent system of honey, while HFCS has
extended information is represented in the chemometric models relatively large amount of free OH in unstructured water with less
on HFCS concentration. Presumably, the additional information H-bonds. This difference explains the increased liquidity of HFCS
may lay in the structure of water molecular system related to compared to Robinia honeys, even at a lower fructose ratio (glu-
moisture and purity, as it can be seen in the peak shifts mentioned cose:fructose = 1:1.21.3 or 1:1.41.7 in HFCS or Robinia honey,
at Fig. 2. respectively). The shift of the dominant water peak described at
Figs. 2 and 3 also supports that the investigated HFCS contains less
H-bonded water than the Robinia honeys.
3.4. Aquaphotomics based evaluation of water structural changes in It must be noted that fructose and glucose show considerable
adulterated honeys differences in the applied 13001600 nm NIR range, i.e. both
mono-saccharides provide signals because of their OH bonds,
Based on the consistency of assigned bands of the subtracted but dominant regions of fructose appear at 14201440 nm, while
spectra and the regression coefficient vectors, ten characteristic those of glucose appear at 15001540 nm (Williams and Norris,
ranges can be defined which are responsible for HFCS detection 2001). These characteristics are not represented in the aquagrams
G. Bzr et al. / Food Chemistry 194 (2016) 873880 879

Fig. 5. (a) Aquagram of the investigated pure Robinia honeys and HFCS. (b) Aquagram of the adulterated mixtures of one honey (H1-purity %) plotted with the mean graph of
the pure honeys and graph of HFCS.

since the changes of the aquagrams are not appearing according to absorption bands of both water and carbohydrates. Using NIR tech-
the ratio and spectral differences of the fructose and glucose in the nique and aquaphotomics, it has been demonstrated that there is a
composites (i.e. the ratio of fructose is higher in honey than in well describable difference in the water molecular structure in
HFCS, still, HFCS is shifted to the 14201440 nm region). We Robinia honey and HFCS. The aquagrams presented in this study
assume that information of water had the most impact in the PLSR emphasize that the investigated unifloral Robinia honeys contain
and PCR calibrations on HFCS adulteration of Robinia honeys, and larger amount of highly organized water than the added industrial
the aquagrams are presenting this information effectively. fructose syrup. The difference may be caused by the larger variety
of molecules dissolved in the multicomponent system of honeys,
4. Conclusions while the simpler matrix of HFCS may have relatively large amount
of less H-bonded, unstructured water. Adulteration of Robinia hon-
Although strong negative correlation was found between the eys with HFCS caused gradual reduction of water molecular struc-
dry matter content and the adulteration level of Robinia honeys tures facilitating the interactions with other molecules. This
mixed with HFCS, the conventional measurement of the dry matter reasoning elucidates the understanding of honeys functionality
content can hardly be used as a quick adulteration monitoring and, therefore, needs further and more detailed investigations.
technique since this measure is allowed to vary in a relatively wide
range even in artisanal Robinia honeys. Based on the NIR spectral Acknowledgements
information of honeys recorded with fiber optic probe, quick ana-
lytical tests can be developed to detect aliment adulteration. The This work is dedicated to the memory of Kroly Kaffka. The
most accurate NIR models for predicting adulteration level with financial support of JSPS Postdoctoral Fellowship for Foreign
the lowest cross-validation error (RMSECV = 1.48%) were achieved Researchers (P12409) and TMOP 4.2.2.-B research Grant is highly
within the whole spectral range of 13001800 nm, containing the acknowledged. Authors would like to thank the assistance of the
880 G. Bzr et al. / Food Chemistry 194 (2016) 873880

Institute of Apiculture and Bee Biology at Research Centre for Farm Kelly, J. D., Petisco, C., & Downey, G. (2006). Potential of near infrared transflectance
spectroscopy to detect adulteration of Irish honey by beet invert syrup and high
Animal Gene Conservation, and Hungrana Starch and Isosugar
fructose corn syrup. Journal of Near Infrared Spectroscopy, 14, 139146.
Manufacturing and Trading Co. Ltd. Kinoshita, K., Miyazaki, M., Morita, H., Vassileva, M., Tang, C., Li, D., ... Tsenkova, R.
(2012). Spectral pattern of urinary water as a biomarker of estrus in the giant
References panda. Scientific Reports, 2, 856. http://dx.doi.org/10.1038/srep00856.
Mehryar, L. (2011). Honey and honey adulteration detection: a review. In 11th
International Congress on Engineering and Food. Athens, Greece.
Barnes, R. J., Dhanoa, M. S., & Lister, S. J. (1989). Standard normal variate Mizuse, K., & Fujii, A. (2012). Tuning of the internal energy and isomer distribution
transformation and de-trending of near-infrared diffuse reflectance spectra. in small protonated water clusters H+(H2O)48: an application of the inert gas
Applied Spectroscopy, 43, 772777. messenger technique. The Journal of Physical Chemistry A, 116, 48684877.
Chandler, D. (2002). Hydrophobicity: Two faces of water. Nature, 417, 491. Ns, T., Isaksson, T., Fearn, T., & Davies, T. (2002). A user-friendly guide to
Chen, L., Wang, J., Ye, Z., Zhao, J., Xue, X., Heyden, Y. V., & Sun, Q. (2012). multivariate calibration and classification. Chichester: NIR Publications.
Classification of Chinese honeys according to their floral origin by near infrared Norris, K. H. (2001). Understanding and correcting the factors which affect diffuse
spectroscopy. Food Chemistry, 135, 338342. transmittance spectra. NIR News, 12(3), 69.
Chen, L., Xue, X., Ye, Z., Zhou, J., Chen, F., & Zhao, J. (2011). Determination of Chinese Norris, K. H. (2002). More on the factors affecting diffuse transmittance and
honey adulterated with high fructose corn syrup by near infrared spectroscopy. reflectance spectra. NIR News, 13(3), 811.
Food Chemistry, 128, 11101114. Osborne, B. G., Fearn, T., & Hindle, P. H. (1993). Practical NIR spectroscopy with
Cho, H. J., & Hong, S. H. (1998). Acacia honey quality measurement by near infrared applications in food and beverage analysis (2nd ed.). Harlow, New York: Longman
spectroscopy. Journal of Near Infrared Spectroscopy, 6, A329A331. Scientific & Technical, Wiley.
Collins, J. R. (1925). Change in the infra-red absorption spectrum of water with Persano Oddo, L., & Piro, R. (2004). Main European unifloral honeys: Descriptive
temperature. Physical Review, 26, 771779. sheets. Apidologie, 35, S38S81.
Cowe, I. A., & McNicol, J. W. (1985). The use of principal components in the analysis Prieto, N., Roeho, R., Lavn, P., Batten, G., & Andrs, S. (2009). Application of near
of near-infrared spectra. Applied Spectroscopy, 39, 257266. infrared reflectance spectroscopy to predict meat and meat products quality: A
Doner, L. W., & White, J. W. (1977). Carbon-13/carbon-12 ratio is relatively uniform review. Meat Science, 83, 175186.
among honeys. Science, 197, 891892. Rambla, F. J., Garrigues, S., & de la Guardia, M. (1997). PLS-NIR determination of
Doner, L. W., White, J. W., & Phillips, J. G. (1979). Gasliquid chromatographic test total sugar, glucose, fructose and sucrose in aqueous solutions of fruit juices.
for honey adulteration by high fructose corn sirup. Journal of Association of Analytica Chimica Acta, 344, 4153.
Official Analytical Chemists, 62, 186189. Reid, L. M., ODonnell, C. P., & Downey, G. (2006). Recent technological advances for
Downey, G. (1996). Authentication of food and food ingredients by near infrared the determination of food authenticity. Trends in Food Science and Technology,
spectroscopy. Journal of Near Infrared Spectroscopy, 4, 4761. 17, 344353.
Downey, G., Fouratier, V., & Kelly, J. D. (2003). Detection of honey adulteration by Segtnan, V. H., Sasic, S., Isaksson, T., & Ozaki, Y. (2001). Studies on the structure of
addition of fructose and glucose using near infrared transflectance water using two-dimensional near-infrared correlation spectroscopy and
spectroscopy. Journal of Near Infrared Spectroscopy, 11, 447456. principal component analysis. Analytical Chemistry, 73, 31533161.
Escuredo, O., Gonzlez-Martn, M. I., Rodrguez-Flores, M. S., & Seijo, M. C. (2015). Tsenkova, R. (2009). Aquaphotomics: dynamic spectroscopy of aqueous and
Near infrared spectroscopy applied to the rapid prediction of the floral origin biological systems describes peculiarities of water. Journal of Near Infrared
and mineral content of honeys. Food Chemistry, 170, 4754. Spectroscopy, 17, 303314.
European Commission (2002). Council Directive 2001/110/EC (20 December 2001) Tsenkova, R., Meilina, H., Kuroki, S., & Burns, D. H. (2009). Near infrared
relating to honey. Official Journal of the European Union, L010, 4752. spectroscopy using short wavelengths and leave-one-cow-out corss-
FAOSTAT (Statistical Database of the Food and Agriculture Organization of the validation for quantification of somatic cells in milk. Journal of Near Infrared
United Nations). URL http://faostat.fao.org/ Accessed 11.12.13. Spectroscopy, 17, 345351.
Fearn, T. (2008). The interaction between standard normal variate and derivatives. White, J. W. J. R. (1980). Detection of honey adulteration by carbohydrate analysis.
NIR News, 19(7), 1617. Journal of the Association of Official Analytical Chemistry, 66, 1118.
Garca-Alvarez, M., Huidobro, J. F., Hermida, M., & Rodrguez-Otero, J. L. (2000). White, J. W., & Doner, L. W. (1978). Mass spectroscopic detection of high-fructose
Major components of honey analysis by near-infrared transflectance corn sirup in honey by use of 18C/12C ratio: Collaborative study. Journal of the
spectroscopy. Journal of Agricultural and Food Chemistry, 48, 51545158. Association of Official Analytical Chemistry, 61, 746750.
Golic, M., Walsh, K., & Lawson, P. (2003). Short-wavelength near-infrared spectra of Williams, P., & Norris, K. (2001). Near-infrared technology in the agricultural and food
sucrose, glucose, and fructose with respect to sugar concentration and industries (2nd ed.). Minnesota, USA: American Association of Cereal Chemists,
temperature. Applied Spectroscopy, 57, 139145. Inc.
Ha, J., Koo, M., & Ok, H. (1998). Determination of the constituents of honey by near Workman, J. (2001). Handbook of organic compounds. London: Academic Press.
infrared spectroscopy. Journal of Near Infrared Spectroscopy, 6, A367A369. Workman, J., & Mark, H. (2006). Limitations in analytical accuracy, Part I: Horwitzs
Headrick, J. M., Diken, E. G., Walters, R. S., Hammer, N. I., Christie, R. A., Cui, J., ... trumpet. Spectroscopy, 21, 1824.
Jordan, K. D. (2005). Spectral signatures of hydrated proton vibrations in water Xantheas, S. S. (1995). Theoretical study of hydroxide ionwater clusters. Journal of
clusters. Science, 308, 17651769. the American Chemical Society, 117, 1037310380.
Huang, H., Yu, H., Xu, H., & Ying, Y. (2008). Near infrared spectroscopy for on/in-line
monitoring of quality in foods and beverages: A review. Journal of Food
Engineering, 87, 303313.

Anda mungkin juga menyukai