You are on page 1of 13

Journal of Food Engineering 64 (2004) 9–21

www.elsevier.com/locate/jfoodeng

Composition, thermal and rheological behaviour


of selected Greek honeys
Athina Lazaridou a, Costas G. Biliaderis a,*, Nicolaos Bacandritsos b,
Anna Gloria Sabatini c
a
Laboratory of Food Chemistry and Biochemistry, Food Science and Technology Department, School of Agriculture,
Aristotle University of Thessaloniki, Thessaloniki, Greece 541 24
b
Institute of Veterinary Research of Athens, N.AG.RE.F., 25 Neapoleos Street, Agia Paraskevi, 153 01 Athens, Greece
c
Instituto Nazionale di Apicoltura, Via di Saliceto 80, I-41128 Bologna, Italy
Received 6 May 2003; accepted 13 September 2003

Abstract
Several chemical and physicochemical properties (sugar composition, water content, water activity, colour, viscosity, thermal
properties) were determined for 33 Greek honeys from different botanical and geographical origin. The water content and water
activity values varied within 13.0–18.9 g/100 g and 0.528–0.663, respectively. Steady shear and dynamic rheological tests revealed
Newtonian behaviour for all samples examined over the temperature range of 20–60 C. The steady shear viscosity (g) and loss
modulus (G00 ) were inversely related to the water content of honey. The temperature dependence of viscosity followed both the
Arrhenius and the Williams–Landel–Ferry models; for the latter model the viscosity data of different samples fitted very well into a
common master curve. The glass transition temperature (Tg ) of honeys, as determined by differential scanning calorimetry, varied
between )34 and )47 C depending on their composition. The plasticizing action of water on honey solids was evident for native
samples as well as among diluted and concentrated honeys; Tg decreased with increasing water content. Despite a broad variation in
sugar composition among the samples, the Tg values vs. water content fitted reasonably well to the Gordon–Taylor empirical
equation.
 2003 Elsevier Ltd. All rights reserved.

Keywords: Honey; Moisture content; Water activity; Colour; Rheology; Arrhenius model; Williams–Landel–Ferry model; Glass transition

1. Introduction ously at room temperature in the form of glucose


monohydrate. Crystallization of honey, commonly
Honey, the viscous and aromatic product prepared called granulation, is an undesirable process in liquid
by bees, mainly from the nectar of flowers or honeydew, honey because it affects the textural properties, making
is a concentrated solution of various sugars. Honey it less appealing to the consumer. Moreover, in many
contains fructose and glucose (60–85%) as the predomi- cases, crystallization of honey results in increased
nant monosaccharides, maltose and sucrose as the most moisture of the liquid phase which can allow naturally
important disaccharides, melezitose as the main trisac- occurring yeast cells to multiply causing fermentation of
charide and other low molecular weight oligosaccha- the product (Doner, 1977). Water content as well as
rides (Doner, 1977; Doner & Hicks, 1982). The water activity are the major factors that influence the
composition of honey (sugars and moisture content) is keeping quality or storability of honey.
responsible for many of the physicochemical properties Sensory and physicochemical properties are very
of honey, such as viscosity, hydroscopicity, and granu- important parameters in determining the quality and
lation. Most honeys are supersaturated solutions of acceptability of honey and many studies have been
glucose, which have a tendency to crystallize spontane- devoted to explore such determinants of product qua-
lity (Al-Khalifa & Al-Arify, 1999; Anupama, Bhat, &
*
Corresponding author. Tel.: +2310-471467/+30-310-998-785; fax:
Sapna, 2003; Bath & Singh, 1999; Popek, 2003; Singh &
+2310-471257/+30-310-471-457. Bath, 1997). The composition and properties of honey
E-mail address: biliader@agro.auth.gr (C.G. Biliaderis). vary with the floral and honeydew sources utilized by
0260-8774/$ - see front matter  2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2003.09.007
10 A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21

Nomenclature

a , b chromatic components (red, yellow) T temperature (C, K)


aw water activity Tg glass transition temperature
C1 , C2 coefficients of WLF equation Tg1 glass transition of dry sample
Ea activation energy (kJ mol1 ) Tg2 glass transition of glassy water
G0 storage modulus (Pa) w2 weight fraction of water
G00 loss modulus (Pa) WLF Williams–Landel–Ferry equation
G–T Gordon–Taylor equation c_ shear rate (s1 )
k constant g viscosity (Pa s)
L lightness component g complex viscosity (Pa s)
p probability level gTg sample viscosity at Tg
r correlation coefficient r shear stress (Pa)
R universal gas constant (8.314 J mol1 K1 ) x angular frequency (rad s1 )

honey-bees, as well as by regional and climatic condi- 2. Materials and methods


tions. Some physicochemical parameters have already
been studied for their use in the identification of the 2.1. Samples
botanical and geographical origin of honey (Gomez
Barez et al., 2000; Popek, 2002; Terrab, Diez, & Here- Thirty-three honey samples were provided by bee-
dia, 2002). keepers with guaranteed botanic origin. These honeys
The rheological behaviour of honey has been investi- were divided into four main groups according to the
gated for shelf-life, proper handling, packing and pro- variety type: honeydew from pine, honeydew from fir,
cessing issues (White, 1978). The rheological properties mixed floral type, and floral from orange blossom.
of honey, like many other physical properties, depend on Samples were sourced from different geographical areas
many factors, including composition and temperature. of Greece; pine honeydew from Thasos (6 samples),
The Arrhenius model is widely used for temperature Halkidiki (6 samples) and Evia (2 samples), fir honey-
dependence of a property but models such Williams– dew from Helmos (2 samples) and Vytina (8 samples),
Landel–Ferry (WLF) that include the glass transition mixed floral from Livadia (3 samples), and orange
temperature (Tg ) as a parameter, have proved equally blossom floral from Argos (4 samples) and Sparti (2
useful for the viscosity–temperature relationship of food samples). The botanical/geographical identification was
systems (Ollett & Parker, 1990; Soesanto & Williams, based on their colour, aroma, taste and location of the
1981; Williams, Landel, & Ferry, 1955). Depending on hives.
concentration, and heating and cooling rates, aqueous
carbohydrate solutions exhibit several thermal events, 2.2. Moisture content and water activity, sugar composi-
the most important being the Tg . At the glass transition tion
temperature, an amorphous material changes from the
rubbery to the glassy state upon cooling, leading to the Refractive indices of honey samples were measured
formation of a hard solid. As the stability of foods is using a refractometer at 20 C and corresponding
mainly dependent on the water content and because Tg is moisture content (%) was calculated using the relation-
also highly sensitive to this parameter, the glass transi- ship between refractive index and water content
tion concept has been proposed as a useful tool for un- (AOAC, 1990).
derstanding the mechanisms of deteriorative processes in Water activity (aw ) of samples was measured at 20 C
food products and for controlling their shelf-life (Slade & using an Aqualab 3TE water activity meter (Decagon
Levine, 1991). Indeed, the glass transition temperature is Devices Inc., Pullman, WA, USA). The determination
often considered as a reference temperature; below Tg , of aw values was performed twice; before and after the
the food is expected to be stable and above this tem- heating of the samples at 50 C for 1 h. This heat
perature, the difference (T  Tg ) between Tg and the treatment was carried out to dissolve crystals or nuclei,
storage temperature T is assumed to control the rate of which might be present in honey and can influence the
physical, chemical and biological changes (Roos, 1995). water activity of the system.
The present study was undertaken to determine cer- The sugar composition was determined by a gas
tain physicochemical properties of selected Greek hon- chromatography (GC) method as described by Sabatini,
eys, and explore some relationships between them. Marcazzan, Colombo, Carpana, and Serra (2001).
A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21 11

2.3. Colour Temperature effects on steady shear viscosity were


analysed using the Arrhenius relationship:
Colour was determined by a Metertech UV/VIS
SP8001 spectrophotometer (Metertech Inc., Taipei, g ¼ g0 eðEa =RT Þ ð4Þ
Taiwan) and a Minolta Dimage 5 digital camera. Sam- where g is the viscosity at temperature T , g0 is a pre-
ples were heated at 50 C for 1 h before measurement to exponential factor, Ea is the activation energy for flow, R
ensure melting of any possibly formed crystals. Colour is the perfect gas constant and T is the absolute tem-
was determined as absorbance at 420 nm after dilution perature (K).
of honey with distilled water at a ratio 1:5 (Bath & The temperature dependence of honey viscosity was
Singh, 1999). Before measurement, mixtures of honey also described using the WLF model (Williams et al.,
and water were heated for better mixing and filtered for 1955):
removal of any coarse particles (plant residues, pollen), !
which may also influence the colour. Images of samples g C1 ðTg  T Þ
log ¼ ð5Þ
were taken by the digital camera with a proper lighting gT g C2 þ ðTg  T Þ
system as described by Papadakis, Abdul-Malek,
Kamdem, and Yam (2000). L , a and b colour para- where Tg is the glass transition temperature, g is the
meters (CIE, 1976) were obtained using the Photoshop viscosity at temperature T , gTg is the viscosity of sample
software (v6.0, Adobe Systems Inc., San Jose, CA). L is at Tg and C1 and C2 are the WLF constants. Experi-
the luminance or lightness component, which ranges mental data were fitted to the model using the Table
from 0 to 100, and a and b are the two chromatic Curve 2D software (v4.0, SPSS Inc., Chicago, IL).
components, which range from )120 to 120 (a from
green to red and b from blue to yellow) (Adobe 2.5. Differential scanning calorimetry
Photoshop 5.0 User Guide for Machintosh and
Windows, 1998). The software uses a scale, ranging The glass transition temperature (Tg ) of honeys was
from 0 to 255, to characterize Lightness, as well as the determined by differential scanning calorimetry (DSC)
values of a and b. To convert these parameters to L , a , using a PL DSC-Gold calorimeter (Polymer Labs. Ltd,
and b the following formulas were used: Epsom, UK). Temperature calibration was made with
cyclohexane, dodecane and octane, whereas heat flow
L ¼ ðLightness=250Þ  100 ð1Þ
calibration was made by reference to the known melting
a ¼ ð240a=255Þ  120 ð2Þ enthalpy of indium and gallium (purity 99.99%) from
Goodfellows Metals (Biliaderis, Lazaridou, & Arvani-
b ¼ ð240b=255Þ  120 ð3Þ toyannis, 1999).
For studying the effect of moisture on Tg , 14 honey
samples were chosen covering a moisture content range
2.4. Rheology of 13.0–18.9%. For increasing the above range the
sample with the lowest moisture content (13%) was
Rheological properties of honey were studied by a either diluted with water or concentrated under vacuum
rotational Physica MCR 300 rheometer (Physica Mess- (at 50 C) to the following levels of moisture content:
technic GmbH, Stuttgart, Germany) using a concentric 26.9%, 23.4%, 20.0%, 16.5%, 11.9%, 11.0% and 10.2%.
cylinder (diameter of cup and bob, 28.92 and 26.66 mm, The native honey samples as well as the diluted and
respectively) geometry; temperature was regulated by a concentrated samples were hermetically sealed into
Paar Physica circulating bath and a controlled peltier stainless steel pans (40–45 mg) and analysed by calo-
system (TEZ 150P/MCR) with an accuracy of ±0.1 C. rimetry under continuous flow of dry N2 gas (20
The data of the rheological measurements were analyzed ml min1 ) to avoid condensation of moisture. First, the
with the supporting rheometer software US200 V2.21. pans were heated from +20 to +50 C at a heating rate
All honey samples were heated at 50 C for 1 h before of 10 C min1 and kept at +50 C for 3–5 min to ensure
rheological measurements to melt any crystals present the melting of any crystals and reach at thermodynamic
and to remove the air bubbles, factors that can influence equilibrium. The samples were then quenched–cooled
the viscosity of honey. Two types of measurements were with liquid N2 to )100 C and reheated to +50 C at the
performed: (a) flow behaviour by measuring steady same heating rate (10 C min1 ). The Tg was determined
shear viscosity (g) and shear stress (r) over a range of in the latter heating scans as the onset temperature of
_ of 0.1–500 s1 at 20, 30, 40, 50 and 60 C;
shear rates (c) the step-like decrease in the heat flow.
and (b) oscillatory measurements to obtain the storage Data analysis to fit experimental values of Tg to the
and loss moduli (G0 , G00 ) and complex viscosity (g ) at a empirical Gordon–Taylor (G–T) model (Gordon &
strain level of 0.1% and a range of angular frequencies of Taylor, 1952) was performed using the Table Curve 2D
3–300 rad s1 at 20 C. software:
12 A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21

w1 Tg1 þ kw2 Tg2 eys). Among the disaccharides, maltose was the most
Tg ¼ ð6Þ abundant one, ranging between 1.9% and 6.7%. The
w1 þ kw2
relatively low levels of sucrose for most samples indicate
where Tg1 is the glass transition temperature of the that the selected honeys were at an advanced stage of
sample at zero moisture content, w1 is the weight frac- ripening. Several trisaccharides were also identified and
tion of dry solids, Tg2 is the glass transition temperature quantified, namely raffinose, erlose, melezitose, panose,
for glassy water, w2 is the weight fraction of water and k isomaltotriose and maltotriose. It was of interest to note
is a constant. The constructed G–T plots were based on that melezitose was present in relatively high amounts
the best data fitting to the equation (i.e. optimization for (9.1–14.4%) for most of the honeydew-fir samples.
both parameters, k and Tg1 ), where a Tg of )138 C was The colour of honey is related to the content of
used for water (Sugisaki, Suga, & Seki, 1968). phenolics, HMF, pollen and minerals (Perez-Arquille,
Conchello, Arino, Juan, & Herresa, 1994). The absor-
bance at 420 nm varied between 0.113 and 0.915 (Table
3. Results and discussion 1) and is in agreement with the finding of other authors
(Bath & Singh, 1999; Singh & Bath, 1997). It is known
3.1. Water content, water activity, sugar composition, that orange blossom honeys are honeys with very light
colour colour, which concurs with the lowest values in the ab-
sorbance range for the samples shown in Table 1. The
The results of analysis of some physicochemical pa- colour parameters L , a and b measured using the
rameters namely, moisture, water activity (aw ), and digital camera were within the range of 35.79–59.56,
colour (absorbance at 420 nm and L , a , and b colour ()5.06)–27.27 and 16.91–42.92, respectively. These val-
parameters) for the Greek honeys are summarized in ues are in close agreement with those found by others
Table 1. The refractive index varied from 1.4892 to researchers using chromatometers (Anupama et al.,
1.5043 and the corresponding moisture content ranged 2003; Popek, 2002, 2003). It is worthy to note, that or-
between 13.0% and 18.9%; these values are within the ange blossom honeys were found to have high values for
range found by other researchers and indicate a proper lightness (L ), and low values for red (a ) and yellow (b )
degree of maturity for these honey samples. In general, components, showing similar responses to the results
the moisture content in different varieties of honey may from absorbance measurements at 420 nm.
be as low as 13% (White, 1978) and as high as 29%
(Junzheng & Changying, 1998). For example, moisture 3.2. Rheological behaviour
contents have been found in the range of 14.0–16.9% for
Saudi honeys (Al-Khalifa & Al-Arify, 1999), 13.8–17.8% Fig. 1 illustrates the steady shear flow curves (Fig. 1a)
for Spanish honeys (Gomez Barez et al., 2000), 15.4– and a typical mechanical spectrum (Fig. 1b) for a Greek
18.1% for Polish honeys (Popek, 2003), 16.8–20.3% for honey sample. All honey samples behaved as Newtonian
Moroccan honeys (Terrab et al., 2002), and 18.7–21.8% fluids at all temperatures of measurement (Fig. 2).
for Indian honeys (Singh & Bath, 1997). The difference Apparent viscosity (g) and complex viscosity (g ) were
in moisture content was significant between all Greek constant, regardless of the shear rate and angular fre-
honeys; however, Greek regulations require <21% quency, respectively (Fig. 1). Moreover, the G00 was
moisture for safety from fermentation. dependent on frequency and greater than G0 at all fre-
Honey is an intermediate moisture food with a water quencies (Fig. 1b). In most of the published works,
activity of about 0.6 and is therefore shelf stable for a honey was reported to have a Newtonian behaviour
reasonable period of time. The low aw (high osmotic (Abu-Jdayil, Al-Majeed Ghzawi, Al-Malah, & Zaitoun,
environment) does not support microbial growth, pre- 2002; Bhandari, D’Arcy, & Chow, 1999; da Costa &
venting fermentation of honey by osmophilic yeast. The Pereira, 2002).
aw values of Greek honeys obtained after heating the Values of various rheological parameters obtained
samples at 50 C varied within the range 0.528–0.663 from steady shear and dynamic measurements for all
(Table 1), whereas for most of the samples the corre- samples are summarized in Table 3. These values ob-
sponding values before heating were found higher. It is tained from measurements at 20 C varied within the
well known that crystal formation in sugars results to wide range of 9.9–200.0 (Pa s) for apparent viscosity (g),
water release, thus increasing water availability. 0.15–19.10 (Pa) for storage modulus (G0 ), 64–1682 (Pa)
Table 2 shows the composition of sugars identified in for loss modulus (G00 ), and 7.7–164.4 (Pa s) for complex
the honey samples. The monosaccharides fructose (22.1– viscosity. The differences among samples could be at-
41.3%) and glucose (13.5–36.3%) were the main sugars, tributed to natural variations in composition (individual
with fructose being always the most abundant. The sugars and water content), as they belong to different
honeys with lower percentages of fructose and glucose plant species-specific varieties and collected from dif-
were those with a non-floral origin (i.e. honeydew hon- ferent geographical locations in Greece. It is clear that
Table 1
Some physicochemical parameters in 33 Greek honeys
Sample No. Botanic/geographical origin Moisture content, Water activity Colour
Xw (g/100 g) (20 C) after
Absorbance at L a b
melting at 50 C
420 nm
1 Honeydew (pine)/Thasos 18.9 0.610 0.387 45.48 12.13 40.15
2 Honeydew (pine)/Thasos 17.4 0.613 0.314 42.98 13.48 40.09
3 Honeydew (pine)/Thasos 18.3 0.615 0.335 43.40 12.75 40.49

A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21


4 Honeydew (pine)/Thasos 13.9 0.567 0.712 40.72 18.35 41.41
5 Honeydew (pine)/Thasos 15.4 0.570 0.702 42.18 15.05 40.42
6 Honeydew (pine)/Thasos 15.2 0.559 0.675 43.80 16.17 41.60
7 Honeydew (pine)/Halkidiki 15.0 0.580 0.703 45.38 9.48 38.48
8 Honeydew (pine)/Halkidiki 14.9 0.576 0.738 43.42 15.21 40.77
9 Honeydew (pine)/Halkidiki 15.4 0.576 0.770 44.32 11.91 36.76
10 Honeydew (pine)/Halkidiki 15.7 0.575 0.791 39.24 17.32 39.42
11 Honeydew (pine)/Halkidiki 15.4 0.577 0.651 43.70 20.76 42.79
12 Honeydew (pine)/Halkidiki 14.8 0.577 0.783 42.30 16.16 40.96
13 Honeydew (pine)/Evia 16.3 0.570 0.593 40.14 19.36 39.96
14 Honeydew (pine)/Evia 14.8 0.663 0.525 44.04 16.58 41.05
15 Honeydew (fir)/Helmos 13.4 0.562 0.557 41.63 19.36 41.53
16 Honeydew (fir)/Helmos 13.0 0.565 0.476 41.63 17.35 40.11
17 Honeydew (fir)/Vytina 13.9 0.570 0.428 39.79 18.49 40.59
18 Honeydew (fir)/Vytina 13.3 0.561 0.469 40.51 16.09 39.67
19 Honeydew (fir)/Vytina 14.6 0.555 0.450 41.66 16.35 39.65
20 Honeydew (fir)/Vytina 15.0 0.556 0.357 43.48 21.55 42.92
21 Honeydew (fir)/Vytina 14.1 0.581 0.405 39.87 18.75 38.95
22 Honeydew (fir)/Vytina 14.0 0.581 0.222 39.80 17.65 38.08
23 Honeydew (fir)/Vytina 15.2 0.609 0.378 41.45 16.16 39.65
24 Honeydew (fir)/Vytina 13.8 0.578 0.395 42.02 13.77 38.71
25 Floral/Livadia 13.8 0.528 0.915 42.97 27.27 39.50
26 Floral/Livadia 14.1 0.528 0.831 35.79 23.49 35.59
27 Floral/Livadia 15.1 0.550 0.682 39.30 21.30 40.84
28 Floral (Orange blossom)/Argos 15.1 0.542 0.183 47.34 1.70 22.72
29 Floral (Orange blossom)/Argos 15.8 0.540 0.161 45.75 0.89 19.85
30 Floral (Orange blossom)/Argos 16.2 0.584 0.127 59.56 )5.06 16.91
31 Floral (Orange blossom)/Argos 17.9 0.577 0.113 45.76 )2.79 18.93
32 Floral (Orange blossom)/Sparti 15.6 0.548 0.134 51.62 7.29 28.93
33 Floral (Orange blossom)/Sparti 15.6 0.546 0.115 47.18 2.96 23.31

13
14
Table 2
Distribution of the levels of various sugars (%) among the selected honey samples
Sample Fructose Glucose Sucrose Trehalose Maltose Isomal- Raffinose Erlose Melezi- Panose Isomalto-triose Malto-triose Malto-tetraose
No. (g/100 g) (g/100 g) (g/100 g) (g/100 g) (g/100 g) tose (g/100 g) (g/100 g) tose (g/100 g) (g/100 g) (g/100 g) (g/100 g)
(g/100 g) (g/100 g)
1 29.9 26.3 0.7 <0.1 4.0 1.4 0.6 3.1 0.1 0.2 0.1 0.2 7.7
2 29.0 25.6 0.8 0.1 3.9 1.4 0.6 3.3 0.1 0.2 0.1 0.2 7.9
3 30.5 26.4 0.8 <0.1 4.2 1.4 0.6 3.3 0.1 0.2 0.1 0.2 8.0

A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21


4 30.5 24.5 0.2 0.1 6.7 3.0 0.5 0.6 0.1 0.5 0.2 0.3 0.7
5 30.5 23.6 0.1 <0.1 6.1 3.2 0.2 0.6 0.1 0.5 0.2 0.3 0.7
6 30.6 23.7 0.1 <0.1 6.2 2.9 0.2 0.6 0.1 0.5 0.2 0.3 0.6
7 22.5 18.4 0.1 0.1 4.4 2.4 0.3 1.0 0.1 0.4 0.1 0.3 0.4
8 23.0 19.6 0.1 <0.1 4.4 2.4 0.6 1.1 0.1 0.4 0.1 0.3 0.4
9 22.2 19.0 0.1 <0.1 4.1 2.0 0.6 1.0 0.1 0.2 0.1 0.1 0.2
10 23.9 21.1 0.1 0.1 4.6 2.1 0.5 0.8 0.1 0.3 0.1 0.2 0.4
11 24.1 21.1 0.1 0.1 4.7 2.2 0.5 0.8 0.1 0.4 0.2 0.3 0.6
12 24.1 21.2 0.1 0.1 4.6 2.0 0.5 0.8 0.1 0.3 0.1 0.3 0.4
13 32.3 28.7 0.6 0.6 4.3 1.2 0.5 1.4 2.0 0.2 <0.1 0.2 0.3
14 29.7 25.5 1.4 0.4 4.2 1.0 0.6 2.8 1.9 0.2 <0.1 0.2 0.2
15 27.3 18.6 0.1 1.7 3.5 1.1 0.8 0.6 9.1 0.6 0.1 0.3 0.3
16 24.1 16.0 0.1 2.9 2.7 0.8 0.9 0.8 10.5 0.4 <0.1 0.3 0.3
17 25.7 16.7 0.1 2.1 3.1 1.1 0.6 0.4 11.0 1.0 0.1 0.4 0.3
18 28.1 19.0 0.2 1.6 3.6 1.0 0.5 0.5 9.8 0.7 <0.1 0.3 0.2
19 32.5 24.5 0.2 0.8 5.4 1.2 1.0 2.1 1.3 0.1 <0.1 0.5 0.5
20 31.1 25.0 0.2 0.8 5.3 0.9 1.0 2.1 1.4 0.1 <0.1 0.4 0.3
21 22.1 13.6 0.1 3.1 1.9 0.8 0.9 0.5 14.4 1.0 <0.1 0.4 0.4
22 22.2 13.5 0.1 3.0 1.9 0.7 0.9 0.5 14.2 0.9 <0.1 0.3 0.3
23 29.0 19.7 0.1 1.5 3.6 0.9 0.3 0.6 10.5 0.6 <0.1 0.2 0.2
24 22.1 14.2 0.2 2.9 2.2 1.0 0.7 0.4 13.0 1.1 0.1 0.4 0.3
25 34.8 29.2 0.2 0.6 4.7 0.9 0.4 0.8 2.5 0.2 <0.1 0.1 0.1
26 34.1 29.0 0.1 0.6 4.6 1.1 0.3 0.8 2.4 0.2 <0.1 0.2 0.2
27 34.9 30.0 0.1 0.7 4.9 1.3 0.4 0.4 2.2 0.2 <0.1 0.3 0.3
28 39.1 32.9 2.7 0.1 4.4 0.4 <0.1 0.9 0.3 <0.1 <0.01 <0.1 <0.01
29 38.1 32.8 2.7 0.1 4.3 0.4 0.5 0.8 0.3 <0.1 <0.01 <0.1 <0.01
30 39.1 36.3 0.6 0.1 3.0 0.4 0.5 0.6 0.2 <0.1 <0.01 <0.1 <0.01
31 39.9 34.7 0.6 0.1 3.3 0.5 0.5 0.6 0.2 <0.1 <0.01 0.1 <0.01
32 39.9 35.4 1.6 0.1 3.6 0.6 0.5 0.8 <0.1 0.1 <0.1 0.1 <0.01
33 41.3 32.7 1.7 <0.1 3.6 0.4 0.5 0.7 <0.1 <0.1 <0.01 <0.1 <0.01
A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21 15

(a) (b)

Fig. 1. Steady shear viscosity profiles of three Greek honey samples (a) and a representative mechanical spectrum of a Greek honey (b) at 20C.

follow the Cox–Merz rule (Cox & Merz, 1958), as ex-


pected for Newtonian fluids. According to this rule the
complex viscosity from small-deformation oscillation
measurements and the steady shear viscosity from ro-
tational measurements superimpose closely at equiva-
lent numerical values of angular frequency (x, rad s1 )
and shear rate (c, _ s1 ) for non-interacting molecular
dispersions.
The temperature effects on honey viscosity are shown
in Fig. 2; as expected, the viscosity was reduced with
increasing temperature. The temperature dependence of
g is adequately described using the Arrhenius relation-
ship (Fig. 2, inset); the correlation coefficients (r2 ) for
each sample are given in Table 3 and were all greater
than 0.91. The observation that the temperature de-
pendence of viscosity follows the Arrhenius relationship
Fig. 2. Temperature effect on viscosity of a representative Greek honey
has been also reported by other researchers for honey
and its Arrhenius plot of viscosity (at 10 s1 ) vs. temperature (inset).
(Al-Malah, Abu-Jdayil, Zaitoun, & Ghzawi, 2001; Bath
& Singh, 1999; Bhandari et al., 1999; Sopade et al.,
2002) as well as for other sugar-rich liquid foods such
the viscosity of honey decreases with water content. As as fruit juice concentrates and sugar syrups (Khalil,
shown in Fig. 3a there was a significant relationship Ramakrishna, Nanjundaswamy, & Patwardhan, 1989;
between moisture content and apparent viscosity, as Manohar, Ramakrishna, & Udayasankar, 1991). The
described by an exponential function: g ¼ 90071 activation energies (Ea ) for flow estimated from the
e0:4997Xw (r2 ¼ 0:80, p < 0:05). Such moisture content slope of the linear relationship of logðgÞ vs. ð1=T Þ are
dependence of viscosity has been also noticed by other presented in Table 3 for all samples and varied between
authors (Anupama et al., 2003; Junzheng & Changying, 69.1 and 93.75 kJ mol1 . Activation energy reflects the
1998; Sopade et al., 2002; Zaitoun, Ghzawi, Al-Malah, sensitivity of viscosity to temperature changes; higher Ea
& Abu-Jdayil, 2001). Moreover, a relationship between means that the viscosity is relatively more sensitive to
moisture content and the loss modulus, estimated by a temperature change. There was an inverse linear
dynamic rheological measurements was observed (Fig. relationship between Ea and moisture content (r2 ¼ 0:61,
3b); the experimental data fitted into an exponential p < 0:01). Although the Arrhenius formalism seems
equation: G00 ¼ 2  106 e0:5547Xw (r2 ¼ 0:72, p < 0:05). It to adequately describe the temperature dependence
is worthy here to note that the g values were very close of honey viscosity, it gives relatively high values for
with the respective g values for each individual sample the activation energy, which are more typical of
(Table 3). It appears therefore that the viscosity data chemical reactions. Similar observations concerning the
16 A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21

Table 3
Rheological parameters of 33 Greek honeys
Sample No. Viscosity (Pa s) Ea (kJ mol1 ) G0 (Pa) (x ¼ 10 rad s1 ) G00 (Pa) (x ¼ 10 rad s1 ) Complex viscosity
(20 C) (20 C) (20 C) (Pa s) (20 C)
1 9.9 72.69 (r2 ¼ 0:97) 3.01 99 9.0
2 10.5 73.62 (r2 ¼ 0:97) 0.40 80 7.7
3 10.7 74.09 (r2 ¼ 0:97) 0.33 111 10.4
4 61.1 89.09 (r2 ¼ 0:97) 4.46 665 62.2
5 55.5 88.13 (r2 ¼ 0:97) 2.88 526 51.4
6 63.6 88.74 (r2 ¼ 0:97) 4.87 708 66.2
7 42.8 82.79 (r2 ¼ 0:97) 3.02 476 43.1
8 44.4 83.79 (r2 ¼ 0:97) 2.39 467 42.2
9 46.8 84.15 (r2 ¼ 0:97) 3.91 469 42.4
10 39.0 80.88 (r2 ¼ 0:96) 2.87 424 38.4
11 37.6 80.55 (r2 ¼ 0:96) 2.22 418 37.8
12 39.8 82.29 (r2 ¼ 0:97) 2.79 429 38.8
13 38.5 79.39 (r2 ¼ 0:95) 2.10 928 34.1
14 38.4 80.36 (r2 ¼ 0:96) 6.51 265 35.9
15 86.3 86.20 (r2 ¼ 0:96) 7.83 980 88.6
16 200.0 93.75 (r2 ¼ 0:96) 10.70 1682 164.4
17 73.2 84.00 (r2 ¼ 0:96) 3.40 772 69.7
18 80.3 85.23 (r2 ¼ 0:96) 4.91 988 89.3
19 67.4 87.33 (r2 ¼ 0:97) 3.50 708 64.0
20 59.0 84.21 (r2 ¼ 0:96) 2.98 624 56.4
21 154.0 89.77 (r2 ¼ 0:96) 19.10 1177 115.0
22 160.0 89.56 (r2 ¼ 0:96) 10.20 1701 167.0
23 67.6 89.65 (r2 ¼ 0:98) 5.58 683 67.5
24 118.0 93.33 (r2 ¼ 0:98) 4.30 1074 105.0
25 79.8 90.22 (r2 ¼ 0:97) 5.64 839 75.8
26 77.9 89.74 (r2 ¼ 0:97) 4.40 803 78.9
27 73.8 88.59 (r2 ¼ 0:97) 5.90 773 69.8
28 28.8 78.87 (r2 ¼ 0:96) 3.75 376 34.0
29 26.2 79.61 (r2 ¼ 0:96) 0.94 108 20.5
30 12.6 83.06 (r2 ¼ 0:99) 0.33 106 10.4
31 13.2 75.40 (r2 ¼ 0:98) 0.15 100 9.9
32 28.6 89.75 (r2 ¼ 0:99) 0.31 208 20.4
33 26.8 90.17 (r2 ¼ 0:99) 2.35 64 26.8

magnitude of Ea derived from viscosity data of three found by other researchers; i.e. a range from )40 to )46
Jordanian honey samples (Ea 95.6–97.7 kJ mol1 , tem- C for honey samples with 15.8–18.0% moisture content
perature range 20–50 C) have been made in a previous (Sopade, Bhandari, Halley, D’Arcy, & Caffin, 2001;
study by Al-Malah et al. (2001). Sopade et al., 2002), from )37.5 to )42.5 C for samples
with m.c. about 17.5% (Cordella et al., 2002), and from
3.3. DSC thermal behaviour )42.5 to )50.7 C for samples with undetermined m.c.
(Kantor, Pitsi, & Thoen, 1999) have been reported. It is
The effect of moisture content (m.c.) on the DSC generally accepted that differences in the composition of
thermal traces of four representative native honey carbohydrate solutions could contribute to the variation
samples is demonstrated in Fig. 4a. The glass transition in the Tg ; i.e. the glass transition temperature is a func-
temperature shifted to lower temperatures with increase tion of both moisture content and the type of solute
of moisture content due to the plasticization effect of (Slade & Levine, 1991). Recently, the Tg value responses
water. Water is a very effective plasticizer for hydro- to the modification of chemical composition of honey
philic components, such as low molecular weight car- has been used, concomitantly with other thermal events
bohydrates (Levine & Slade, 1988); this effect has been detected by DSC, to develop a new method for adulte-
related to the ability of water molecules to weaken hy- ration detection in this product (Cordella et al., 2002);
drogen bonds, dipole–dipole, and intra- and inter mole- these authors have found that adulterations of honey by
cular interactions (Matveev, Grinberg, & Tolstoguzov, industrial sugar syrups can be detected calorimetrically
2000). The glass transition temperatures of the honey up to a minimum of 5–10% addition.
samples are summarized in Table 4, varying between The plasticizing action of water is also obvious at
)34.6 and )47.2 C for a moisture content range of comparative DSC traces of diluted and concentrated
13.0–18.9 g/100 g. These values are similar to those honey samples using a native sample (16), as illustrated
A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21 17

(a) (a)

(b)
(b)

Fig. 4. DSC thermal scans for four honey samples with different
moisture content (a) and for sample 16, and diluted and concentrated
sample 16 (b); arrows indicate the position of the onset temperature
considered as glass transition.
Fig. 3. Effect of moisture content on viscosity (g) (a) and loss modulus
(G00 ) (b) of Greek honeys.
sition at 0 C present in all DSC curves of Fig. 4 was not
related to crystallization of free water from the sample
in Fig. 4b. With increasing moisture content of honey but is an artifact due to moisture condensation from the
from 10.2 to 26.9 g/100 g the Tg decreased significantly atmosphere; i.e. the magnitude of this transition was
from )25.0 to )69.5 C. Similar trends in Tg for diluted found independent of the moisture content of the sam-
honeys samples have been reported by Kantor et al. ple and it represented ice melting of water in amounts
(1999). These authors have also noticed onto the DSC less than 0.22% of the amount of water present in any-
scans water crystallization from honey/water mixtures one of the honey samples examined. With reference
with less than approximately 85% honey content. Their to the concentrated sugar solutions, such as honeys, a
DSC traces showed the presence of a typical endother- rapid reduction in temperature also prevents solute nu-
mic peak in the range )20 to 0 C linked to the free cleation as a result of the lack of sufficient mobility of
water of the sample; with decreasing honey content this the sugar molecules to assemble into a crystal lattice.
peak moved towards 0 C and the enthalpy of melting The Tg –water content relationship presented in Fig. 5
increased. In the present study, the sample with the reveal more clearly the sensitivity of honey to water-
highest moisture content (26.9%) corresponds to 84% plasticization. Despite the large variation in sugar
honey content. For all the examined samples there was composition among the samples (Table 2), the experi-
no evidence of water or sugar crystallization (Fig. 4). mental data for all honeys fitted successfully well
This could be attributed to the effectiveness of quench- (r2 ¼ 0:99) to the empirical G–T model (Eq. (6)). Esti-
cooling with liquid nitrogen in preventing freezing of mated values of the G–T parameters were 3.14 for the k
water during Tg measurement and the absence of Ôfree- and 288.0 K (or 15 C) for the Tg1 , i.e. the glass transi-
water’ from the native honey samples. The minor tran- tion temperature for dry honey solids. Consequently, the
18 A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21

0.99
0.99
0.99
0.99
0.99
0.98
0.99
0.99
0.99
0.99
0.99
0.94
0.99
0.99
r2
C2 (K)
30.90
24.81
25.07
24.75
22.44
14.89
17.11
15.12
20.92
16.91
20.92
13.95
29.71
20.20
WLF (Tg from DSC experimental data)

17.68
18.98
19.38
19.89
20.41
25.18
24.56
23.17
21.60
24.17
21.52
25.01
17.20
22.68
C1
logðgTg Þ (Pa s)
13.10
14.77
15.26
15.73
16.50
21.72
21.09
20.00
17.75
20.79
17.64
21.59
12.85
18.66

Fig. 5. Relationship of glass transition and moisture content for honey;


the solid line gives the G–T plot of the experimental data from DSC
measurements.

G–T model could be an important tool for the predic-


tion of Tg . Another usefulness of the above modeling is
0.99
0.99
0.99
0.99
0.99
0.98
0.99
0.99
0.99
0.99
0.99
0.95
0.99
0.99

the determination of Tg1 , a temperature that it could not


2
r

be measured experimentally, as it is impossible to re-


logðgTg Þ (Pa s)

move the strongly bound water from honey.


Being a molecular phenomenon, the glass transition
has been often related to many food properties that
11.52
11.57
11.57
11.57
11.69
11.75
11.81
11.85
11.68
11.89
11.62
11.80
11.60
11.61

comprise mechanical/structural relaxation processes. The


Glass transition temperatures of Greek honeys and estimated parameters of the WLF model
WLF (Ôuniversal’ constants)

WLF equation (Eq. (5)) includes the glass transition


temperature as a parameter and has been suggested to be
Tg , K (C) (predicted)

applicable in predicting the temperature dependence of


214.09 ()58.91)
213.40 ()59.60)
226.73 ()46.27)

222.72 ()50.28)
219.65 ()53.35)
230.53 ()42.47)
223.17 ()49.83)

227.68 ()45.32)
227.42 ()45.58)
209.88 ()63.12)
214.92 ()58.08)
219.96 ()53.04)

viscosity, a primary property that describes mechanical


226.04()46.96)

225.65()47.39)

relaxation of mobile components. This equation has been


claimed as a more appropriated model than the Arrhe-
nius relationship in the rubbery domain; i.e. between Tg
and about Tg þ 100 C (Roos, 1995; Slade & Levine,
1991). The WLF equation specifies a much stronger
Tg , K (C) experimental

temperature dependence of viscosity compared to that


predicted by the Arrhenius formalism. In this context, the
()47.15)
()45.55)
()37.23)
()38.69)
()39.85)
()40.04)
()35.77)
()34.60)
()38.69)
()36.64)
()37.23)
()43.80)
()43.94)
()44.09)
data from DSC

Tg is considered as a reference temperature: below Tg ,


the viscosity is very high, whereas above this temperature,
225.85
227.45
235.77
234.31
233.15
232.96
237.23
238.40
234.31
236.36
235.77
229.20
229.06
228.90

the difference between storage and processing tempera-


tures (T  Tg ) is assumed to control the rate of viscosity
changes in the product. Generally, the viscosity in the
Moisture con-

glassy state varies between 107 and 1016 Pa s as has been


tent (g/100 g)

estimated and/or extrapolated for a range of materials by


various procedures (Maltini & Manzocco, 1998; Ollett &
18.9
17.4
13.9
15.4
15.4
16.3
13.0
13.9
14.6
14.1
14.1
16.2
17.9
15.6

Parker, 1990; Peleg, 1992; Slade & Levine, 1993; Soesanto


& Williams, 1981; Williams et al., 1955).
Sample No.

3.4. WLF modeling of viscosity


Table 4

The viscosity data of 14 honey samples were fitted to


13
16
17
19
21
26
30
31
33
1
2
4
5
9

the WLF model to test its applicability at temperatures


A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21 19

three honey samples. The Ôuniversal constants’ of the


WLF equation are actually average values, which have
been extracted from data on numerous glass-forming
liquids and their use is not successful for all materials.
For each data set of g vs. T (each honey sample con-
sidered individually), the WLF equation with Ôuniversal’
constants was fitted to yield the corresponding values of
Tg and gTg . Table 3 shows these predicted values for the
glass transition temperature (Tg ) ranging from )59.6 to
)42.5 C and for the glass viscosity (gTg ) values which
ranged from 1011:5 to 1011:9 Pa s. Although Tg is a pa-
rameter that is measured by various techniques, a high
viscosity value such as gTg , is not accessible experimen-
tally by any of the existing rheometers and viscometers.
The predicted Tg values differed from the corresponding
DSC values obtained experimentally by 7–19 C (Table
4). It is well known that the glass transition reflects a
Fig. 6. Temperature dependence of viscosity for 14 honey samples range of temperatures rather than a single temperature.
according to the WLF formalism; the solid line shows the WLF Even for single components (e.g. starch) the glass entails
equation using the Ôuniversal’ constants. changes over a temperature range of 10–20 C (Bili-
aderis, 1998) and thus the Tg value depends on the
technique and experimental conditions used for its de-
from 20 to 60 C (Fig. 6); i.e. between 60 and 125 C termination. Moreover, in a previous study for con-
above Tg of honeys. In general, as already mentioned, centrated fructose (69%) solutions differences up to 21
the WLF equation is not intended for use much below C between Tg s values obtained by DSC measurements
Tg , (in the glassy state) or in the very low viscosity re- and Tg s values estimated by the application of WLF on
gime (<10 Pa s) occurring at temperatures typically 100 viscosity data have been reported (Maltini & Manzocco,
C or more above Tg , where Arrhenius kinetics apply 1998).
(Slade & Levine, 1991; Soesanto & Williams, 1981; So- Recently, there have been conflicting reports on how
pade et al., 2002). In a study for the application of WLF to handle the constants in the WLF model, whether to
relationship in starch hydrolysates, D’Haene and Van use the Ôuniversal’ constants or to allow them to vary for
Liederkerke (1996) considered the latter limit restrictive a good fit to the experimentally data (Peleg, 1992; Roos,
and proved that the principles of the WLF-theory can 1995). In a second attempt to apply the WLF model on
be applied down to viscosity levels of 0.1 Pa s. These viscosity data of Greek honeys with varying the C1 and
viscosity levels are comparable to viscosity values for C2 constants and using as reference temperature the Tg
Greek honeys measured at high temperatures. Our vis- values experimentally determined from DSC, the cor-
cosity data were also extended to higher temperatures relation coefficients were the same with the previous
than usual with success. Moreover, these high temper- analysis using the Ôuniversal’ constants (Table 4). The C1
atures are usually used in honey processing for ease of and C2 constants were in the range of 17.20–25.18 and
handling with minimal quality (e.g. hydroxymethyl- 13.95–30.90 K, respectively (Table 4). Moreover, the
furfural, HMF) degradation and thus it would be useful predicted values of glass viscosities varied widely from
the prediction of viscosity changes under these condi- 1012:9 to 1021:7 Pa s and for most of the samples were
tions. The WLF equation has been previously found to extremely high which is probably unrealistic. However,
adequately describe viscosity data of Jordanian (tem- the latter findings are in agreement with a recent study
perature range 20–50 C; Al-Malah et al., 2001) and by Sopade et al. (2002) for Australian honeys. In their
Australian (temperature range 2–40 C; Sopade et al., study, although the temperature dependence of viscosity
2002) honeys. was badly predicted using the Ôuniversal’ WLF con-
Very good fits (r2 > 0:95) to the WLF model were stants, the WLF model was successfully applied if the C1
obtained for all 14 data sets (Table 4) as showed in Fig. and C2 were allowed to vary. The estimated values for
6. All data sets could be fitted to a single master curve the Australian honeys have been reported within the
using the WLF model with the fixed Ôuniversal’ con- range 13.7–21.1, 55.9–118.7 C, and 4 · 107 )4 · 1020
stants, C1 ¼ 17:44 and C2 ¼ 51:6 K (Williams et al., Pa s for C1 , C2 , and gTg , respectively (Sopade et al.,
1955). These results concur with the findings of 2002). These researchers claimed that the WLF con-
Al-Malah et al. (2001) who also found that the WLF stants should be allowed to vary in order to use them
equation, using the Ôuniversal’ constants, adequately as parameters for comparison of the temperature sen-
described the temperature-dependence of viscosity for sitivity among different samples. The WLF model with
20 A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21

different constants than the Ôuniversal’ ones has been Al-Khalifa, A. S., & Al-Arify, I. A. (1999). Physicochemical charac-
also applied by Kerr and Reid (1994) on viscosity data teristics and pollen spectrum of some Saudi honeys. Food Chem-
istry, 67, 21–25.
of sugar (glucose and sucrose) and maltodextrin solu- Al-Malah, K. I. M., Abu-Jdayil, B., Zaitoun, S., & Ghzawi, A.-M.
tions. Maltini and Manzocco (1998) have applied the (2001). Application of WLF and Arhenius kinetics to rheology of
WLF formalism in both ways (with Ôuniversal’ and with selected dark-colored honey. Journal of Food Process Engineering,
varying constants) on viscosity data for concentrated 24, 341–357.
solutions of various materials (polymers, fructose, lactic Anupama, D., Bhat, K. K., & Sapna, V. K. (2003). Sensory and
physico-chemical properties of commercial samples of honey. Food
acid, and glycerol); in their study the estimated values Research International, 36, 183–191.
for glass viscosities have been found about 1014 cP. AOAC (1990). In K. Helrich (Ed.), Official methods of analysis (15th
Soesanto and Williams (1981), using the Ôuniversal’ edn). Arlington, VA: Association of Official Analytical Chemists.
constants on viscosity data of highly concentrated (91.9– Bath, P. K., & Singh, N. (1999). A comparison between Helianthus
97.9%) fructose–sucrose mixtures (1:7), indicated that annuus and Eucalyptus lanceolatus honey. Food Chemistry, 67, 389–
397.
gTg may not be invariant; i.e. they showed that gTg varies Bhandari, B., D’Arcy, B., & Chow, S. (1999). Rheology of selected
from about 4.5 · 1014 to 1.5 · 1015 mPa s as the mole Australian honeys. Journal of Food Engineering, 41, 65–68.
fraction of sugar varies from 0.4 to 0.7. Finally, Peleg Biliaderis, C. G. (1998). Structures and phase transitions of starch
(1992) has reported that for a variety of polymers and polymers. In R. H. Walter (Ed.), Polysaccharide association
amorphous sugars, the magnitude of C1 and C2 may structures in Food (pp. 57–168). New York: Marcel Dekker.
Biliaderis, C. G., Lazaridou, A., & Arvanitoyannis, I. (1999). Glass
vary considerably from the Ôuniversal’ constants de- transition and physical properties of polyol-plasticized pullulan-
pending on the material studied, the measured property starch blends at low moisture. Carbohydrate Polymers, 40, 29–47.
and the selected reference temperature. CIE (1976). Committee TC-1.3 CIE, Technical note (1974). Journal of
Optical Society of America, 64, 896–897.
Cordella, C., Antinelli, J. F., Aurieres, C., Faucon, J. P., Cabrol-Bass,
D., & Sbirrazzuoli, N. (2002). Use of differential scanning
4. Conclusions calorimetry (DSC) as a new technique for detection of adulteration
in honeys. 1. Study of adulteration effect on honey thermal
All honey samples exhibited Newtonian flow behav- behavior. Journal of Agriculture Food Chemistry, 50, 203–208.
Cox, W. P., & Merz, E. H. (1958). Correlation of dynamic and steady
iour and the viscosity in the temperature range of 20–60 shear flow viscosities. Journal of Polymer Science, 28, 619–622.
C can be predicted using the Arrhenius relationship; the D’Haene, P., & Van Liederkerke, B. (1996). Viscosity prediction of
calculated activation energies for flow were inversely starch hydrolysates from single point measurements. Starch, 48,
related to the moisture content. Steady shear viscosity 327–334.
and the loss modulus values (G00 ) at 10 s1 or rad s1 da Costa, C. C., & Pereira, R. G. (2002). The influence of propolis on
the rheological behaviour of pure honey. Food Chemistry, 76, 417–
decreased exponentially with increasing moisture con- 421.
tent of honey. The glass transition temperature (Tg ) of Doner, L. W. (1977). The sugars of honey––a review. Journal of the
the honey samples varied between )34 and )47 C (over Science of Food and Agriculture, 28, 443–456.
the entire moisture content range, 13.0–18.9%) and ex- Doner, L. W., & Hicks, K. B. (1982). Lactose and the sugras of honey
hibited strong depression with the moisture; the Tg data and maple: reactions, properties, and analysis. In D. R. Lineback &
G. E. Inglett (Eds.), Food Carbohydrates (pp. 74–112). West Port,
fitted reasonably well to the Gordon–Taylor empirical CT: AVI Publishing Company.
equation. The WLF equation was also suitable to model Gomez Barez, J. A., Garcia-Villanova, R. J., Elvira Garcia, S., Rivas
the rheological behaviour of honey viscosity above the Pala, T., Gonzalez Paramas, A. M., & Sanchez Sanchez, J. (2000).
glass transition (T  Tg ¼ 60–125 C) using the universal Geographical discrimination of honeys through the employment of
values of the WLF model (C1 ¼ 17:44 and C2 ¼ 51:6 K) sugar patterns and common chemical quality parameters. European
Food Research and Technology, 210, 437–444.
where all samples fitted onto a common master curve. Gordon, M., & Taylor, J. S. (1952). Ideal copolymers and the second-
order transitions of synthetic rubbers. I. Non-crystalline copoly-
mers. Journal of Applied Chemistry, 2, 493–500.
Junzheng, P., & Changying, J. (1998). General rheological model for
Acknowledgement natural honeys in China. Journal of Food Engineering, 36, 165–168.
Kantor, Z., Pitsi, G., & Thoen, J. (1999). Glass transition temperature
The authors wish to thank Mr. A. Pavlis for his of honey as a function of water content as determined by
analytical assistance. differential scanning calorimetry. Journal of Agriculture Food
Chemistry, 47, 2327–2330.
Kerr, W. L., & Reid, D. S. (1994). Temperature dependence of the
viscosity of sugar and maltodextrin solutions in coexistence with
References ice. Lebensmittel-Wissenschaft und-Technologie, 27, 225–231.
Khalil, K. E., Ramakrishna, P., Nanjundaswamy, A. M., & Patward-
Abu-Jdayil, B., Al-Majeed Ghzawi, A., Al-Malah, K. I. M., & han, M. V. (1989). Rheology of banana juice. Journal of Food
Zaitoun, S. (2002). Heat effect on rheology of light- and dark- Engineering, 10, 231–240.
colored honey. Journal of Food Engineering, 51, 33–38. Levine, H., & Slade, L. (1988). Principles of ‘‘cryostabilization’’
Adobe Photoshop 5.0 User Guide for Machintosh and Windows. technology from structure/property relationship of carbohydrate/
(1998). Adobe Systems, Inc., San Jose, CA, USA. water systems––a review. Cryo-Letter, 9, 21–63.
A. Lazaridou et al. / Journal of Food Engineering 64 (2004) 9–21 21

Maltini, E., & Manzocco, L. (1998). Experimental and predicted Slade, L., & Levine, H. (1991). Beyond water activities: recent
viscosities of model solutions at temperatures above the glass advances based on an alternative approach to the assessment of
transitions. In proceedings of ISOPOW 7 International Symposium food quality and safety. Critical Reviews in Food Science and
on water management in the design and distribution of quality Nutrition, 30, 115–360.
foods, Helsinki, Filand, pp. 57–60. Slade, L., & Levine, H. (1993). The glassy state phenomenon in food
Manohar, B., Ramakrishna, P., & Udayasankar, K. (1991). Some molecules. In J. M. V. Blanshard & P. J. Lillford (Eds.), The glassy
physical properties of Tamarind Tamarindus indica L. juice state in foods (pp. 35–101). Sutton Bonington, UK: Nottingham
concentrates. Journal of Food Engineering, 13, 241–258. University Press.
Matveev, Y. I., Grinberg, V. Y., & Tolstoguzov, V. B. (2000). The Soesanto, T., & Williams, M. C. (1981). Volumetric interpretation of
plasticizing effect of water on proteins, polysaccharides and their viscosity for concentrated and dilute sugar solutions. Journal of
mixtures. Glassy state of biopolymers, food and seeds. Food Physical Chemistry, 85, 3338–3341.
Hydrocolloids, 14, 425–437. Sopade, P. A., Bhandari, B., Halley, P., D’Arcy, B., & Caffin, N.
Ollett, A. L., & Parker, R. (1990). The viscosity of supercooled (2001). Glass transition in Australian honeys. Food Australia, 53,
fructose and its glass transition temperature. Journal of Texture 399–404.
Studies, 21, 355–362. Sopade, P. A., Halley, P., Bhandari, B., D’Arcy, B., Doebler, C., &
Papadakis, S. E., Abdul-Malek, S., Kamdem, R. E., & Yam, K. L. Caffin, N. (2002). Application of the Williams–Landel–Ferry model
(2000). A versatile and inexpensive technique for measuring color to the viscosity–temperature relationship of Australian honeys.
of foods. Food Technology, 54(12), 48–51. Journal of Food Engineering, 56, 67–75.
Peleg, M. (1992). On the use of the WLF model in polymers and foods. Sugisaki, M., Suga, H., & Seki, S. (1968). Calorimetric study of the
Critical Reviews in Food Science and Nutrition, 32, 59–66. glassy state IV. Heat capacities of glassy water and cubic ice.
Perez-Arquille, C., Conchello, P., Arino, A., Juan, T., & Herresa, A. Bulletin of the Chemical Society of Japan, 41, 2591–2599.
(1994). Quality evaluation of spanish rosemary (Rosmarius offici- Terrab, A., Diez, M. J., & Heredia, F. J. (2002). Characterization of
nalis) honey. Food Chemistry, 51, 207–210. Moroccan unifloral honeys by their physicochemical characteris-
Popek, S. (2002). A procedure to identify a honey type. Food tics. Food Chemistry, 79, 373–379.
Chemistry, 79, 401–406. White, J. W., Jr. (1978). Honey. Advances in Food Research, 24, 288–
Popek, S. (2003). Identification of honey types. Nahrung/Food, 47, 39–40. 354.
Roos, Y. H. (1995). Phase transitions in foods. New York: Academic Williams, M. L., Landel, R. F., & Ferry, J. D. (1955). The temperature
Press. dependence of relaxation mechanisms in amorphous polymers and
Sabatini, A. G., Marcazzan, G. L., Colombo, R., Carpana, E., & other glass-forming liquids. Journal of the American Chemical
Serra, G. (2001). The analytical determination of sugars in honey. Society, 77, 3701–3707.
Industrie Alimentari, 40(404), 623–627. Zaitoun, S., Ghzawi, A. -M., Al-Malah, K. I. M., & Abu-Jdayil, B.
Singh, N., & Bath, P. K. (1997). Quality evaluation of different types of (2001). Rheological properties of selected light colored Jordanian
Indian honey. Food Chemistry, 58, 129–133. honey. International Journal of Food Properties, 4, 139–148.

You might also like