Sei sulla pagina 1di 9

Food Chemistry 214 (2017) 717–725

Contents lists available at ScienceDirect

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

Chemical profiling of the major components in natural waxes


to elucidate their role in liquid oil structuring
Chi Diem Doan a,d,⇑, Chak Ming To a, Mike De Vrieze b, Frederic Lynen b, Sabine Danthine c, Allison Brown a,
Koen Dewettinck a, Ashok R. Patel a,⇑
a
Laboratory of Food Technology and Engineering, Faculty of Bioscience Engineering, Ghent University, 653 Coupure Links, 9000 Ghent, Belgium
b
Separation Science Group, Department of Organic and Macromolecular Chemistry, Faculty of Science, Ghent University, 281 Krijgslaan, 9000 Ghent, Belgium
c
Department of Food Technology, Gembloux Agricultural University, Passage des Déportés 2, B-5030 Gembloux, Belgium
d
Department of Food Technology, College of Agriculture and Applied Science, Cantho University, Viet Nam

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

Article history: Elucidating the composition of waxes is of utmost importance to explain their behavior in liquid oil
Received 28 October 2015 structuring. The chemical components (hydrocarbons – HCs, free fatty acids – FFAs, free fatty alcohols
Received in revised form 19 June 2016 – FALs and wax esters – WEs) of natural waxes were analyzed using HPLC–ELSD and GC–MS followed
Accepted 20 July 2016
by evaluation of their oil structuring properties. The gel strength, including the average storage modulus
Available online 21 July 2016
and oscillation yield stress, displayed a negative correlation with FALs and a positive correlation with
HCs, FFAs and WEs. The components dictating the gel strength are HCs, FFAs and WEs in a descending
Chemical compounds studied in this article:
order of importance. The consistency of the oleogels increased with the increasing amount of FFAs and
Heptacosane (PubChem CID: 11636)
Nonacosane (PubChem CID: 12409)
HCs and the decreasing amount of WEs and FALs. The presence of more WEs results in a strong but brittle
Hentriacontane (PubChem CID: 12410) gel with a high initial flow yield stress. We believe these results might be useful in selecting the right
Palmitic acid (PubChem CID: 985) waxes to combine in certain fat-based food products.
Stearic acid (PubChem CID: 5281) Ó 2016 Elsevier Ltd. All rights reserved.
Arachidic acid (PubChem CID: 10467)
Octacosanoic acid (PubChem CID: 10470)
Myricic acid (PubChem CID: 5461027)
1-Triacontanol (PubChem CID: 68972)
1-Dotriacontanol (PubChem CID: 96117)

Keywords:
Wax
Chemical components
HPLC
GC–MS
X-ray
Gelation

1. Introduction gelling behavior at varying critical gelling concentrations of waxes,


depending on the type of wax and solvent, the wax-wax and wax-
In recent years, there has been an emerging interest in solvent interaction, the cooling and shear rate, the temperature
employing waxes as effective structuring agents to entrap a high and time of cooling (Blake & Marangoni, 2015b; Dassanayake,
amount of liquid oils (>90%wt) into solid-like systems, called Kodali, Ueno, & Sato, 2012). The macroscopic properties of an
oleogels, which are self-standing, thermoreversible and viscoelas- oleogel are also influenced by the nature of interactions between
tic (Marangoni & Garti, 2011). Wax-based oleogels exhibit different the microstructural components occurring at the junction zones
(permanent or transient), which exist between intersecting crystal
branches or entangled crystal fibers (Carriere & Inglett, 2000; Lee &
⇑ Corresponding author at: Laboratory of Food Technology and Engineering, Inglett, 2007). The amount of such interactions can be increased or
Faculty of Bioscience Engineering, Ghent University, 653 Coupure Links, 9000 decreased by altering the number of aggregated crystal clusters,
Ghent, Belgium.
resulting in a change of gel properties (Carriere & Inglett, 2000).
E-mail addresses: Chi.DoanDiem@UGent.be, ddchi@ctu.edu.vn (C.D. Doan),
Patel.Ashok@UGent.be (A.R. Patel).
In addition, the gelling behavior is also governed by the wax crystal

http://dx.doi.org/10.1016/j.foodchem.2016.07.123
0308-8146/Ó 2016 Elsevier Ltd. All rights reserved.
718 C.D. Doan et al. / Food Chemistry 214 (2017) 717–725

morphologies (the rod-like or platelets crystals will result in the (C44) and lignoceryl lignocerate (C48) were purchased from
formation of a strong gel), which are specified by the nature and Sigma-Aldrich Chemie GmbH, Schnelldorf, Germany. HPLC-grade
chain length of the inner chemical components present in waxes solvents used for analysis were hexane (EMD Milipore Corpora-
(Hwang, Kim, Singh, Winkler-Moser, & Liu, 2012). Therefore, a tion), methyl-tertbutylether (MTBE – Sigma-Aldrich), chloroform
better understanding of the chemical composition of waxes is of (Sigma-Aldrich) and acetic acid (Sigma-Aldrich). The standards
utmost importance to explain their behavior in liquid oil network were prepared in chloroform/hexane (1:1) solvent at concentra-
development. tions ranging between 1 and 2000 ppm, which spanned the same
Natural waxes are lipid in nature and from a chemical concentrations of wax constituents in HPLC analysis.
standpoint, they are comprised of complicated mixtures, such as:
n-alkanes or hydrocarbons (HCs), free fatty acids (FFAs), free
2.1.2. Preparation of wax samples
fatty alcohols (FALs), wax esters (WEs), ketones, and sterols
Sunflower wax (SW – melting temperature Tm = 71.7 °C) and
(Kolattukudy, 1976; Regert, 2009). However, information regard-
bees wax (BW, Tm = 49.9 °C) were provided by Koster Keunen
ing the chain length, the ratio, and percentage of components in
Holland BV Co (Bladel, The Netherlands). Oryza Sativa rice bran
each chemical class of waxes is still lacking. Furthermore, the
wax (RBW, Tm = 70.2 °C), candelilla wax (CLW, Tm = 56.8 °C),
chemical composition in each wax can vary depending on its origin
carnauba Brazilian wax (CRBW, Tm = 60.9 °C), carnauba wild wax
and extraction methods. Thus, it is important to have a complete
(CRWW, Tm = 60.7 °C) and berry wax (BEW, Tm = 16.3 °C) were
analysis related to the percentage of each chemical class
provided by Kalhwax GmbH & Co. KG (Trittau, Germany). These
(HC, FFA, FAL and WE) in each wax as well as the chain length of
waxes were dissolved in chloroform/hexane 1:1 solvent at a
different components in each wax class.
concentration of 2000 ppm for HPLC analysis.
Various techniques are available for the chemical characteriza-
tion of waxes. The different wax classes can first be separated using
thin layer chromatography (TLC) (Kanya, Rao, & Sastry, 2007) 2.1.3. Analysis of wax composition using HPLC–ELSD
before being subjected to next steps, such as Fourier Transform The analytical protocol from Hwang et al., was modified to fit on
infrared spectroscopy (Odlyha, 1995), supercritical gas chromatog- the Alliance HPLC system (Hwang et al., 2002). The three solvents
raphy (Hamilton, 1995), liquid chromatography (Asperger, used in the mobile phase were solvent A (0.2% acetic acid in
Engewald, & Fabian, 2001), gas chromatography–flame ionization hexane), solvent B (0.2% acetic acid in MTBE) and solvent C
detection (Marinach, Papillon, & Pepe, 2004), pyrolysis gas chro- (isopropanol). Two HPLC pumps were operated in gradient modes
matography (Regert, Langlois, & Colinart, 2005). TLC is a fast and by a Waters 2690 separation module combined with a gradient
widely used technique for separating complex mixtures; however, controller. A guard column (4  3.0 mm ID silica cartridge in a
the quality of achievable separation through TLC represents only Security Guard cartridge system; Phenomenex) was connected to
one side, and the unambiguous assignment of the obtained spots the main column Agilent Zorbax 5 lm Rx-Sil (4.6 mm i.d. 
to defined compounds (Fuchs, Süß, Nimptsch, & Schiller, 2009). 250 mm length) and both were heated at 40 °C. The Alltech 3300
High pressure liquid chromatography (HPLC) equipped with an ELSD detector was operated at 60 °C with a nitrogen (N2) flow of
UV–vis detector and evaporative light scattering detector (ELSD) 1.7 mL/min. The temperature during HPLC–ELSD operation should
detector has also been proven to be an effective tool due to the be controlled around 40–75 °C (depending on the melting
high sensitivity regarding the achievable chromatographic resolu- temperature of the measured wax, but lower than evaporation
tion and peak tailing (Fuchs et al., 2009; Moreau, Kohout, & Singh, temperature of diluted solvents) to avoid the solidification of
2002). For substances lacking useful ultraviolet chromophore like waxes inside HPLC lines and columns, and to enhance the column
waxes, ELSD is a robust choice because it can detect non-volatile polarity (Moreau et al., 2002). By injecting 10 lL of the standard
substances that block the light path (Hwang, Cuppett, Weller, & stock at different concentrations, a calibration curve was plotted
Hanna, 2002; Megoulas & Koupparis, 2005). Therefore, the objec- to link the relation between the concentration and area for further
tive of this study is to characterize the composition of natural analysis. The elution program and washing step are respectively
waxes using HPLC–ELSD/GC–MS and, to understand the role of described in Tables 1A and S1.
different chemical constituents in oil structuring. To make a com-
prehensive case, seven natural waxes (rice bran wax, sunflower
wax, bees wax, candelilla wax, carnauba Brazilian wax, carnauba Table 1A
wild wax and berry wax) were used at their minimum gelling Gradient program on the Alliance HPLC system.
concentrations to transform liquid rice bran oil (RBO) into oleogels.
Gradient
These wax-based oleogels were comparatively evaluated by
Time (min) Solvent
exploring oscillation and flow measurements, solubility, differen-
tial scanning calorimetry, powder X-ray diffraction and polarized A (%) B (%) C (%)
light microscopy to gain a deeper insight into their crystallization 0 100 0 0
and gelation properties. 7 100 0 0
9 95 5 0
14 95 5 0
16 75 25 0
2. Materials and methods 20 75 25 0
24 40 60 0
2.1. Analytical HPLC–ELSD 28 40 60 0
30 0 40 60
35 0 50 50
2.1.1. Preparation of the standard solutions 50 40 60 0
Standard stocks including hexadecane (C16), docosane (C22), 55 75 25 0
pentacontane (C50), palmitic acid (C16), heptadecanoic acid (C17), 65 75 25 0
oleic acid (C18:1), stearic acid (C18), eicosanoic acid (C20), docosa- Solvent A: 0.2% acetic acid in hexane.
noic acid (C22), 1-triacontanol (C30), lignoceryl alcohol or Solvent B: 0.2% acetic acid in MTBE.
1-tetracosanol (C24), lignoceric acid (C24), behenyl arachidate Solvent C: isopropanol (IPA).
C.D. Doan et al. / Food Chemistry 214 (2017) 717–725 719

2.2. Analytical GC–MS in Ghent (Belgium). Fatty acid composition in RBO were analyzed
by an interscience thermofocus GC (Shimadzu Co., Japan) equipped
2.2.1. Fraction collector with a TX-2330 column and a flame ionization detector, including
A Waters fraction collector II was attached to the HPLC for the oleic acid (44.2%w/w), linoleic acid (30.0%w/w), palmitic acid
preparative collection of different wax classes, without connecting (19.9%w/w), stearic acid (2.1%w/w) and arachidic acid (1.1%w/w).
to the ELSD detector. The collector was programmed to collect each To prepare the oleogels, natural waxes were dispersed in RBO at
fraction at an interval of 4 min with an initial holding time of 2 min their critical gelling concentration (RBW: 5.0%wt, SW: 1.0%wt, BW:
(fraction 1 of HCs and WEs at minute 2–6, fraction 2 of FFAs at min- 1.5%wt, CLW: 1.0%wt, CRBW: 4.0%wt, CRWW: 2.0%wt and BEW:
ute 10–14, fraction 3 of FALs at minute 18–22). All fractions were 1.0%wt) (Doan, Van de Walle, Dewettinck, & Patel, 2015) (Details
collected from a 50 lL injection of a 5000 ppm wax sample at a of determining the minimum gelling concentration is described
flow rate of 1 mL/min after the separation with HPLC. Each fraction in Supporting information). The wax-oil mixtures were heated at
contained 4 mL and was derived for further analysis in GC–MS. 90 °C under mild agitation (200 rpm) using a magnetic stirrer
(Model EM300T, Labotech Inc., Germany) for 10–30 min until clear
2.2.2. Derivatization solutions were obtained. The clear oily solutions were subse-
FFAs were derived according to the method of Bicalho et al., quently transferred into plastic cups before being cooled down to
using a 5%v/v acetyl chloride/methanol (AcCl/MeOH) reagent 5 °C at a cooling rate of 2 °C/min and stored at 5 °C overnight in
(Bicalho, David, Rumplel, Kindt, & Sandra, 2008). One mL of reagent a thermal cabinet for further experiments.
was added to the sample; the reaction proceeded for 30 min at
90 °C and was terminated by adding 1 mL of bi-distilled water. 2.3.2. Thermal behavior
The fatty acid methyl esters (FAMEs) were extracted twice with Onset crystallization (Tc, onset), onset melting (Tm, onset), peak
1 mL of hexane before the injection in the GC–MS. melting (Tm) and peak crystallization (Tc) temperatures of neat
AcCl reagent (0.5 mL) was added to FAL fractions and the waxes and wax-based oleogels were determined in triplicate using
reaction proceeded for 30 min before AcCl was evaporated with a a Q1000 DSC and analyzed using the software TA Universal Analy-
nitrogen flow. The sample was re-dissolved with 1 mL of hexane sis (TA Instruments, New Castle, Delaware, USA). The cell constant
in a 1.5 mL crimp neck vial for the injection in GC–MS. and temperature were set with indium (TA Instruments). A tem-
Wax esters were saponified according to a procedure of perature calibration was done using azobenzene (Sigma-Aldrich,
Colombini et al., with a slight modification (Lluveras, Bonaduce, Bornem, Belgium) and undercane (Acros Organics, Geel, Belgium).
Andreotti, & Colombini, 2009). The sample was added with 1 mL The sample (neat waxes and wax-based oleogels) was first placed
of 10%wt KOH/EtOHaq for saponification, assisted by microwaves inside an aluminum pan, being sealed with an aluminum lid (TA
at 90 °C for 2 h (power  40 W) in a microwave oven equipped Instruments). The sample weight can vary from 6.0 to 8.0 mg.
with a robotic arm for sampling (Biotage Microwave Synthesizer, The sealed aluminum pans were first located at the sample holder
Model Initiator 8). and followed the thermal program: heating at 90 °C for 10 min to
erase all the crystalline history before being cooled to 5 °C at a
cooling rate of 2 °C/min, keeping isothermally at 5 °C for 20 min
2.2.3. Analytical GC–MS and reheating to 90 °C at a heating rate of 5 °C/min.
The major components in the derivatized samples were deter-
mined using an Agilent 7890A GC System, equipped with an 2.3.3. Rheological behavior
Agilent Technologies 5975C inert MSD with a triple-acid detector All the rheological measurements were carried out using an
and an automatic injector (Agilent technologies 7683B Series advanced rheometer AR2000ex (TA Instruments, New Castle,
injector). The column was a fused capillary column HP-5MS USA). The average storage modulus (G0 LVR) in the linear viscoelastic
(5%wt diphenyl/95%wt dimethyl-polysiloxane, 30 m  0.25 mm, region (LVR) of wax-based oleogels were measured at 5 °C using a
film thickness 0.25 lm, Agilent Technologies, Palo Alto, CA) with parallel plate geometry (cross-hatched plate-plate geometry;
the GC injector setting at 250 °C. Samples of 1 lL were injected diameter, / = 40.0 mm; gap = 1000 lm) by logarithmically increas-
using a 10 lL injector under splitless mode. The temperature pro- ing the oscillation stress from 0.01 to 100 Pa at a frequency of
gram (Table 1B) was applied to all fractions. Auxiliary temperature 1.0 Hz. Steady state flow measurements were performed by
was set at 300 °C. The mass spectrometer was operated in the EI logarithmically increasing the shear rate from 0.1 to 20 s1. The
positive mode (70 eV). The temperature of the MS ion source was obtained flow curves were fitted to the general Herschel-Bulkley
kept at 230 °C and the temperature of the MS quadrupole was (Eq. (1)) model.
150 °C. The detected low and high mass were 40 and 550 m/z,
respectively. r ¼ ro þ K c_ n ð1Þ
where
2.3. Preparation, gelation and thermal behavior of wax-based oleogels
r: shear stress (Pa)
2.3.1. Sample preparation ro : yield stress (Pa)
Rice bran oil (RBO) (onset and peak melting temperatures, K: consistency index (Pa.sn)
Tm, onset = 73.95 °C, Tm = 11.14 °C) (Suriny brand, Surin Bran c_ : shear rate (s1)
Oil Company, Thailand) was purchased from a local supermarket n: flow index

2.3.4. Powder X-ray diffraction spectroscopy (XRD)


Table 1B
Polymorphism of the wax crystals in oleogels was investigated
GC–MS temperature program.
by XRD using a Bruker D8-Advanced Diffractometer (Bruker,
Rate (°C/min) Value (°C) Hold time (min) Run time (min) Germany) (k Cu = 1.54178 Å, 40 kV, and 30 mA), equipped with
0 50 1 1 an Anton Paar temperature control system composed of a TTK
Rate 1 11 175 0 12.364 450 low-temperature chamber connected to a waterbath (Lauda)
Rate 2 10 240 0 18.864
and heating device (TCU 110 Temperature Control Unit) (Anton
Rate 3 10 320 15 41.864
Paar, Graz, Austria). The samples were heated at 90 °C for 10 min
720 C.D. Doan et al. / Food Chemistry 214 (2017) 717–725

and subsequently cooled to 5 °C at a cooling rate of 5 °C/min. The wax from oil refineries contained 66% of WEs, 10% of FALs and 16%
long and short-spacing runs were recorded at 5 °C using a Lynxeye of FFAs (Kanya et al., 2007). The difference can be explained by the
detector (Bruker, Germany). difference in wax origin and analysis methods. SW used by Sindhu
et al., was crude wax and purified in the laboratory, while SW in
2.4. Statistical analyses the present study was a pure product received from Kalhwax
Company. The difference in sensitiveness between TLC and
The experimental data are expressed as means ± standard devi- HPLC–ELSD are also reasonable for the variation in separation
ation of three repetitions. The statistical tests were performed results. CRBW and CRWW, as expected, had a similar chemical
using SPSS version 22. One-way Anova test, two-sample T-test, composition, with a major WE fraction and a medium amount of
Kruskall-Wallis test and Mann-Whitney U test were applied FALs. BW and CLW showed a multi-component profile with respec-
depending on the compliance of data to a normal distribution. tive WEs and n-alkanes being prominent. BW also contained a
Dunnet T’3 post-hoc test was used to find the significant difference considerable amount of HCs and FFAs while CLW contained
in the Anova test when no homogeneity of variances is found. additional WEs and FFAs.

3. Results and discussion 3.2. Identification of the chain length of chemical components in
natural waxes using GC–MS
3.1. Characterization of chemical classes in natural waxes using
HPLC–ELSD 3.2.1. Analysis of fatty hydrocarbons, free fatty acids and free fatty
alcohols
HPLC–ELSD has proven its effectiveness as a tool for the separa- After the separation through the fraction collector, the hydro-
tion of polar and non-polar lipidic classes (Moreau et al., 2002; carbon fractions of BW and CLW (resulting from HPLC–ELSD) were
Nordbäck & Lundberg, 1999). The use of hexane and MTBE subjected to GC–MS without derivatization. The data in Table 3
combined with 0.2% acetic acid provided a better elution through show that the major constituents in the hydrocarbon class of BW
the silica column for lipidic compounds which differ in polarity and CLW are odd-numbered long chains, among which C31
(Asikin et al., 2012). contributes the most to CLW composition (accounting for more
A mixture of 10 standards (docosane, heptadecanoic acid, oleic than 80%) while BW contains a considerable amount of C27, C29
acid, stearic acid, eicosanoic acid, docosanoic acid, triacontanol, and C31. These results are comparable to those previously reported
tetracosanol, behenyl arachidate and lignoceryl lignocerate) was (Toro-Vazquez et al., 2007; Tulloch, 1973).
injected into HPLC–ELSD system to determine the retention time The long-chain FFAs and FALs (>20 carbons) were not volatile
of each lipidic class. Because the sterol compound is acknowledged enough and the presence of the free polar head group hampered
as a minor component in wax, it can be omitted in this study the movement through the column. By first deriving the FFAs
(Asikin et al., 2012). The application of the modified elution and FALs into their non-polar derivatives, the resultant peak shape
program (Table 1A) led to distinctive retention times between and resolution in GC–MS chromatogram were improved (Asikin
the four chemical classes in order of increasing polarity. Fatty et al., 2012). The fractions of FFAs and FALs were derivatized by
hydrocarbons (HCs) belong to the same chemical class as the elu- applying the acid-catalyzed esterification and acetylation, convert-
tion solvent (0.2% acetic acid in hexane) and eluted first, followed ing them into fatty acid methyl esters (FAMEs) and alcohol
by WEs, FFAs and FALs (Fig. S1). For the standard mixture with high acetates, respectively. The use of GC–MS allows us to refine the
purity, each compound could be distinguished separately in its chromatograms by extracting the ions with a certain m/z value
corresponding class, except for the two wax esters which eluted (m/z = mass number/charge number of ions). The mass value of
together (Fig. S1H). For the separation of neat waxes, each peak 74 m/z is highly characteristic in the analysis of FAMEs because
represented one single chemical class because all the constituents of the McLafferty rearrangement. The chain length and degree of
in one class eluted together due to their similar polarity (Figs. S1A, packing determined the order of elution for components within a
S1B, S1C, S1D, and S1G). Carnauba wax was an exception since chemical class, which was clearly demonstrated by the broad range
peak 4 of the FALs was clearly separated (Figs. S1E and S1F). This and discriminatory capacity of the GC–MS. The smaller chain
may have happened because the individual alcohols in carnauba length and lower degree of packing enabled the molecules to enter
wax significantly differ in polarity. the vapor phase, which reduced the retention time. The com-
After the separation of four lipidic classes in HPLC–ELSD, an pounds were identified by comparing the retention time of the
integration was performed to compare the percentage of these derivatized samples with the standard, and using the NIST library
chemical compounds in the waxes (Table 2). RBW, SW and BEW database, which further confirmed the accuracy of the method.
are considered as mono-component materials with more than Short-chain components (<14 carbons), unsaturated compo-
90% of the prominent constituent: WEs in RBW and SW; and FA nents and components with an odd-alkyl chain were present in
in BEW. SW contained 96.23% of WEs, 3.29% of FFAs, 0.32% of FALs very low amounts, indicating that the FFA portion of natural waxes
and 0.17% of HCs, which is different from TLC/GC–MS results consisted largely of components with an even alkyl chain between
obtained by Sindhu Kanya et al., who reported that sunflower seed 16 and 32 carbon atoms. The main FFA components in each wax

Table 2
The chemical composition of the waxes analyzed through HPLC–ELSD on the Alliance system.

RBW SW BW CLW CRBW CRWW BEW


* a a b c a a
HC 0.29 ± 0.29 0.17 ± 0.15 26.84 ± 1.04 72.92 ± 2.23 0.41 ± 0.30 0.16 ± 0.09 0.03 ± 0.01a
WE* 93.49 ± 2.63E 96.23 ± 0.19E 58.00 ± 0.68C 15.76 ± 0.35B 62.05 ± 3.03D 58.99 ± 4.57C 0.02 ± 0.02A
FFA* 6.00 ± 2.12a,b 3.29 ± 0.16a 8.75 ± 0.75b,c 9.45 ± 1.14c 6.80 ± 0.76a,b 6.32 ± 3.65a,b 95.70 ± 1.11e
FAL* 0.22 ± 0.22A 0.32 ± 0.38A 6.42 ± 0.90B 2.20 ± 1.02A,B 30.74 ± 2.48C 34.53 ± 3.45C 4.24 ± 1.10A,B

HC: hydrocarbon, WE: wax ester, FFA: free fatty acid and FAL: free fatty alcohol.
RBW: rice bran wax, SW: sunflower wax, BW: bees wax, CLW: candelilla wax, CRBW: carnauba Brazilian wax, CRWW: carnauba wild wax, and BEW: berry wax.
*
a, b, c, d, e, A, B, C, D, E: Different letters in the same format style, within the same row, indicate significant differences at p < 0.05 according to Dunnet T’3 post-hoc test (SPSS 22).
C.D. Doan et al. / Food Chemistry 214 (2017) 717–725 721

Table 3
Analytical results of fatty hydrocarbons, free fatty acids and free fatty alcohols, fatty acid moieties and fatty alcohol moieties of wax esters from GC–MS.

RBW SW BW CLW CRBW CRWW BEW


Hydrocarbon
C27 // // 40.28 ± 3.43 0.15 ± 0.13 // // //
C29 // // 25.63 ± 1.90 6.26 ± 2.57 // // //
C31 // // 18.05 ± 0.85 82.48 ± 1.95 // // //
C33 // // 3.12 ± 1.19 7.68 ± 0.85 // // //
FFA
C16 20.76 ± 2.55b 33.18 ± 2.90d 30.11 ± 5.57cd 17.84 ± 1.78a 16.17 ± 1.85a 28.57 ± 6.04c 82.13 ± 2.49e
C18 5.54 ± 0.61b 20.42 ± 1.83e 6.84 ± 1.54b 1.90 ± 0.29a 17.78 ± 1.79d 5.38 ± 1.78b 11.26 ± 1.67c
C20 8.70 ± 1.58b 22.12 ± 2.05d 1.74 ± 0.64a 1.31 ± 0.29a 10.57 ± 2.14c 9.14 ± 0.65bc 0.39 ± 0.19a
C22 17.52 ± 3.12e 9.83 ± 2.26d 4.33 ± 1.39b 1.60 ± 0.17a 6.28 ± 2.40c 5.50 ± 0.35bc //
C24 28.33 ± 3.91d 4.00 ± 0.95b 27.80 ± 2.84d 0.64 ± 0.19a 15.47 ± 3.49c 13.37 ± 3.79c //
C26 3.36 ± 0.76b 2.26 ± 0.40b 6.55 ± 0.76c 0.90 ± 0.29a 6.17 ± 1.15c 5.25 ± 2.30c //
C28 3.34 ± 0.71a 2.85 ± 0.09a 7.20 ± 0.30b 3.13 ± 0.24a 10.58 ± 0.18c 9.78 ± 2.24c //
C30 3.03 ± 0.80b 0.99 ± 0.73a 6.60 ± 0.31d 31.38 ± 0.82e 4.70 ± 2.39c 4.50 ± 1.17c //
FAL
C24 25.30 ± 2.08c 25.25 ± 1.13c 12.01 ± 2.11b 1.77 ± 0.12a // // //
C26 26.15 ± 0.36d 24.11 ± 2.91c 9.46 ± 2.60b 4.53 ± 1.25a // // //
C28 16.16 ± 1.41c 17.74 ± 1.68c 12.95 ± 1.35b 11.48 ± 1.45b 6.38 ± 1.64a 5.67 ± 0.37a //
C30 15.39 ± 1.36b 10.46 ± 3.00a 29.23 ± 0.20c 42.28 ± 2.30d 13.94 ± 1.22b 13.22 ± 0.38b //
C32 9.47 ± 2.50b 5.92 ± 0.63a 20.40 ± 2.74d 14.57 ± 1.97c 65.39 ± 2.71e 68.29 ± 0.82e //
C34 4.70 ± 0.15b // 3.94 ± 0.28ab 2.83 ± 0.20a 11.17 ± 1.05c 10.95 ± 1.20c //
Wax ester fraction
FA moieties RBW SW BW CLW CRBW CRWW BEW
b a f e c d
C16 8.63 ± 1.47 4.89 ± 1.85 80.27 ± 4.50 34.16 ± 3.91 15.16 ± 5.19 19.35 ± 0.46 79.89 ± 4.55f
C18 4.97 ± 0.69a 4.84 ± 1.49a 9.54 ± 1.57b 17.44 ± 1.01d 16.69 ± 3.91d 19.00 ± 9.47e 13.25 ± 2.38c
C20 22.95 ± 1.25e 48.65 ± 6.87f 0.51 ± 0.15a 6.31 ± 3.23b 16.02 ± 0.11d 13.06 ± 2.67c //
C22 26.75 ± 2.19d 24.14 ± 1.55d 0.43 ± 0.17a 10.87 ± 2.56bc 12.19 ± 0.77c 9.47 ± 1.48b //
C24 27.83 ± 1.17e 5.87 ± 1.27b 0.47 ± 0.19a 1.20 ± 0.70a 22.38 ± 5.35d 18.65 ± 4.11c //

FAL moieties
C18 2.80 ± 0.93a 1.82 ± 0.60a 2.93 ± 0.85a 20.70 ± 7.09c 13.69 ± 3.10b 21.48 ± 1.13c 66.02 ± 2.42d
C22 6.92 ± 1.04 9.48 ± 1.43 // // 3.08 ± 0.09 2.55 ± 0.51 //
C24 29.46 ± 4.24e 34.41 ± 4.06f 25.19 ± 3.50d 2.05 ± 0.08a 3.61 ± 1.66ab 4.41 ± 0.92b 8.77 ± 2.09c
C26 23.94 ± 1.10e 30.07 ± 1.10f 14.20 ± 0.94d 8.18 ± 1.16c 1.73 ± 0.87a 1.46 ± 0.64a 3.82 ± 0.29b
C28 13.70 ± 1.31d 12.15 ± 1.98c 13.45 ± 0.51cd 18.72 ± 0.79e 3.15 ± 1.98b 1.60 ± 0.52a 3.32 ± 0.07b
C30 11.16 ± 2.92b 4.01 ± 1.59a 26.12 ± 1.70c 26.64 ± 3.05c 11.95 ± 0.74b 10.37 ± 2.25b 3.05 ± 0.08a
C32 5.31 ± 2.68b 2.20 ± 0.28a 13.57 ± 1.73d 7.99 ± 2.48c 51.74 ± 6.57f 47.89 ± 7.30e 3.53 ± 0.06ab

The symbol ‘‘//” refers to compounds of which the amount is less than the detection limit.
a, b, c, d, e: Different letters within the same row indicate significant differences at p < 0.05 – according to Dunnet T’3 post-hoc test (SPSS 22).
HC: fatty hydrocarbon, WE: wax ester, FFA: free fatty acid, FAL: free fatty alcohol, FA moieties: fatty acid moieties and FAL moieties: fatty alcohol moieties.
RBW: rice bran wax, SW: sunflower wax, BW: bees wax, CLW: candelilla wax, CRBW: carnauba Brazilian wax, CRWW: carnauba wild wax, and BEW: berry wax.
Bold values refer to the values of the major components in natural waxes.

are summarized in Table 3. Palmitic acid was present in all waxes in crude SW (34.80%wt) (Kanya et al., 2007) and in RBW
and contributed remarkably to the FFA compositions of BEW which (Dassanayake, Kodali, Ueno, & Sato, 2009), respectively. Triacon-
contained more than 90%wt of short-chain components (C16 and tanol (C30) was identified as the primary FAL in CLW (42.28%wt)
C18). Two types of carnauba wax (CRBW and CRWW) are both and BW (29.23%wt) which coincided with the findings from End-
products derived from the leaves of the palm, Copernicia cerifera, lein and Peleikis (Endlein & Peleikis, 2011). The FAL chromatogram
which were expected to show a similar composition. Lignoceric of BW is clearly indicated in Fig. S2C. There was no significant
acid (C24) was reported as the major FFA in carnauba wax difference in FAL composition between CRBW and CRWW. The
(Asperger et al., 2001), and there was the same proportion of multi principal components ranged from 30 to 34 carbon atoms,
FFAs in both carnauba waxes used in this study. CLW is a unique which confirms the results of Asperger et al., who reported C30
wax containing a high percentage of longer alkyl chain FFAs (triacontanol) and C32 (dotriacontanol) to be the main FAL in
(C24–C32). The length of FFAs in SW and RBW ranged from 16 to carnauba wax (Asperger, Engewald, & Fabian, 1999).
24 carbon atoms which is congruent with Kanya et al., who
identified the predominant FFAs in crude SW to be arachidic acid 3.2.2. Fatty acid and fatty alcohol moieties from wax esters in natural
(C20) and behenic acid (C22) (Kanya et al., 2007). The chromatogram waxes
in Fig. S2B suggests that more than 95% of the FFA fraction in BW The saponification procedure severed the wax esters into fatty
consisted of FFAs with an even chain length ranging from 12 to 34 acid and fatty alcohol moieties which were derived for the
carbon atoms, in which C16 and C24 were the primary constituents. characterization in GC–MS. Table 3 shows that the FA and FAL
The results are consistent with those from Tulloch et al., (Tulloch, moieties of the wax ester fraction consisted of components with
1973). an even alkyl chain ranging between 16 and 32 carbon atoms.
The free FAL portion in natural waxes was largely comprised of The prevalent FA moieties of WEs were behenic acid (C22) and
components with an even alkyl chain ranging between 22 and 34 lignoceric acid (C24) in RBW, and arachidic acid (C20) in SW, which
carbon atoms. SW and RBW possessed the similar FAL composition, are comparable with the reported values (Kanya et al., 2007; Vali,
which contained mainly C24, C26, C28 and C30 FAL. In comparison, Ju, Kaimal, & Chern, 2005). In general, the chain-lengths of FA
C24 FAL and C30 FAL were reported to be the principal alcohol moieties did not exceed 24 carbons. The dominant FA moieties of
722 C.D. Doan et al. / Food Chemistry 214 (2017) 717–725

WEs in BW were C16 and C18 (Fig. S2D). FA moieties of WEs in BEW broadened and shifted to much lower temperatures as compared
were rich in short-chain components (C16–C18) with approximately to the peaks of the neat waxes (Fig. S4). As a result from the
80% of palmitic acid. FA moieties of WEs in CRBW and CRWW were dilution effect of various wax constituents, only the predominant
similar, with C16, C18 and C24 FA as the predominant constituents. components in that wax govern the crystallization of wax-based
The FALs in the WE fraction ranged between 18 and 32 carbon oleogels. A high amount of long-chain WEs causes RBW and SW
atoms (Table 3). The prevalent FAL moieties of WEs in RBW and SW to crystallize in RBO at very high peak temperatures
were lignoceryl (C24) and ceryl (C26) alcohol, which differs from the (53.33 ± 0.10 °C and 49.33 ± 0.16 °C, respectively). For BW-oleogel,
report of Vali et al., who found C30 as the primary FAL of WEs in the more intensive peak is the first one (Tc = 42.30 ± 0.12 °C), which
RBW (Vali et al., 2005). However, the FAL moieties in SW are com- is specified by the most prominent component in BW (WEs), followed
parable to those reported by Kanya et al. and Carelli et al., but in by the second peak (Tc = 28.15 ± 0.25 °C) originating from HC fraction
different percentages (Carelli, Frizzera, Forbito, & Crapiste, 2002; (the second major chemical class in BW). For CLW-oleogel, the most
Kanya et al., 2007). C24, C26, C28, C30 and C32 were the predominant intensive crystallization peak at 39.82 ± 0.18 °C originated from HCs
FALs present in WEs of BW (Fig. S2E). Triacontanyl palmitate (C46) (mainly 82.48% of hentriacontane C31), emerging after the smaller
was established as the principal WE in BW (Tulloch, 1973), which peak at 43.66 ± 0.23 °C of WE portion. CRBW and CRWW possessing
is confirmed in the current study by a high relative amount of C16 similar chemical composition, were expected to display similar crys-
FA moiety (80.27%wt) and C30 FAL (26.12%wt) moiety. Triacontanol tallization behavior in RBO. The most prominent peaks at high tem-
(C30) was also the dominant FAL in WEs of CLW, followed by C18 peratures (Tc, CRBW = 53.89 ± 0.15 °C and Tc, CRWW = 56.09 ± 0.20 °C)
and C28 FAL. are indicative of WE fraction. For CRBW-gel, this was followed by
The results suggest that the chain length of wax esters in these two minor peaks at 46.54 ± 0.21 °C (of FAs) and 36.29 ± 0.10 °C (of
natural waxes varied from 32 to 68 carbon atoms. Most natural FALs). However, CRWW-oleogel displayed only one minor peak at
WEs were comprised of a short-chain FA moiety bound to a 33.75 ± 0.12 °C (of FALs). BEW-oleogel showed its late crystallization
longer-chain FAL moiety. The hydrolysis of WEs to its building at a low temperature (Tc, onset = 9.22 ± 0.03 °C, Tc = 6.82 ± 0.12 °C)
blocks and the subsequent analysis makes it challenging to deter- because of 82% of short-chain palmitic acid (C16) existing in BEW.
mine how the moieties are bound to each other, but it demon- At the minimum gelling concentration, the oleogel prepared from
strates a rationale to explore innovative techniques for chemical SW, BW, CLW and BEW exhibited a higher G0 LVR and critical stress
characterization of natural waxes. The suggesting predominant as compared to those of RBW, CRBW and CRWW. Among the strong
WEs in waxes might be: C44–C50 in SW, C44–C52 in RBW, C40–C48 gels, CLW and BEW showed the highest consistency and a very low
in BW, C34–C52 in CLW, C34–C56 in CRBW and CRWW, and flow yield stress at the region of low shear. In the weak gel group,
C34–C38 in BEW. RBW displayed a higher storage modulus on linear visco-elastic
region (G0 LVR) but lower critical stress and oscillation yield stress
3.3. Macro- and microscopic analysis of oleogels prepared from rice compared to CRWW. As expected, CRBW and CRWW, which have a
bran oil at minimum gelling concentration of natural waxes similar chemical composition, displayed similar weakness in gelation
although a higher amount of CRBW (4%wt) was required for the
The oleogels were prepared in RBO at the minimum gelling con- gelation of RBO compared to CRWW (2%wt).
centration of each wax: 1.0%wt (SW, CLW, and BEW); 1.5%wt X-ray diffraction measurements were done at two ranges of
(BW); 5.0%wt (RBW), 2.0%wt (CRWW) and 4.0%wt (CRBW) (Doan angles: wide diffraction angle (WAXD) for short-spacing runs and
et al., 2015). The rheological gelling behavior of these oleogels small diffraction angle (SAXD) for long-spacing runs, which
(the average storage modulus in LVR or G0 LVR, critical stress, provides information on the lateral and longitudinal packing of
oscillation yield stress, flow yield stress and consistency index K), molecules in crystalline structures, respectively (Bouzidi, Li, &
Tc, onset and Tm, onset were subsequently investigated (Table 4). Narine, 2014) (Fig. 1). In the wide diffraction angles (2 theta > 15°),
Under cooling, single or mono-component lipidic classes such all the wax-based oleogels (except for BEW-oleogel) exhibited
as FFAs, FALs, HCs or WEs are able to form crystalline building two separated intensive and narrow peaks at d-spacing
blocks that can immobilize a high amount of liquid oil into a of 0.415 nm and 0.372 nm (d = n * wavelength/2sin(theta) =
self-supporting solid-like system (oleogel) (Marangoni, 2012). As 1.54178/2 * sin(theta)), which are indicative of the orthorhombic
discussed, waxes are multi-component materials with a prominent sub-cell structure (b0 morphology). The BEW-oleogel had only
compound and other minor constituents. Hence, the gelation one extremely small but true peak at d-value of 0.415 nm (SAXD),
mechanism from the waxes is considered as a complicated activity which is a characteristic of a hexagonal symmetry. It is noted that
of a mixture containing multi-gelators. The wax-based oleogel is a neat BEW exhibited a low melting and crystallization temperature
result from the dilution of neat wax in RBO, explaining why the (16.25 ± 0.10 °C) while other waxes demonstrated higher thermal
crystallization and melting peaks of the wax-based oleogels were profiles (>50 °C). According to Larsson, when the orthorhombic

Table 4
Rheological and thermal properties of wax-based oleogels prepared at minimum gelling concentration.

RBW SW BW CLW CRBW CRWW BEW


Concentration of wax (%) 5.0 1.0 1.5 1.0 4.0 2.0 1.0
Viscous modulus (G0 LVR) (Pa) 54.08 ± 0.02b 433.34 ± 31.44e 249.33 ± 47.14c 340.77 ± 36.14d 30.16 ± 0.49a 38.71 ± 1.48a 274.88 ± 30.06c
Critical stress (Pa) 0.037 ± 0.005a 0.1995 ± 0.00c 0.1585 ± 0.00b 0.3981 ± 0.00d 0.063 ± 0.00a 0.063 ± 0.00a 0.6309 ± 0.00e
Oscillation yield stress (Pa) 4.53 ± 0.41bc 11.95 ± 0.38e 2.51 ± 0.00b 8.77 ± 2.13d 0.39 ± 0.02a 0.63 ± 0.00a 6.31 ± 0.00cd
Flow yield stress (Pa) 2.12 ± 0.26c 5.82 ± 0.03e 3.81 ± 0.09d 0.00 ± 0.00a 0.45 ± 0.04b 3.56 ± 0.30d 0.00 ± 0.00a
Consistency index K (Pa.sn) 2.42 ± 0.09ab 3.66 ± 0.47b 5.42 ± 0.35c 10.97 ± 0.75d 2.26 ± 0.40ab 1.37 ± 0.33a 10.96 ± 0.34d
Tc, onset (°C) 56.5 ± 0.04de 54.58 ± 0.6d 42.4 ± 0.17c 34.78 ± 2.07b 55.95 ± 0.82d 59.73 ± 0.51e 9.22 ± 0.03a
Tm, onset (°C) 54.39 ± 0.17e 34.54 ± 0.42d 19.6 ± 0.03b 19.94 ± 1.04b 23.45 ± 0.95c 24.56 ± 0.08c 6.54 ± 0.38a

Tc, onset: onset crystallization temperature.


Tm, onset: onset melting temperature.
RBW: rice bran wax, SW: sunflower wax, BW: bees wax, CLW: candelilla wax, CRBW: carnauba Brazilian wax, CRWW: carnauba wild wax, and BEW: berry wax.
a, b, c, d, e, f: Different letters within the same row, indicate significant differences at p < 0.05 – according to Tukey post-hoc test (SPSS 22).
C.D. Doan et al. / Food Chemistry 214 (2017) 717–725 723

Fig. 1. X-ray diffractogram of wax-based oleogels at 5 °C.

waxes show a phase transition at temperatures close to their melt- developed in a unidirection, resulting in the needle-like form as
ing points, the orientation order of the ‘‘zig-zag” planes are lost and seen in Fig. 2. Chambers, Ritchie, and Booth (1976) reported that
the molecules can rotate around their length axis to form a hexag- WEs dominated the platelet crystals, which was confirmed by Koch
onal symmetry, which could be the case for BEW-oleogel (Larsson and Ensikat in 2008 (Koch & Ensikat, 2008). For RBW-oleogel,
& Larsson, 1994). All the wax-based oleogels showed clear peaks besides the existence of some platelet-crystals, RBW-oleogel also
(with different intensities) at the low angle region, providing displayed the long dendritic crystals, which interconnected to form
evidence for a lamellar packing. The SAXD diffractograms of the a branched network with many voids resulting in the weak gelling
oleogels from RBW, SW and BEW show the appearance of one ability of RBW. As discussed above, RBW and SW possessed similar
intense and strong reflection (0 0 1) peak, followed by the intensity chemical composition, in which the most remarkable difference is
decrease of other higher-order peaks. This intensity reduction can substantially higher amount of C24-COOH and C24 moiety in RBW
be explained by the layers of different thickness being stacked than in SW (28% vs 4%, respectively). In addition, neat RBW dis-
together. In addition, Ensikat et al. stated that this intensity played 2 distinct peaks on the crystallization curve as compared
reduction, or even extinction of certain (0 0 1) peaks is the unique to only one intensive peak of neat SW (Fig. S4A). We are convinced
property of some compounds present in plant waxes, and is due that the minor crystallization peak of neat RBW at 59.05 ± 0.13 °C
to the specific position of the oxygen atoms in the long-chain originated from FA fraction (the second prominent compound in
molecules (Ensikat, Boese, Mader, Barthlott, & Koch, 2006). The RBW) and is related to the appearance of the big dendritic crystals.
intense and narrow long-spacing peaks at the first half 2h angles Although this minor peak diminished on the crystallization curve
are indicative of the double-layer order structure when the com- of RBW-oleogel due to the dilution effect, the related FA fraction
pounds with terminal polar groups (FAs, primary FALs) crystallize still existed in a relatively sufficient amount to form big denritic
with a head-to-head orientation (Koch & Ensikat, 2008). This is crystals. Their presence would interfere with the crystallization
true for the oleogels prepared from RBW, SW and BEW. The of the major WE peak, which is important for the formation of
oleogels prepared from BW, CLW, CRBW and CRWW had only platelet crystals and a strong gelling network. This explains why
one (0 0 1) SAXD peak (Table S2) with much lower intensity, indi- RBW was not efficient in structuring RBO, this is in complete con-
cating a significantly lower order of the layer structure, which is a trast to the results obtained by Sato et al. (Dassanayake et al.,
feature of mixtures of compounds with varying chain lengths 2009), who reported that RBW could structure olive oil and salad
(Ensikat et al., 2006). oil into good gels, even at 1.0%wt. We also compared the structur-
In 2009, Sato et al., reported that a difference in intensity ing properties of RBW and SW (the type used in this study) in
between long-spacing and short-spacing peaks (as is the case of gelling different oils (high oleic sunflower oil, olive oil, peanut
the oleogels from BW, CLW) is the result of a strong anisotropy oil, walnut oil and hazelnut oil). We found that the crystal habits
in crystal growth rates between directions vertical to the lamellar of RBW and SW were independent on the oil type. RBW always
plane and directions within the lamellar plane. The Van der Waals exhibited the big dendritic crystals and formed weak oleogels
interaction between the long hydrocarbon chains and the polar while SW always revealed the platelet crystals connecting in a
ester groups (the molecular interactions within the lamellar plane) dense and strong network. It can thus be claimed that the minor
are far stronger than the molecular interactions through the FA fraction in RBW formed the dendritic molecules which inter-
terminal –CH3 groups. This mechanism leads to the appearance fered with the network development of platelet crystals, resulting
of rod-like or platelet morphologies, which are desirable for in a weak gel structure.
the formation of a strong three-dimensional gelling network The molecular order of the minor components might influence
(Fig. 2) (Dassanayake et al., 2009; Hwang et al., 2012). This is con- the wax crystal shape, and the two morphological wax types
firmed by the strong gelation of SW, CLW, BW and BEW in RBO, (tubules and platelets) were able to self-assemble (Birkett &
which exhibited the tiny needle-like crystals, which are actually Talbot, 2009). Clusters of self-assembled crystals might form an
the edges of platelet crystals (Blake & Marangoni, 2015a). It is aggregate in different ways leading to a stronger or weaker net-
important to note that the crystals shown in Fig. 2 were heated work. The crystalline aggregates of CRBW (Fig. S3) and CRWW
and re-crystallized between the narrow space of the glass slide (Fig. 2) oleogels revealed a similar morphology of spherulites and
and the cover slide. Under cooling, the edges of the platelets dendritic morphology explaining the weaker gel formation: the
724 C.D. Doan et al. / Food Chemistry 214 (2017) 717–725

Fig. 2. Microstructure of wax-based oleogels at 5 °C.

low critical stress (0.063 Pa) and quick flowing characteristic at The gelator with a higher amount of FFAs and HCs can be com-
very small oscillation and shear stress regions (Table 4). CRBW bined in food products which require consistency, smoothness and
and CRWW are comprised of approximately 60% of WEs, 6% of FFAs plasticity like whipped cream, ice cream, confectionery products
and 30% of FALs, which were expected to positively contribute to (spread, fillings), and spreading margarine. However, FFAs could
the oil gelation. The long-chain FALs (P24 carbons) in CRBW and be oxidized during heat processing and storage, resulting in off-
CRWW also govern the oil structuring through H-bonding; how- flavor. For ice cream, berry wax with a high amount of FAs (low
ever, the activity of FALs has a threshold limit. The hydrogen bond- melting temperature) can be a good choice because of the low-
ing between very long chain FALs present in CRBW and CRWW temperature working condition. Waxes with a prominent amount
(C30, C32, and C34) becomes stronger as the size and mass of the of WEs (SW, CRBW, CRWW or RBW) can be utilized as a good alter-
FALs increases (Mudge, 2005). Such hydrogen bonding in long native for shortening or baking margarine.
chain FALs, according to Zhu and Dordick, contributes to the
overall gelator-gelator interaction which compromises the 4. Conclusion
solvent-gelator H-bonding, resulting in a poor gelation (Zhu &
Dordick, 2006). In this study, the crystallization and gelling activity of natural
A statistical Pearson correlation (details in Supporting informa- waxes in RBO has been explained by the presence of different
tion) was executed to understand the relationship between the chemical components in waxes. More WEs will result in a strong
gelling properties and chemical composition in waxes (Table S3). and brittle gel, which is expressed in a high storage modulus and
There is a positive correlation between oscillation yield stress flow yield stress of the wax-based oleogels. HCs and FFAs con-
and the average storage modulus (G0 LVR) which displays the tribute to the consistency and stability of the wax-based oleogels
strength of the oleogels, and between the critical stress and (a high critical stress and high consistency index). However, the
consistency index K which reflects the flexibility, smoothness chain length needs to be taken into account because more short-
and consistency of the oleogels. Based on the Pearson correlation, chain FFAs (<16 carbons) result in wax crystallization at low tem-
the gel strength including G0 LVR and oscillation yield stress has a peratures, which can limit its ability in the food application. FALs
negative correlation with FALs and a positive correlation with also have an important contribution to the oil structuring. How-
HCs, FFAs and WEs. The gel strength will increase with the decreas- ever, the strong H-bonding between very long-chain FALs (>30 car-
ing amount of FALs and the increasing amounts of HCs, FFAs, and bons) in waxes and RBO compromises the oil-wax H-bonding and
WEs. Similarly, the critical stress and consistency index K have a hinders the gelation. The results of our study provide important
positive correlation with FFAs and HCs, and a negative correlation information relating to the elucidation of wax components in liq-
with WEs and FALs. The flow yield stress is positively correlated to uid oil structuring, which can be exploited to select a wax with
the amount of WEs and negatively correlated to FALs, FFAs and properties desirable for food application.
HCs. More WEs will lead to a strong but brittle gel with a high
initial flow yield stress. The brittleness is determined by the Acknowledgements
high gel strength and a narrow LVR as well as yield zone, in which
the slope of the curve gives information about the breakage This research has been funded with support from the European
of intermolecular forces holding up the structure (Patel, Commission. This publication reflects the views only of the author,
Babaahmadifooladi, Lesaffer, & Dewettinck, 2015; Uhlherr et al., and the Commission cannot be held responsible for any use which
2005). We would like to stress that the oil model used in this may be made of the information contained therein. Authors also
research was RBO. Other liquid oils with different triacylglycerols thank Marie Curie Career Integration Grant (SAT-FAT-FREE) and
will differently contribute to the gelling properties of waxes, for Vandemoortele Centre for microscope support, and colleague Iris
which the research is ongoing. Tavernier and Kim Moens for scientific inputs.
C.D. Doan et al. / Food Chemistry 214 (2017) 717–725 725

Appendix A. Supplementary data Hwang, K. T., Cuppett, S. L., Weller, C. L., & Hanna, M. A. (2002). HPLC of grain
sorghum wax classes highlighting the separation of aldehydes from wax esters
and steryl esters. Journal of Separation Science, 25, 619–623.
Supplementary data associated with this article can be found, in Kanya, T. S., Rao, L. J., & Sastry, M. S. (2007). Characterization of wax esters, free fatty
the online version, at http://dx.doi.org/10.1016/j.foodchem.2016. alcohols and free fatty acids of crude wax from sunflower seed oil refineries.
Food Chemistry, 101(4), 1552–1557.
07.123.
Koch, K., & Ensikat, H.-J. (2008). The hydrophobic coatings of plant surfaces:
Epicuticular wax crystals and their morphologies, crystallinity and molecular
References self-assembly. Micron, 39(7), 759–772.
Kolattukudy, P. E. (1976). Chemistry and biochemistry of natural waxes . Elsevier
Asikin, Y., Takahashi, M., Hirose, N., Hou, D. X., Takara, K., & Wada, K. (2012). Wax, Scientific Pub. Co.
policosanol, and long-chain aldehydes of different sugarcane (Saccharum Larsson, K., & Larsson, K. (1994). Lipids: Molecular organization, physical functions and
officinarum L.) cultivars. European Journal of Lipid Science and Technology, 114 technical applications. UK: Oily Press Dundee.
(5), 583–591. Lee, S., & Inglett, G. (2007). Effect of an oat b-glucan-rich hydrocolloid (C-trim30) on
Asperger, A., Engewald, W., & Fabian, G. (1999). Analytical characterization of the rheology and oil uptake of frying batters. Journal of Food Science, 72(4),
natural waxes employing pyrolysis–gas chromatography–mass spectrometry. E222–E226.
Journal of Analytical and Applied Pyrolysis, 50(2), 103–115. Lluveras, A., Bonaduce, I., Andreotti, A., & Colombini, M. P. (2009). GC/MS analytical
Asperger, A., Engewald, W., & Fabian, G. (2001). Thermally assisted hydrolysis and procedure for the characterization of glycerolipids, natural waxes,
methylation – A simple and rapid online derivatization method for the gas terpenoid resins, proteinaceous and polysaccharide materials in the same
chromatographic analysis of natural waxes. Journal of Analytical and Applied paint microsample avoiding interferences from inorganic media. Analytical
Pyrolysis, 61(1), 91–109. Chemistry, 82(1), 376–386.
Bicalho, B., David, F., Rumplel, K., Kindt, E., & Sandra, P. (2008). Creating a fatty acid Marangoni, A. G. (2012). Organogels: An alternative edible oil-structuring method.
methyl ester database for lipid profiling in a single drop of human blood using Journal of the American Oil Chemists’ Society, 89(5), 749–780.
high resolution capillary gas chromatography and mass spectrometry. Journal of Marangoni, A. G., & Garti, N. (2011). Edible oleogels: Structure and health implications.
Chromatography A, 1211(1), 120–128. AOCS Press.
Birkett, J., & Talbot, G. (2009). Fat-based centres and fillings. In Science and Marinach, C., Papillon, M.-C., & Pepe, C. (2004). Identification of binding media in
technology of enrobed and filled chocolate, confectionery and bakery products works of art by gas chromatography–mass spectrometry. Journal of Cultural
(pp. 101–122). Heritage, 5(2), 231–240.
Blake, A. I., & Marangoni, A. G. (2015a). Plant wax crystals display platelet-like Megoulas, N. C., & Koupparis, M. A. (2005). Twenty years of evaporative light
morphology. Food Structure, 3, 30–34. scattering detection. Critical Reviews in Analytical Chemistry, 35(4), 301–316.
Blake, A. I., & Marangoni, A. G. (2015b). The use of cooling rate to engineer the Moreau, R. A., Kohout, K., & Singh, V. (2002). Temperature-enhanced alumina HPLC
microstructure and oil binding capacity of wax crystal networks. Food method for the analysis of wax esters, sterol esters, and methyl esters. Lipids, 37
Biophysics, 10(4), 456–465. (12), 1201–1204.
Bouzidi, L., Li, S., & Narine, S. S. (2014). Lubricating and waxy esters. VI. Effect of Mudge, S. M. (2005). Fatty alcohols – A review of their natural synthesis and
symmetry about ester on crystallization of linear monoester isomers. Symmetry, environmental distribution. The Soap and Detergent Association, 132, 1–141.
6(3), 655–676. Nordbäck, J., & Lundberg, E. (1999). High resolution separation of non-polar lipid
Carelli, A. A., Frizzera, L. M., Forbito, P. R., & Crapiste, G. H. (2002). Wax composition classes by HPLC–ELSD using alumina as stationary phase. Journal of High
of sunflower seed oils. Journal of the American Oil Chemists’ Society, 79(8), Resolution Chromatography, 22(9), 483–486.
763–768. Odlyha, M. (1995). Investigation of the binding media of paintings by
Carriere, C., & Inglett, G. (2000). Effect of processing conditions on the viscoelastic thermoanalytical and spectroscopic techniques. Thermochimica Acta, 269,
behavior of nu-trimx: An oat-based b-glucan-rich hydrocolloidal extractive1. 705–727.
Journal of Texture Studies, 31(2), 123–140. Patel, A., Babaahmadifooladi, M., Lesaffer, A., & Dewettinck, K. (2015). Rheological
Chambers, T., Ritchie, I., & Booth, M. A. (1976). Chemical models for plant wax profiling of organogels prepared at critical gelling concentrations of natural
morphogenesis. New Phytologist, 77(1), 43–49. waxes in a triacylglycerol solvent. Journal of Agricultural and Food Chemistry, 63
Dassanayake, L. S. K., Kodali, D. R., Ueno, S., & Sato, K. (2009). Physical properties of (19), 4862–4869.
rice bran wax in bulk and organogels. Journal of the American Oil Chemists’ Regert, M. (2009). Direct mass spectrometry to characterise wax and lipid materials.
Society, 86(12), 1163–1173. Organic Mass Spectrometry in Art and Archaeology, 97–129.
Dassanayake, L. S. K., Kodali, D. R., Ueno, S., & Sato, K. (2012). Crystallization kinetics Regert, M., Langlois, J., & Colinart, S. (2005). Characterisation of wax works of art by
of organogels prepared by rice bran wax and vegetable oils. Journal of Oleo gas chromatographic procedures. Journal of Chromatography A, 1091(1),
Science, 61(1), 1–9. 124–136.
Doan, C. D., Van de Walle, D., Dewettinck, K., & Patel, A. R. (2015). Evaluating the oil- Toro-Vazquez, J. F., Morales-Rueda, J. A., Dibildox-Alvarado, E., Charó-Alonso, M.,
gelling properties of natural waxes in rice bran oil: Rheological, thermal, and Alonzo-Macias, M., & González-Chávez, M. (2007). Thermal and textural
microstructural study. Journal of the American Oil Chemists’ Society, 92(6), properties of organogels developed by candelilla wax in safflower oil. Journal
801–811. of the American Oil Chemists’ Society, 84(11), 989–1000.
Endlein, E., & Peleikis, K.-H. (2011). Natural waxes-properties, compositions and Tulloch, A. (1973). Comparison of some commercial waxes by gas liquid
applications. SÖFW-Journal, 137(4). chromatography. Journal of the American Oil Chemists Society, 50(9), 367–371.
Ensikat, H., Boese, M., Mader, W., Barthlott, W., & Koch, K. (2006). Crystallinity of Uhlherr, P., Guo, J., Tiu, C., Zhang, X.-M., Zhou, J.-Q., & Fang, T.-N. (2005). The shear-
plant epicuticular waxes: Electron and X-ray diffraction studies. Chemistry and induced solid–liquid transition in yield stress materials with chemically
Physics of Lipids, 144(1), 45–59. different structures. Journal of Non-Newtonian Fluid Mechanics, 125(2), 101–119.
Fuchs, B., Süß, R., Nimptsch, A., & Schiller, J. (2009). MALDI-TOF-MS directly Vali, S. R., Ju, Y.-H., Kaimal, T. N. B., & Chern, Y.-T. (2005). A process for the
combined with TLC: A review of the current state. Chromatographia, 69(1), preparation of food-grade rice bran wax and the determination of its
95–105. composition. Journal of the American Oil Chemists’ Society, 82(1), 57–64.
Hamilton, R. J. (1995). Waxes: Chemistry, molecular biology and functions. Oily Press Zhu, G., & Dordick, J. S. (2006). Solvent effect on organogel formation by low
Dundee. molecular weight molecules. Chemistry of Materials, 18(25), 5988–5995.
Hwang, H.-S., Kim, S., Singh, M., Winkler-Moser, J. K., & Liu, S. X. (2012). Organogel
formation of soybean oil with waxes. Journal of the American Oil Chemists’
Society, 89(4), 639–647.

Potrebbero piacerti anche