Accepted Manuscript Principles and applications of spectroscopic techniques for evaluating food protein conformational changes: A review Kaiqiang Wang, Da-Wen Sun, Hongbin Pu, Qingyi Wei PII:
S0924-2244(17)30090-0
DOI:
10.1016/j.tifs.2017.06.015
Reference:
TIFS 2035
To appear in:
Trends in Food Science & Technology
Received Date: 16 February 2017 Revised Date:
12 June 2017
Accepted Date: 13 June 2017
Please cite this article as: Wang, K., Sun, D.-W., Pu, H., Wei, Q., Principles and applications of spectroscopic techniques for evaluating food protein conformational changes: A review, Trends in Food Science & Technology (2017), doi: 10.1016/j.tifs.2017.06.015. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT 1
Principles and Applications of Spectroscopic Techniques for Evaluating
2
Food Protein Conformational Changes: A Review
3
Kaiqiang Wang1,2,3, Da-Wen Sun1,2,3,4∗, Hongbin Pu1,2,3, Qingyi Wei1,2,3
5
1
School of Food Science and Engineering, South China University of Technology, Guangzhou 510641, China 2
6
RI PT
4
Academy of Contemporary Food Engineering, South China University of Technology, Guangzhou Higher Education Mega Center, Guangzhou 510006, China
Engineering and Technological Research Centre of Guangdong Province on Intelligent Sensing and Process
M AN U
3
8
SC
7
Control of Cold Chain Foods, Guangzhou Higher Education Mega Center, Guangzhou 510006, China
9 4
10
Food Refrigeration and Computerized Food Technology (FRCFT), Agriculture and Food Science Centre, University College Dublin, National University of Ireland, Belfield, Dublin 4, Ireland
11
TE D
12
Abstract
14
Background: Proteins are essential nutrients required in various body functions and normal human
15
life. However, in the food industry, the application of proteins especially those of plant origin have
16
been limited due to their poor functionality. Although nowadays, diverse modification techniques are
17
usually employed to improve their performance in food products, it is also important that effective
18
methods for monitoring the resultant conformational changes induced during protein modification
19
are developed.
20
Scope and approach: In this review, the relationship between protein conformation and functionality
21
is briefly discussed. Thereafter, the underlying principles behind five selected spectroscopic
AC C
EP
13
∗
Corresponding author. Tel: +353-1-7167342; Fax: +353-1-7167493.
E-mail address:
[email protected]. Website: www.ucd.ie/refrig; www.ucd.ie/sun. 1
ACCEPTED MANUSCRIPT techniques i.e. Fourier transform infrared, Raman, circular dichroism, fluorescence, and ultraviolet
23
spectroscopies is introduced and their recent applications for monitoring conformational changes that
24
occur during physical, chemical or enzymatic modification of proteins are addressed. In addition, the
25
advantages and limitations of each spectroscopic technique are comparatively discussed and
26
perspectives on the current situation alongside future trends are highlighted.
27
Key findings and conclusions: Spectroscopic techniques present an attractive panacea for evaluation
28
of conformational changes during protein modification. Although certain challenges especially with
29
complex food materials require urgent attention thus, more robust spectroscopic solutions should be
30
exploited in the future.
31
Keywords: Protein conformation, Fourier transform infrared spectroscopy, Raman spectroscopy,
32
circular dichroism spectroscopy, fluorescence spectroscopy, ultraviolet spectroscopy
M AN U
SC
RI PT
22
33
1. Introduction
35
Proteins are considered principal components for body metabolism and general human life. In recent
36
years, there has been an increasing interest in utilizing proteins from various sources as functional
37
ingredients in food products, primarily due to their high nutritional value and unique functionality.
38
The functional properties of protein in food including water- and fat-binding capacities, gel forming
39
and rheological behaviors, emulsifying capabilities, foaming capabilities etc., providing desirable
40
sensory characteristics such as structure, texture, flavor, and color during food product formulation
41
(Buckow et al., 2013; Hu et al., 2015b; Lam & Nickerson, 2013; Whitford, 2013). To cite an
42
example, the emulsifying properties of proteins permit the formation of emulsions owing to their
43
amphiphilic nature and film-forming capacity, therefore profitable in various food products, drugs
44
and nutrient delivery tactics (Lam & Nickerson, 2013). However, some native proteins, especially
45
those of plant origin exhibit poor functionalities during food manufacturing. For instance, the poor
46
digestibility of red kidney bean protein isolates (Yin et al., 2008), low solubility of soy protein
AC C
EP
TE D
34
2
ACCEPTED MANUSCRIPT isolates (Hu et al., 2013), poor gelation property of rapeseed protein isolates (He et al., 2014), as well
48
as poor solubility, foaming and emulsifying properties of wheat gluten proteins (Agyare et al., 2009).
49
As a result, certain modification treatments including physical (Zhang et al., 2017), chemical
50
(Shilpashree et al., 2015), and enzymatic methods (Wang et al., 2016a; 2016b) are required in order
51
to obtain optimal nutritive value such as high bioactive activities (Perreault et al., 2017), high
52
digestibility (Yin et al., 2008) and low allergenicity (Li et al., 2012; Tong et al., 2012). Such
53
treatments also ameliorate the functional properties of proteins including gelling strength (He et al.,
54
2014), emulsification (Raikos, 2010) and their ability to form foams (Morales et al., 2015).
55
Furthermore, it is well known that the modification of food protein functionality is accompanied with
56
conformational changes, which influence the quality of the end-product. For example, Abaee et al.
57
(2017) found that the hardness of the non-heat-treated whey protein cold-set hydrogels prepared at
58
pH 9.0 was significantly higher than that of the samples obtained at pH 8.0 and pH 7.0. Their results
59
suggested that base-induced denaturation and unfolding of β-lactoglobulin at pH 9.0 caused the
60
formation of more disulfide bonds and hydrophobic interactions, accounting for the increased
61
α-helical structures. Li et al. (2014) illustrated that high-intensity ultrasound could induce
62
conformational changes including a decrease in the α-helical contents and increase in β-sheet, β-turns,
63
and unordered contents, contributing to protein aggregation and gel formation in meat, leading to
64
enhanced gel texture. On the other hand, Rahaman et al. (2016) reported that different processing
65
approaches could affect the conformational changes related to digestibility and allergenicity of food
66
proteins in peculiar ways. Therefore, it is essential to develop effective detection methods for
67
evaluating conformational changes in protein and understanding the relationship between its
68
structural and functional characteristics during processing. Several chemical methods for indirect
69
reflection of protein denaturation in meat are available based on detection of protein solubility, free
70
sulfhydryl content, surface hydrophobicity, or myofibrillar ATPase activity (Chen et al., 2016).
71
However, most of these above-mentioned methods are time-consuming, environmentally unfriendly
AC C
EP
TE D
M AN U
SC
RI PT
47
3
ACCEPTED MANUSCRIPT and do not obtain conformational changes related to specific secondary and tertiary structures.
73
Undoubtedly, spectroscopic techniques in the last few years have significantly improved, however,
74
they are developed and used for protein structure analysis for decades. These techniques include
75
X-ray diffraction (XRD) (Jenkins et al., 2013), nuclear magnetic resonance (NMR) (Mao et al.,
76
2014), Fourier transform infrared (FTIR) (Zhang et al., 2015), Raman (Li et al., 2014), circular
77
dichroism (CD) (Chandrapala et al., 2012), fluorescence (Ruffin et al., 2014) and ultraviolet (UV)
78
(Barrios-Peralta et al., 2012) spectroscopies. These spectroscopic techniques are particularly suitable
79
for probing structural conversion such as folding and unfolding. In particular, XRD and NMR
80
spectroscopies are used to obtain structural information of proteins in high resolution. In fact, XRD
81
is regarded as one of the best methods for protein structure analysis and even minimal
82
conformational changes are detectable if only crystallization in two alternative conformations is
83
possible. However, XRD does not allow for real time conformational transitions analysis. While
84
NMR spectroscopy is only able to detect low molecular weight proteins and has its limitations for
85
applying to proteins larger than a few hundred residues (Demchenko, 2013).
86
Nonetheless, other spectroscopic techniques including FTIR, Raman, CD, fluorescence and UV are
87
simple, rapid, convenient, and have gained increase popularity for monitoring conformational
88
changes during protein modification. To the best of knowledge, a review, which specifically
89
addresses their applications in this area, is currently unavailable. Thus, the current review presents
90
the recent advances pertaining to their application in detection of conformational changes in proteins
91
alongside their advantages and limitations. The underlying principles of aforementioned
92
spectroscopic techniques are also summarized. In addition, certain areas that could be further
93
exploited and future research trends are proposed.
AC C
EP
TE D
M AN U
SC
RI PT
72
94 95
2. Relationship between conformational and functional properties of food protein
96
The conformational and functional properties of food protein are closely related to various inherent 4
ACCEPTED MANUSCRIPT (e.g. amino acids composition) and external (e.g., pH, temperature and ionic strength) factors. The
98
chemical nature of the amino acids side-chain groups decides the shape and overall hydrophobicity
99
of proteins. Generally, proteins tend to assume an elongated rodlike shape when they contain a large
100
number of hydrophilic amino acids residues distributed uniformly in its sequence; in contrast, they
101
tend to assume a globular shape when they contain a large number of hydrophobic residues. In the
102
native form of proteins, the hydrophobic segments are mostly buried inside the core. The surface
103
hydrophobicity and hydrophilicity characteristics of protein surface can mostly affect their solubility
104
characteristics, which govern several functionalities such as thickening, foaming, emulsification, and
105
gelation properties (Hettiarachchy, 2012). The pH value affects protein solubility in aqueous
106
solutions. At the isoelectric pH, the hydrophobic interaction between proteins reaches maximum,
107
inhibiting unfolding of the protein molecules and resulting in the minimum solubility. When proteins
108
are exposed to moderately pH values above or below the isoelectric point, the electrostatic repulsion
109
and ionic hydration promote the solubilization of protein and some functional properties are
110
improved, which might be related to unfolding of the protein and/or activation of buried sulf-hydryl
111
groups. The ionic strength of a solution also determines the overall charge of the protein molecule.
112
Ionic
113
hydrophilicity-hydrophobicity characteristics of the protein surface. Moreover, temperature induced
114
protein denaturation could also alter the surface hydrophobicity of proteins (Damodaran, 1997).
115
Functionalities of denatured food proteins are differ from their native states. Therefore, physical,
116
chemical or enzymatic modifications often used to alter the conformation and functionalities of food
117
proteins. The exposure of hydrophobic regions, for example, can lower the solubility, affect the
118
surface activity as well as alter the water and oil holding abilities of proteins (Shen & Tang, 2012).
119
The unfolding of protein structure can expose more amino acids residues, facilitating the proteolysis
120
and improve the digestibility of proteins (Perreault et al., 2017). In addition, changes in the
121
conformation can inhibit the activity of anti-nutritional factors or protein toxins and lower the
affects
protein
solubility (salting-in
or
salting-out)
dependent
on
the
AC C
EP
strength
TE D
M AN U
SC
RI PT
97
5
ACCEPTED MANUSCRIPT allergenicity (Liu et al., 2013; Rahaman et al., 2016). More details about the relationship between
123
protein conformation and functionalities can be found elsewhere (Hettiarachchy, 2012; Whitford,
124
2013). Apart from the modification conditions, changes in the structural and functional properties of
125
food protein also depend on the state of protein (e.g., fibrous or globular, wet or dry, in liquid or
126
frozen state, and presence of foreign substances or not) (Kuan et al., 2013). For instance, globular
127
proteins that possess complex three-dimensional shape depart considerably from fibrous proteins,
128
showing elongated structures that lack true tertiary structure. Due to the cavities in the folded state of
129
globular proteins, they tend to be more susceptible to hydrostatic pressure than fibrous proteins
130
(Ustunol, 2014). The conformational changes of proteins can be detected by spectroscopic
131
techniques, and spectral characteristics related to protein conformational properties are introduced in
132
details in the current review.
M AN U
SC
RI PT
122
133
3. Principles of spectroscopic techniques
135
3.1 Fourier transform infrared (FTIR) spectroscopy
136
Infrared (IR) spectrum in the range of 400 to 4000 cm-1 arises from the absorption of energy by
137
chemical bonds, primarily stretching and bending motions, and has been recognized as a powerful
138
technique for the structural and chemical characterization of proteins (Carbonaro & Nucara, 2010).
139
Generally speaking, the amide I band (1700-1600 cm-1) of an IR spectrum is primarily attributed to
140
C=O stretching vibrations (approximately 80%) with some in-plane N-H bending and C-H stretching
141
modes. In particular, the C=O stretching vibrations in proteins mainly depend on their various
142
secondary structures and inter- or intramolecular effects, including molecular geometry and
143
hydrogen bonding pattern, which makes the amide I band being the most sensitive IR spectral region
144
to predict the secondary structural components of proteins (Kong & Yu, 2007).
145
When the FTIR absorption of protein is measured in solutions, the strong IR absorbance of H2O
146
centered at 1640 cm-1 from O-H-O bending mode may interfere the determination. In order to
AC C
EP
TE D
134
6
ACCEPTED MANUSCRIPT eliminate the interference, the IR absorption spectrum of protein can be acquired in solid state, i.e.,
148
for FTIR experiment, a protein sample of 2 mg is generally mixed with 198 mg of transparent media
149
(usually KBr) to form pellet (1-2 mm thick) for measurement. Besides, the diffuse reflectance and
150
attenuated total reflection (ATR) FTIR spectrum have also been developed for evaluating food
151
protein secondary structure.
152
The observed amide I band is a complex of several overlapping components that corresponding to
153
specific secondary structures, including α-helices, β-sheets, β-turns, and random coils. However,
154
these broad underlying components bands are instrumentally unresolvable (Carbonaro & Nucara,
155
2010). In order to enhance the resolution of individual underlying component and quantitatively
156
estimate the relative contributions of various secondary structures, Fourier self-deconvolution (FSD)
157
fitting and second derivative analysis should be used to achieve maximum band narrowing, degrade
158
the signal-to-noise ratio and identify different types of secondary structures present in proteins. For
159
instance, Figure 1 presents the handling process of IR spectra in amide I region of pulsed electric
160
field (PEF) treated egg white protein (EWP) (Qian et al., 2016). The peak location of the
161
components does not shift, whereas their areas change significantly. The calculation of the portion of
162
the components indicates that a reduction of α-helices is accompanied by an increase of β-sheets
163
during PEF treatment.
SC
M AN U
TE D
EP
AC C
164
RI PT
147
165
3.2 Raman spectroscopy
166
Raman spectrum has been proven to provide effective information about protein secondary structures
167
and the microenvironment of protein side chains, which can be used as a valid tool for evaluating
168
protein denaturation. The spectrum can be processed identically to the infrared spectroscopy data. In
169
common with IR spectra, among several distinct vibrational modes of the -CO-NH- amide, the most
170
available bands of Raman spectra for determining protein secondary structures are the amide I and
171
III regions. The frequency position of these bands depends strongly on the protein state, the 7
ACCEPTED MANUSCRIPT environment and the intermolecular interactions (Blanpain-Avet et al., 2012).
173
Figure 2 shows the Raman spectra of β-conglycinin by high intensity ultrasound (20 kHz at 400 W)
174
treatment for 0, 5, 20 and 40 min (Hu et al., 2015a). Apart from the amide I and III regions, the major
175
vibrational motions of the side chains, including inter-chain disulfide bands, tryptophan (Trp) bands,
176
tyrosine (Tyr) bands, and aliphatic hydrophobic residues in Raman spectra can also be analyzed to
177
provide some information about protein tertiary structure. The Raman bands located at 760, 880,
178
1361 cm-1 are ascribed to Trp residues. A sharp line at 1361 cm-1 is suggested as an indicator of
179
buried Trp residues (Ferrer et al., 2011). However, the Trp bands at 760 and 880 cm-1 have been
180
proposed as an indicator of the strength of H-bonding and hydrophobicity of the indole ring. An
181
increase in the intensity in these bands indicates that Trp residues are buried, and in contrast a
182
decrease is connected with the opposite phenomenon (Gómez et al., 2013). The Raman Tyr residues
183
vibrations are located at about 850 cm-1 and 830 cm-1, and the ratio of Tyr doublet (I850/I830) is
184
sensitive to the nature of hydrogen bonding and ionization of phenolic hydroxyl groups. Tyr doublet
185
is considered as a good indicator for evaluating the degree of Tyr residues exposed or buried (Gómez
186
et al., 2013).
187
In Raman spectra, the stretching vibration located in the range of 500-550 cm-1 derive from disulfide
188
bridges that are formed by two cysteines, and the peaks around 510, 525, and 545 cm-1 can be
189
assigned to disulfide bonds in gauche-gauche-gauche (g-g-g), gauche-gauche-trans (g-g-t), and
190
trans-gauche-trans (t-g-t) conformations, respectively (Li, 2012). Disulfide bonds belong to the
191
secondary bonds that could maintain the tertiary structure of proteins, thereby changes in disulfide
192
bonds correlate with the alteration of protein tertiary structure. In addition, the vibration of aliphatic
193
amino acid residues in Raman spectra near 2800-3000 cm-1 (C-H stretching) and 1440-1465 cm-1
194
(C-H bending) have also been investigated for monitoring protein conformational changes. Although
195
the indicators of changes in C-H bending vibration are debatable, these changes can also provide
196
information on proteins hydrophobic interactions and conformational changes due to processing
AC C
EP
TE D
M AN U
SC
RI PT
172
8
ACCEPTED MANUSCRIPT 197
(Sheng et al., 2016).
198
3.3 Circular dichroism (CD) spectroscopy
200
CD spectroscopy is another well-established spectroscopic technique for determining secondary
201
structures, folding and binding properties of proteins. In fact, when the plane polarized light passes
202
through a modulator that subjects to an alternating 50 kHz electric field, it will be split into the
203
rotating left-handed (counter-clockwise) and the right-handed (clockwise) circularly polarized
204
components (Kelly et al., 2005). If the two circularly polarized components have the same amplitude,
205
the recombination of the components can regenerate radiation polarized in the original plane.
206
Conversely, the resulting recombined component radiation is then elliptically polarized. The
207
principle of CD spectroscopy is based on the unequal absorption of the two circularly polarized
208
components (∆A=AL-AR). In order to observe CD signals, the sample should be optically active. As
209
all amino acids except glycine are asymmetric and hence optically active, protein structures have
210
been widely studied by CD spectroscopy.
211
Normally, CD spectra are collected in high-transparent rectangular or cylindrical quartz cells. Buffers
212
for dissolving protein samples should be transparent and not contain any materials that are optically
213
active. In addition, protein solutions for CD measurement should be at least 95% pure with
214
concentrations between 0.005 and 5 mg/mL depending on the pathlength of the cell. When CD
215
spectrum is acquired, software such as CDPro, CONTIN, SELCON3, DICROPROT, and CDSSTR
216
and computational methods such as singular value decomposition, optimization algorithms,
217
regression, and neural networks are used for analyzing the CD spectrum. A detailed introduction of
218
the acquirement and analysis of the CD spectrum can be found in a previous review (Martin &
219
Schilstra, 2008).
220
The CD spectrum of a protein in the far UV region is dominated by a weak but broad n→π*
221
transition at about 220 nm and an intense π→π* transition centered around 200 nm of amide groups,
AC C
EP
TE D
M AN U
SC
RI PT
199
9
ACCEPTED MANUSCRIPT and are influenced by the geometries of the polypeptide backbones, making it one of the most
223
commonly used techniques to determine the protein secondary structure content (Whitmore &
224
Wallace, 2008). Generally, α-helices show a strong positive band at 191-193 nm and a typical double
225
negative bands at 208-210 and 222 nm, β-sheets produce an intense positive band at about 195-200
226
nm and a negative band at about 216-218 nm, whereas random coils have a strong negative bands at
227
195-200 nm and a much weaker band (either positive or negative) between 215 and 230 nm. On the
228
other hand, the CD spectrum in near UV region (250-320 nm) can provide useful information related
229
to aromatic chromophores (Phe, Tyr, and Trp residues) of proteins in asymmetric environment, which
230
has been widely used to assess the tertiary and occasionally quaternary structures of proteins during
231
processing. Generally speaking, Phe residues have sharp fine structure between 255 and 270 nm with
232
peaks observed near 262 and 268 nm, whereas the bands arising from Tyr and Trp residues are
233
located at 275-282 nm and 290-305 nm in near-UV CD spectrum, respectively (Kelly et al. 2005;
234
Martin & Schilstra, 2008). An increase in the band magnitudes and intensities is an indication of
235
structural changes, which are related to the loss of native-like structure and increasing interactions of
236
the aromatic amino acid residues during processing (He et al., 2014). Figure 3 shows the CD spectra
237
of myoglobin in solution, absorbed at tricaprin oil/water interface and at hexadecane oil/water
238
interface (Day et al., 2014). As shown in Figure 3a, myoglobin is a high helical protein with strong
239
positive ellipticity at 193 nm and two distinctive negative minima at 208 and 222 nm respectively.
240
Upon adsorption to oil/water interfaces, the reduction in the intensity of peaks at 193, 208 and 222
241
nm suggests the loss of helical structure. Figure 3b shows the near-UV CD spectra, providing the
242
evidence of the change in the Trp residues environment.
243
In recent years, with the improvements in instrumentation for conventional CD and advent of
244
synchrotron radiation circular dichroism (SRCD), lower wavelength bands (as low as 140 nm) and
245
more information about proteins conformation of the spectra are obtainable. SRCD spectroscopy can
246
provide important static and dynamic structural information on proteins in solution. Furthermore, the
AC C
EP
TE D
M AN U
SC
RI PT
222
10
ACCEPTED MANUSCRIPT 247
high flux of a synchrotron source increases the signal-to-noise levels of the CD spectrum, allowing
248
measurement with turbid samples (e.g., in the presence of lipids, salts, and detergents) (Wallace &
249
Janes, 2010).
250
3.4 Fluorescence spectroscopy
252
Fluorescence is the emission of photons due to the absorption of UV or visible light of chromophores
253
that can emit photons. The principles of fluorescence generation can be elucidated by a Jablonski
254
diagram (Karoui & Blecker, 2011). In general, a spectrofluorimeter system comprises of six
255
components: a light source, which is typically a mercury or xenon lamp for emitting UV or Vis light,
256
a sample holder, two monochromator and/or filter(s) with one for selecting the excitation
257
wavelengths and the other for selecting the emission wavelengths, a detector for converting the
258
emitted light to the electronic signal, and a data acquisition unit (Karoui & Blecker, 2011).
259
Fluorescence spectroscopy as one of the oldest and powerful analytical methods has been extensively
260
studied for the analysis of molecular structure and function, as well as protein conformations. In
261
terms of proteins, Trp, Tyr, and Phe residues are natural chromophores and are responsible for
262
fluorescence. The intrinsic Trp fluorescence spectrum is generally used to study protein unfolding
263
and dynamics (Albani, 2008). Figure 4 shows the intrinsic Trp fluorescence emission spectra of black
264
bean protein dispersion with different ultrasonic treatments (Jiang et al., 2014). When Trp residues
265
are fully or partially buried in the hydrophobic core of protein interiors (before ultrasonic treatment,
266
in the environment with a low polarity), Trp fluorescence emission maximum wavelength (λmax) is <
267
330 nm, whereas λmax shifts to a longer wavelength (bathochronic shift) in the presence of a polar
268
environment and the loss of the protein tertiary or quaternary structure (after ultrasonic treatment).
269
For Tyr residues, their λmax locates at about 305 nm. Tyr residues are more fluorescent than Trp
270
residues in solutions, but the fluorescence quantum yield significantly decreases when Tyr is present
271
in proteins. This phenomenon may be interpreted by the fact that the protein tertiary or quaternary
AC C
EP
TE D
M AN U
SC
RI PT
251
11
ACCEPTED MANUSCRIPT structure suppresses Tyr fluorescence. Besides, energy transfer from Tyr to Trp residues occurs in
273
proteins, inducing a certain extent of quenching the Tyr fluorescence. Consequently, Tyr fluorescence
274
is lack of sensitivity to the polarity of the environment and is only used as an intrinsic fluorescent
275
probe in studying Trp-lacking proteins (Munishkina & Fink, 2007). On the other hand, Phe with a
276
fluorescence λmax at near 280 nm is rarely used as fluorescent probe because of its relatively low
277
quantum yield.
278
Moreover, extrinsic fluorescent dyes such as 1- anilinonaphthalene-8-sulfonic acid (ANS),
279
4,4’-bis-1-anilinonaphthalene-8-sulfonate (Bis-ANS), and nile red can attach to proteins via covalent
280
interactions and/or non-covalent interactions, and therefore providing additional possibility for
281
studying proteins conformational changes (Hawe et al., 2008). However, it should be noted that the
282
extrinsic fluorescent dye may also change the protein properties.
M AN U
SC
RI PT
272
283
3.5 UV spectroscopy
285
The near-UV absorption spectra of aromatic amino acid residues in proteins contain abundant
286
information related to protein conformations. UV spectroscopy uses the algorithms for molecular
287
surface topography and for the accessibility of certain groups of atoms to the solvent. The algorithms
288
also enable analysis of the three-dimensional arrangement of atoms and groups within the
289
environment of chromophore groups (Demchenko, 2013). The spectral peaks in the range of 250-265
290
nm correspond to Phe residues, and those in the region 265-280 nm attribute to Tyr-Trp electronic
291
interactions, whereas peaks above 285 nm are identified exclusively as Trp contributions (Randall et
292
al., 2016). Both changes in protein conformation and dissociation, as well as protein denaturation
293
may lead to the change in the microenvironment of one or more aromatic amino acid residues.
294
However, the larger bandwidth of UV absorption spectrum masks most useful information. Other
295
components such as cysteine and histidine can also contribute to the UV absorbance of proteins
296
resulting in rather structure less spectra, thus making it difficult to detect small changes that occur
AC C
EP
TE D
284
12
ACCEPTED MANUSCRIPT due to varying local environment of the chromophores. Figure 5a displays the normal UV spectra of
298
lysozyme before and after PEF treatment. The poor separation of the UV absorption bands causes
299
non-informative for detailed analysis of protein tertiary structure (Zhao & Yang, 2008).
300
Spectrophotometry is an analytical method that using mathematical transformation of normal
301
spectral curve into a derivative, which can extract qualitative and quantitative information from
302
overlapping bands of the analytes. In the derivative near-UV spectrum, the ability to detect and to
303
measure minor spectral features is considerably enhanced (Chang et al., 2017). As a result, the
304
numerical derivative spectra are often calculated to achieve a higher resolution with increased
305
sensitivity in near-UV regions of proteins. As shown in Figure 5b, the UV absorption peaks related to
306
Phe, Tyr and Trp can be easily differentiated from the second-derivative UV spectra of lysozyme.
307
The blue shift of the second-derivative peaks after PEF treatment indicates an unfolding of the
308
tertiary structure. Changes in amplitude can be best described by calculating the ratio (r = a/b) of the
309
two peak to trough values between differences in second-derivative absorbance peaks. The value of
310
‘r’ is greatly relevant to solvent polarity for Tyr, while it is rarely dependent on solvent polarity for
311
Trp.
SC
M AN U
TE D EP
312
RI PT
297
4. Applications in evaluating proteins conformational changes
314
Protein modification commonly refers to purposive alteration in protein conformation by physical,
315
chemical or enzymatic treatments. In such a case, even small alteration in protein conformation are
316
capable of causing significant changes in their physicochemical and functional properties, thus
317
enhancing their utilization as ingredients in the food industry. Therefore, evaluating the
318
conformational changes of the modified protein can be helpful for further understanding of the
319
relationship between the structural and functional properties, and for gaining proteins with desired
320
functional properties suitable for many food formulations. Table 1 summarizes applications of
321
spectroscopic techniques for monitoring proteins conformational changes due to different treatments.
AC C
313
13
ACCEPTED MANUSCRIPT 322
4.1 Physical treatments
324
The modification of protein by physical treatment is generally safe with no chemical reagent addition.
325
Particularly, heat treatment is commonly used. Gelation capacity is an important property of food
326
proteins during heat treatment, and heat-induced denaturation process is accompanied by protein
327
unfolding and an exposure of reactive groups (such as sulfhydryl groups and hydrophobic groups),
328
which is believed to be crucial for protein gel formation (Raikos, 2010). In one study, Liu et al.
329
(2011)
330
intra-/intermolecular disulfide exchanges between pork and fish protein during heating triggered a
331
distinguished gelation process. They found that heating induced unfolding of α-helices and formation
332
of β-sheets was associated with the exposure of reactive groups that were beneficial for
333
protein-protein interaction and gelling. Recently, Wang et al. (2017b) employed UV spectroscopy,
334
intrinsic fluorescence spectroscopy, and FTIR spectroscopy to interpret the mechanism of
335
heat-induced wheat gluten gel formation and observed a pronounced transition towards β-sheet-like
336
structures.
337
Apart from heat treatment, high-intensity ultrasound (HIU), high pressure (HP), dynamic
338
high-pressure microfluidization (DHPM), pulsed electric field (PEF) have also been extensively
339
studied as physical treatment methods for food protein modification. Li et al. (2014) explored HIU
340
(20 kHz, 450 W, and 6 min) for modifying the functional properties of pale, soft and exudative (PSE)
341
chicken breast meat. Raman spectroscopy indicated that HIU treatment appeared to induce changes
342
in the spatial structure of myosin, rendering the unfolding of α-helices, followed by a significant
343
increase in β-sheets, β-turns, and unordered contents. The conformational changes contributed to
344
protein aggregates and gel formation in the meat system, which in turn explained the improvement in
345
gel texture and water retention of the treated PSE-like meat gels. Jia et al. (2010) attempted to utilize
346
ultrasound treatment to accelerate the enzymatic hydrolysis of wheat germ protein. An increase of
Raman
spectroscopy
demonstrated
that
the
difference
Tyr
ratios
and
AC C
EP
TE D
M AN U
SC
used
RI PT
323
14
ACCEPTED MANUSCRIPT ANS fluorescence intensity of wheat germ protein with the increase of ultrasonic power was
348
observed, indicating the exposure of more hydrophobic groups and regions inside the protein and the
349
unfolded protein structure being more beneficial to alcalase hydrolysis. Furthermore, the angiotensin
350
converting enzyme (ACE) inhibitory activity of wheat germ protein hydrolysate treated with
351
ultrasound was higher than that without ultrasound treatment.
352
HP processing is a novel nonthermal method that can be used as a physical treatment for protein
353
modification. Unlike the heating-denatured counterparts, which disrupt protein structure by
354
transferring nonpolar hydrocarbons from the hydrophobic core toward the water, the pressure
355
treatment allows penetration of water into the hydrophobic region interior of the protein matrix.
356
Consequently, the hydration patterns of the protein side chains would significantly affect the
357
structural dynamic properties under high-pressure conditions, and the structure stability is primary
358
influenced by its conformational flexibility to compensate for losses of noncovalent bonds due to
359
relocation of water molecules (Buckow et al., 2013). HP effect potentially accounts for the exchange
360
of S-S/SH and alteration of non-covalent bonds (ionic, hydrophobic, and hydrogen bridges) of
361
proteins. Although the primary structure remains unchanged, the secondary, tertiary, and quaternary
362
structures of protein would unfold and disassociate during HP processing (Tabilo-Munizaga et al.,
363
2014). Li et al. (2012) provided direct evidence that high hydrostatic pressure (HPP) treatment could
364
reduce the allergenic property of soy protein isolates (SPI) and improve the security of SPI for cow
365
milk allergic babies. Meanwhile, the extrinsic emission fluorescence spectroscopy suggested an
366
approximate 11.5-fold increase of fluorescence intensity and a blue shift of λmax from 516 to 466 nm.
367
Their CD spectral analysis also indicated that there was a significant increase in helices and turns
368
contents while reduction occurred in strands and unordered contents. Li et al. (2012) speculated that
369
the epitopes of SPI allergens could be closely related to its secondary structures. In addition, effects
370
of HHP on the modification of the conformational and functional properties of proteins in condensed
371
systems have been extensively studied in recent years. Based on FTIR spectral analysis, Savadkoohi
AC C
EP
TE D
M AN U
SC
RI PT
347
15
ACCEPTED MANUSCRIPT et al. (2016) showed the conformational behavior of soy glycinin after HPP treatment and indicated
373
that the pressurization induced the loosing of β-sheet and α-helical structures and the concomitant
374
increase of random coils in soy glycinin samples with 10, 30 and 60% (w/w) solid contents, while the
375
twelve disulphide linkages assisted in retaining secondary structure in concentrated systems (> 70%,
376
w/w).
377
DHPM technology uses the combined forces of high velocity impact, high-frequency vibration,
378
instantaneous pressure drop, intense shear, cavitation and ultra-high pressures of up to 200 MPa, with
379
a short treatment time (less than 5 s) and continuous operation. Zhong et al. (2012) evaluated the
380
relationship between the antigenicity and conformation of β-lactoglobulin (β-LG) subjected to
381
DHMP treatment. UV, CD and fluorescence spectra characterized that the conformational unfolding
382
and aggregation of β-LG under DHMP were dramatically related to its antigenicity. At low level of
383
DHMP treatment (0.1 - 80 MPa), the disaggregation and unfolding of β-LG were accompanied by an
384
increase in the antigenicity, and the aggregation of β-LG at pressures greater than 80 MPa
385
contributed to a decrease of antigenicity. In another study, Hu et al. (2011) analyzed CD spectra and
386
demonstrated some α-helices and β-turns converted to β-sheets in peanut allergen Ara h2 samples
387
after 60 MPa DHMP treatment, which corresponded with an obvious reduction of antigenicity and
388
presumably was caused by either burying or damaging of the conformational IgG-binding epitopes.
389
On the other hand, PEF is widely used in the food industry for inactivation of microorganisms, which
390
can also be used to preserve nutrients and modify the structure and function of proteins in order to
391
achieve specific and/or desired functional properties. Qian et al. (2016) discussed the effect of PEF
392
on structural properties of egg white protein (EWP) in solid state. UV spectral analysis suggested
393
more Trp exposure, while FTIR spectral analysis revealed quantitative change of the relative portions
394
of molecular secondary structure of EWP. The decrease of α-helices and increase of β-sheets would
395
cause the alteration of the functional properties of proteins.
AC C
EP
TE D
M AN U
SC
RI PT
372
396
16
ACCEPTED MANUSCRIPT 4.2 Chemical treatments
398
Chemical treatments mainly aim to modify -NH2, -OH, -SH, or -COOH in protein side chains, and is
399
an effective approach to improve further the physicochemical and functional properties of food
400
proteins. Chemical modification can be achieved by phosphorylation, Maillard reaction, deamidation,
401
acylation, oxidative and so on.
402
Phosphorylation modification of proteins is very important either in biological systems or in vitro for
403
improving the physicochemical and functional properties of proteins. The major effect of
404
phosphorylation is to increase the net negative charge on protein surface and alter protein
405
conformations. Kaewruang et al. (2014) studied phosphorylation gelatin from the skin of unicorn
406
leatherjacket in solutions (pH 7.0, 65 oC, 1 and 3 h), and their FTIR spectral analysis showed that the
407
phosphate incorporated might affect the helical structure of gelatin mainly via the increased repulsion
408
between charged residues in gelatin chains. Enomoto et al. (2010) employed dry-heating (pH 4.0, 85
409
o
410
OVA (Re-OVA), and their CD spectra of N-OVA and Re-OVA showed double minima at 208 and
411
222 nm, whereas these minima were slightly decreased through phosphorylation by dry-heating.
412
These phenomena revealed that the secondary structure of OVA was scarcely affected by
413
phosphorylation. However, the tertiary structure was significantly altered as suggested by
414
fluorescence spectra.
415
Maillard reaction generally occurs spontaneously during long-time storage or heat treatment, which
416
can effectively improve some functionalities without the addition of extraneous chemicals. Spotti et
417
al. (2014) reported that the molecular weight of dextrans (DX) affected the structural and rheological
418
characteristics of WPI/DX conjugates obtained by Maillard reaction. According to fluorescence
419
spectra, the fluorescence intensity of WPI decreased compared with native WPI, and the effect
420
accentuated more in the case of lower molecular weights of DX. In addition, Zhang et al. (2012)
421
used near-UV CD spectroscopic technique to analyze the tertiary conformations of soy
M AN U
SC
RI PT
397
AC C
EP
TE D
C, 1 day) to phosphorylate ovalbumin (OVA) derived from egg white (N-OVA) and recombinant
17
ACCEPTED MANUSCRIPT β-conglycinin/DX conjugate prepared by Maillard reaction in a crowded liquid system, and showed
423
that the decrease of the near-UV CD spectral intensity of β-conglycinin/DX conjugate was due to the
424
loss of β-conglycinin tertiary structure caused by an exposure of aromatic side chains.
425
Deamidation can convert the amide groups of protein side chains into acid groups and thus increases
426
the number of negative charges, which is attractive for modifying plant proteins that contain a
427
number of glutamine and asparagine residues. In one work, a stronger absorption located at 1641
428
cm-1 or 1640 cm-1 was observed in all deamidated barley proteins, indicating the formation of more
429
flexible or extended structures. An increase of the absorption at 1608 cm-1 in deamidated barley
430
proteins was ascribed to the exposure of hidden parts in native protein molecules during deamidation.
431
These conformational changes might arise from the increase of electronic repulsion and loss of
432
hydrogen bonding, which would facilitate emulsion formation (Zhang et al., 2015). More recently,
433
Liao et al. (2016) investigated the intrinsic fluorescence emission spectra of citric-acid-deamidated
434
wheat gluten (CDWG) and found an increase of a deamidated degree from 25% to 55% and a λmax
435
red shift of CDWG, indicating the expansion of the wheat gluten structure and exposure of
436
hydrophobic groups, which resulted in a significant change in the tertiary structure.
437
In addition, it has been reported that acylation of proteins would lead to an increase in the
438
electrostatic repulsion forces in the protein, hence resulting in a transformation of the conformational
439
and functional characteristics. Acetic anhydride and succinic anhydride are the two most commonly
440
used acylation agents for modifying proteins structural and functional properties. Yin et al. (2010)
441
reported that there was a good relationship between physicochemical properties and conformational
442
features of acetylated and succinylated kidney bean protein isolate (KPI). In their study, intrinsic
443
fluorescence and CD spectroscopic techniques were carried out to investigate tertiary and secondary
444
conformational changes of KPI during acylation. Acetylation and succinylation caused significant
445
and gradual decreases in fluorescence intensity with increasing anhydride-to-protein ratio to 1.0,
446
suggesting protein unfolding during acylation. The increase in negative ellipticity showed the
AC C
EP
TE D
M AN U
SC
RI PT
422
18
ACCEPTED MANUSCRIPT conformational transformation to α-helices or random coils under acylation. Shilpashree et al. (2015)
448
studied succinylation on milk protein concentrate (MPC) and found that succinylation could be used
449
for ameliorating the functional properties of MPC and its application could be extended at a
450
succinylation degree of 90.43%. Simultaneously, the red shift from 347 to 359 nm of the maximum
451
fluorescence wavelength demonstrated that the succinylation led to denaturation of MPC.
452
It should also be noted that proteins are vulnerable to oxidative damage because of their abundance
453
in foods and high oxidation reaction rates. Protein oxidation is a covalent interaction, which could
454
induce protein fragmentation, cross-linking, and conformational changes, leading to decrease of its
455
nutritional value and functional characteristics. Spectrofluorometric methods were demonstrated to
456
be the feasible techniques in evaluating proteins oxide. Wu et al. (2009) investigated intrinsic Trp
457
fluorescence spectra to trace 2, 2’-azobis (2-amidinopropane) dihydrochloride (AAPH) mediated soy
458
protein oxidation, and found AAPH resulted in a decrease in Trp fluorescence intensity and a blue
459
shift of λmax from 330 to 319 nm. In addition, their CD spectra suggested that oxidation led to a
460
gradual loss of α-helix and β-sheet structures. Furthermore, it was found that these conformational
461
changes decreased the water holding capacity and gel strengthen of soy protein gel.
462
TE D
M AN U
SC
RI PT
447
4.3 Enzymatic treatments
464
Compared with physical and chemical treatment, enzymatic modification is more acceptable for
465
improving the functional properties of proteins due to milder process conditions required, easier
466
control of the reaction, high efficiency of modification and less formation of by-products. Among the
467
enzymatic modification, the enzymatic proteolysis, cross-linking and deamidation are commonly
468
used. Proteolysis can result in hydrolysis of peptide bonds and affect proteins primary structure.
469
Therefore, secondary and tertiary structural changes result from proteolysis are not discussed here.
470
Unlike the proteolysis process, transglutaminase (TGase) could catalyze intra- and intermolecular
471
isopeptide bonds cross-linking by an acyl transfer reaction between glutamine (acyl donors) and
AC C
EP
463
19
ACCEPTED MANUSCRIPT lysine residues (acyl acceptors) of the proteins. Currently, a number of researches reported that
473
TGase could be utilized to improve the textural properties of various proteins and spectroscopic
474
techniques are thus used to monitor protein conformational changes. For example, Herrero et al.
475
(2008) used Raman spectra of amide I and amide III bands to analyze structural conversion of
476
cross-linked meat proteins with different amounts of TGase. A significant decrease in α-helices
477
accompanied by an increase in β-sheets and turns percentages upon addition of TGase was observed.
478
Herrero et al. (2008) also observed positive correlation of springiness with β-sheet structure and
479
negative correlation with α-helices content, positive correlation of adhesiveness with α-helices and
480
turns and negative correlation with β-sheets, and positive correlation of hardness and springiness
481
with turns. Furthermore, studies on Trp band at 759 cm-1, Tyr doublet ratio (I850/I830), Raman bands at
482
1450 cm-1 and 2935 cm-1 suggested an alteration of tertiary structure of meat proteins. In another
483
study, CD spectra were analyzed to assess the TGase-induced structural alteration of soybean
484
proteins. The ellipticity around 195, 208 and 216 nm decreased in TGase cross-linked soybean
485
protein, showing that TGase modification induced a more random secondary structure of soybean
486
protein (Song & Zhao, 2014).
487
In recent years, protein-glutaminase (PG) deamidation has gradually gained more and more attention
488
because it is more desirable than chemical deamidation. Miwa et al. (2013) reported that
489
PG-deamidated WPI tended to form a soft texture gel with a higher water-binding capacity, which to
490
some extent was attributed to the structural changes and would be meaningful for practical uses.
491
Structural analysis of WPI by using CD spectroscopy and fluorescence spectroscopy revealed that
492
the partial disruption of the tertiary structures of WPI proteins with respect to Trp residues shift to a
493
more-polar environment by the electrical repulsion of the negative charge derived from carboxyl
494
groups under PG deamidation.
AC C
EP
TE D
M AN U
SC
RI PT
472
495
496
5. Advantages and Limitations 20
ACCEPTED MANUSCRIPT Spectroscopic techniques including FTIR, Raman, CD, fluorescence, and UV spectroscopy have
498
been successfully applied for evaluating conformational changes in proteins. In contrast to their
499
traditional counterparts, known to reflect food protein denaturation indirectly, spectroscopic
500
techniques present considerable advantages. For example, FTIR spectroscopy provides high-quality
501
spectra with spectrum in the region of 1600-1700 cm-1 highly sensitive to alterations in secondary
502
structure. Raman spectroscopy can obtain information about molecular vibrations related to
503
secondary structures, side chains of proteins and interference resulting from H2O is rare in Raman
504
spectra. In addition, it is feasible to analyze samples in many cases with fewer sample preparation
505
procedures, thereby providing the potential for direct, non-destructive, and faster detection of
506
conformational changes in situ. CD spectroscopy is uniquely sensitive to the detection of protein
507
conformational changes at a low concentration. Besides, CD spectra in the far-UV and near-UV
508
region can provide some useful information related to protein secondary structures and tertiary
509
structures, respectively. On the other hand, UV spectroscopy and fluorescence are inexpensive and
510
easy to operate, and can effectively monitor changes in the tertiary structure of proteins. In particular,
511
data acquisition in fluorescence spectroscopy occurs within nanoseconds, thereby making it possible
512
to investigate in-depth thermodynamics involving multiple experiments under different conditions.
513
Table 2 provides a summary of the advantages of these spectroscopic techniques. It should be noted
514
that other methods, including differential scanning calorimetry (Kazemi et al., 2011), sodium dodecyl
515
sulfate-polyacrylamide gel electrophoresis (Perreault et al., 2017), scanning tunneling microscopy
516
(Rinke et al., 2014), hydrogen/deuterium exchange mass spectrometry (Li et al., 2008), small-angle
517
X-ray scattering and multi-angle laser light scattering in conjunction with a size exclusion
518
chromatography (Zhao et al., 2012) could be used with spectroscopies complementarily, which offers
519
further advantages.
AC C
EP
TE D
M AN U
SC
RI PT
497
520
However, certain limitations are associated with some of these spectroscopic techniques. Notably,
521
reliable evaluation of conformational changes in protein is only possible with high sample purity and 21
ACCEPTED MANUSCRIPT strictly defined environmental conditions. Most food systems comprise of complex matrices,
523
containing not only proteins but also some other components such as lipids, starches, pigments, etc.
524
which interfere during analysis and overlap the information of protein spectra, posing difficulties in
525
monitoring protein conformation in real food systems. Even for proteins with high purity, the FTIR
526
spectrum of protein in aqueous conditions may experience interference by H2O. Although this
527
interference can be eliminated by using D2O solution, the H-D substitution may change the protein
528
structural characteristics somewhat in comparison with its native state. In some cases, Raman
529
spectroscopy is greatly restricted by inherent weaker Raman scattering and stronger disturbance of
530
biological fluorescence, which could hamper the gaining of high-resolution Raman spectra. Besides,
531
lengthy laser radiation can generate heat that may alter the conformation of samples, thus affecting
532
the measurement accuracy. In the case of CD, UV and fluorescence spectroscopies, only diluted
533
protein samples can be analyzed. However, diluted samples are not the usual concentrations of
534
protein in food products. Therefore, the sample preparation procedure will be time-consuming. In
535
addition, CD spectra can only provide low-resolution secondary structural information because the
536
accuracy of this technique highly depends on the reference databases verified by other techniques.
537
On the other hand, another drawback of FTIR, Raman and CD spectrometers is their high cost in the
538
instruments.
539
6. Conclusions and future trends
540
In this review, the principles and recent applications of five spectroscopic techniques including FTIR,
541
Raman, CD, UV and fluorescence spectroscopies for protein conformation evaluation are described.
542
Although these techniques cover several advantages for measuring the secondary or tertiary
543
structures, some challenges still exist to realize protein conformation detection in real foodstuffs due
544
to their complex components. Therefore, spectroscopic solutions for protein conformational
545
monitoring of complex food systems should be further developed in the future.
546
Noteworthy, there are increasing studies of Raman spectroscopy to evaluate protein conformational
AC C
EP
TE D
M AN U
SC
RI PT
522
22
ACCEPTED MANUSCRIPT changes in meat products. This broadens the perspective to detect proteins in food products with few
548
preparation procedures, and provides the potential for on-line and in-situ monitoring protein
549
conformational changes in the course of different types of treatments, which has practical
550
significance for food protein processing. On the other hand, due to the inherent weak Raman
551
scattering effect and strong biological fluorescence interference of Raman spectroscopy,
552
development of new techniques such as surface-enhanced Raman spectroscopy might a promising
553
way to gain a spectrum with strong signals. Considering the use of far-UV CD for protein secondary
554
structure analyses greatly depends on the reference databases, consequently, amelioration of the CD
555
spectrum processing software and computational approaches, as well as establishment of more
556
enormous and accurate reference databases should be considered in the future work. Fluorescence
557
and UV spectrophotometers have relatively low cost with simple operations and future research
558
interests can be focused on software improvements and overlapping spectral processing.
559
It should be pointed out that any spectroscopic technique should not be used alone. In terms of UV
560
and fluorescence spectroscopies, their limitations on the detection of peptide backbone structures
561
should be compensated for other complementary approaches, and combined spectroscopic
562
techniques are promising for increasing prediction accuracy of protein conformations transformation.
563
Apart from the aforementioned techniques, development of other emerging spectroscopies such as
564
using electromagnetic spectra in terahertz (THz) frequency ranges (0.1 to 10 THz) for protein
565
conformation measurement is attractive in future work. Although fundamental research and
566
applications of THz spectroscopy for protein conformational monitoring are still in its infancy, with
567
the development of THz sources and detector, this technique could be a novel and powerful
568
nondestructive technique with great potential for denatured proteins detection. It is hoped that future
569
studies could develop robust solutions for proteins conformational monitoring for the food industry.
AC C
EP
TE D
M AN U
SC
RI PT
547
570 571
Acknowledgments 23
ACCEPTED MANUSCRIPT The authors are grateful to the International S&T Cooperation Program of China (2015DFA71150)
573
for its support. This research was also supported by the Collaborative Innovation Major Special
574
Projects of Guangzhou City (201508020097, 201604020007, 201604020057), the Guangdong
575
Provincial Science and Technology Plan Projects (2015A020209016, 2016A040403040), the Key
576
Projects of Administration of Ocean and Fisheries of Guangdong Province (A201401C04), the
577
National Key Technologies R&D Program (2015BAD19B03), the International and Hong Kong -
578
Macau - Taiwan Collaborative Innovation Platform of Guangdong Province on Intelligent Food
579
Quality Control and Process Technology & Equipment (2015KGJHZ001), the Guangdong Provincial
580
R & D Centre for the Modern Agricultural Industry on Non-destructive Detection and Intensive
581
Processing of Agricultural Products and the Common Technical Innovation Team of Guangdong
582
Province on Preservation and Logistics of Agricultural Products (2016LM2154).
M AN U
SC
RI PT
572
583
References
585
Abaee, A., Madadlou, A., & Saboury, A. A. (2017). The formation of non-heat-treated whey protein cold-set hydrogels
587 588
via non-toxic chemical cross-linking. Food Hydrocolloids, 63, 43-49. Agyare, K. K., Addo, K., & Xiong, Y. L. (2009). Emulsifying and foaming properties of transglutaminase-treated wheat gluten hydrolysate as influenced by pH, temperature and salt. Food Hydrocolloids, 23(1), 72-81.
EP
586
TE D
584
Albani, J. R. (2008). Principles and applications of fluorescence spectroscopy: John Wiley & Sons.
590
Barrios-Peralta, P., Pérez-Won, M., Tabilo-Munizaga, G., & Briones-Labarca, V. (2012). Effect of high pressure on the
591
interactions of myofibrillar proteins from abalone (Haliotis rufencens) containing several food additives. LWT-Food
592
Science and Technology, 49(1), 28-33.
AC C
589
593
Blanpain-Avet, P., Hédoux, A., Guinet, Y., Paccou, L., Petit, J., Six, T., & Delaplace, G. (2012). Analysis by Raman
594
spectroscopy of the conformational structure of whey proteins constituting fouling deposits during the processing in
595
a heat exchanger. Journal of Food Engineering, 110(1), 86-94.
596 597 598 599
Buckow, R., Sikes, A., & Tume, R. (2013). Effect of high pressure on physicochemical properties of meat. Critical Reviews in Food Science and Nutrition, 53(7), 770-786. Carbonaro, M., & Nucara, A. (2010). Secondary structure of food proteins by Fourier transform spectroscopy in the mid-infrared region. Amino Acids, 38(3), 679-690. 24
ACCEPTED MANUSCRIPT 600
Chandrapala, J., Zisu, B., Kentish, S., & Ashokkumar, M. (2012). The effects of high-intensity ultrasound on the
601
structural and functional properties of α-Lactalbumin, β-Lactoglobulin and their mixtures. Food Research
602
International, 48(2), 940-943.
603 604
Chang, C., Li, X., Li, J., Niu, F., Zhang, M., Su, Y., & Yang, Y. (2017). Effect of enzymatic hydrolysis on characteristics and synergistic efficiency of pectin on emulsifying properties of egg white protein. Food Hydrocolloids, 65, 87-95. Chen, H., Diao, J., Li, Y., Chen, Q., & Kong, B. (2016). The effectiveness of clove extracts in the inhibition of hydroxyl
606
radical oxidation-induced structural and rheological changes in porcine myofibrillar protein. Meat Science, 111,
607
60-66.
609
Coleman, J., & Mitchell, C. (2016). The thermal and storage stability of bovine haemoglobin by ultraviolet-visible and
SC
608
RI PT
605
circular dichroism spectroscopies. Journal of Pharmaceutical Analysis, 6(4), 242-248. Damodaran, S. (1997). Food proteins and their applications (Vol. 80). CRC Press.
611
Day, L., Zhai, J., Xu, M., Jones, N. C., Hoffmann, S. V., & Wooster, T. J. (2014). Conformational changes of globular
612
proteins adsorbed at oil-in-water emulsion interfaces examined by synchrotron radiation circular dichroism. Food
613
Hydrocolloids, 34, 78-87.
M AN U
610
Demchenko, A. P. (2013). Ultraviolet spectroscopy of proteins: Springer Science & Business Media.
615
Enomoto, H., Ishimaru, T., Li, C.-P., Hayashi, Y., Matsudomi, N., & Aoki, T. (2010). Phosphorylation of ovalbumin by
616
dry-heating in the presence of pyrophosphate: Effect of carbohydrate chain on the phosphorylation level and heat
617
stability. Food Chemistry, 122(3), 526-532.
TE D
614
Ferrer, E. G., Gómez, A. V., Añón, M. C., & Puppo, M. C. (2011). Structural changes in gluten protein structure after
619
addition of emulsifier. A Raman spectroscopy study. Spectrochimica Acta Part A: Molecular and Biomolecular
620
Spectroscopy, 79(1), 278-28.
EP
618
George, P., Kasapis, S., Bannikova, A., Mantri, N., Palmer, M., Meurer, B., & Lundin, L. (2013). Effect of high
622
hydrostatic pressure on the structural properties and bioactivity of immunoglobulins extracted from whey protein.
623
Food Hydrocolloids, 32(2), 286-293.
624 625 626 627
AC C
621
Gómez, A. V., Ferrer, E. G., Añón, M. C., & Puppo, M. C. (2013). Changes in secondary structure of gluten proteins due to emulsifiers. Journal of Molecular Structure, 1033, 51-58. Guo, X., & Xiong, Y. L. (2013). Characteristics and functional properties of buckwheat protein–sugar Schiff base complexes. LWT-Food Science and Technology, 51(2), 397-404.
628
Han, Y., Wang, J., Li, Y., Hang, Y., Yin, X., & Li, Q. (2015). Circular dichroism and infrared spectroscopic
629
characterization of secondary structure components of protein Z during mashing and boiling processes. Food 25
ACCEPTED MANUSCRIPT 630 631 632 633 634
Chemistry, 188, 201-209. Hawe, A., Sutter, M., & Jiskoot, W. (2008). Extrinsic fluorescent dyes as tools for protein characterization. Pharmaceutical research, 25(7), 1487-1499. He, R., He, H.-Y., Chao, D., Ju, X., & Aluko, R. (2014). Effects of high pressure and heat treatments on physicochemical and gelation properties of rapeseed protein isolate. Food and Bioprocess Technology, 7(5), 1344-1353. Herrero, A. M., Cambero, M., Ordonez, J., De la Hoz, L., & Carmona, P. (2008). Raman spectroscopy study of the
636
structural effect of microbial transglutaminase on meat systems and its relationship with textural characteristics.
637
Food Chemistry, 109(1), 25-32.
639
Hettiarachchy, N. S., Sato, K., Marshall, M. R., & Kannan, A. (2012). Food proteins and peptides: chemistry, functionality, interactions, and commercialization. CRC Press.
SC
638
RI PT
635
Hu, C. q., Chen, H. b., Gao, J. y., Luo, C. p., Ma, X. j., & Tong, P. (2011). High‐pressure microfluidisation‐induced
641
changes in the antigenicity and conformation of allergen Ara h 2 purified from Chinese peanut. Journal of the
642
Science of Food and Agriculture, 91(7), 1304-1309.
643 644
M AN U
640
Hu, H., Cheung, I. W., Pan, S., & Li-Chan, E. C. (2015a). Effect of high intensity ultrasound on physicochemical and functional properties of aggregated soybean β-conglycinin and glycinin. Food Hydrocolloids, 45, 102-110. Hu, H., Wu, J., Li-Chan, E. C., Zhu, L., Zhang, F., Xu, X., Fan, G., Wang, L., Huang, X., & Pan, S. (2013). Effects of
646
ultrasound on structural and physical properties of soy protein isolate (SPI) dispersions. Food Hydrocolloids, 30(2),
647
647-655.
TE D
645
Hu, H., Zhu, X., Hu, T., Cheung, I. W., Pan, S., & Li-Chan, E. C. (2015b). Effect of ultrasound pre-treatment on
649
formation of transglutaminase-catalysed soy protein hydrogel as a riboflavin vehicle for functional foods. Journal of
650
Functional Foods, 19, 182-193.
EP
648
Jenkins, J. E., Sampath, S., Butler, E., Kim, J., Henning, R. W., Holland, G. P., & Yarger, J. L. (2013). Characterizing the
652
secondary protein structure of black widow dragline silk using solid-state NMR and X-ray diffraction.
653
Biomacromolecules, 14(10), 3472-3483.
654 655
AC C
651
Jia, J., Ma, H., Zhao, W., Wang, Z., Tian, W., Lin, L., & He, R. (2010). The use of ultrasound for enzymatic preparation of ACE-inhibitory peptides from wheat germ protein. Food Chemistry, 119(1), 336-342.
656
Jiang, L., Wang, J., Li, Y., Wang, Z., Liang, J., Wang, R., Chen, Y., Ma, W., Qi, B., & Zhang, M. (2014). Effects of
657
ultrasound on the structure and physical properties of black bean protein isolates. Food Research International, 62,
658
595-601.
659
Kaewruang, P., Benjakul, S., & Prodpran, T. (2014). Characteristics and gelling property of phosphorylated gelatin from 26
ACCEPTED MANUSCRIPT
663 664 665 666 667 668 669 670 671 672 673 674
review. Food and Bioprocess Technology, 4(3), 364-386. Kazemi, S., Ngadi, M. O., & Gariépy, C. (2011). Protein denaturation in pork longissimus muscle of different quality groups. Food and Bioprocess Technology, 4(1), 102-106. Kelly, S. M., Jess, T. J., & Price, N. C. (2005). How to study proteins by circular dichroism. Biochimica et Biophysica
RI PT
662
Karoui, R., & Blecker, C. (2011). Fluorescence spectroscopy measurement for quality assessment of food systems-a
Acta (BBA)-Proteins and Proteomics, 1751(2), 119-139.
Kong, J., & Yu, S. (2007). Fourier transform infrared spectroscopic analysis of protein secondary structures. Acta Biochimica et Biophysica Sinica, 39(8), 549-559.
SC
661
the skin of unicorn leatherjacket. Food Chemistry, 146, 591-596.
Kuan, Y. H., Bhat, R., Patras, A., & Karim, A. A. (2013). Radiation processing of food proteins–a review on the recent developments. Trends in Food science & Technology, 30(2), 105-120.
Lam, R. S., & Nickerson, M. T. (2013). Food proteins: a review on their emulsifying properties using a structure–
M AN U
660
function approach. Food Chemistry, 141(2), 975-984.
Li, H., Zhu, K., Zhou, H., & Peng, W. (2012). Effects of high hydrostatic pressure treatment on allergenicity and structural properties of soybean protein isolate for infant formula. Food Chemistry, 132(2), 808-814. Li, K., Kang, Z.-L., Zhao, Y.-Y., Xu, X.-L., & Zhou, G.-H. (2014). Use of high-intensity ultrasound to improve functional
676
properties of batter suspensions prepared from PSE-like chicken breast meat. Food and Bioprocess Technology,
677
7(12), 3466-3477.
TE D
675
Li, Y. Q. (2012). Structure Changes of Soybean Protein Isolates by Pulsed Electric Fields. Physics Procedia, 33, 132-137.
679
Li, Y., Williams, T. D., & Topp, E. M. (2008). Effects of excipients on protein conformation in lyophilized solids by
680
EP
678
hydrogen/deuterium exchange mass spectrometry. Pharmaceutical Research, 25(2), 259-267. Liao, L., Han, X., Chen, L.-p., Ni, L., Liu, Z.-b., Zhang, W., & Chen, Q. (2016). Comparative characterization of the
682
deamidation of carboxylic acid deamidated wheat gluten by altering the processing conditions. Food Chemistry, 210,
683
520-529.
AC C
681
684
Liu, C., Zhao, M., Sun, W., & Ren, J. (2013). Effects of high hydrostatic pressure treatments on haemagglutination
685
activity and structural conformations of phytohemagglutinin from red kidney bean (Phaseolus vulgaris). Food
686
Chemistry, 136(3), 1358-1363.
687 688 689
Liu, R., Zhao, S.-M., Xie, B.-J., & Xiong, S.-B. (2011). Contribution of protein conformation and intermolecular bonds to fish and pork gelation properties. Food Hydrocolloids, 25(5), 898-906. Liu, Y., Zhao, G., Zhao, M., Ren, J., & Yang, B. (2012). Improvement of functional properties of peanut protein isolate 27
ACCEPTED MANUSCRIPT 690
by conjugation with dextran through Maillard reaction. Food Chemistry, 131(3), 901-906.
691
Mao, B., Tejero, R., Baker, D., & Montelione, G. T. (2014). Protein NMR structures refined with Rosetta have higher
692
accuracy relative to corresponding X-ray crystal structures. Journal of the American Chemical Society, 136(5),
693
1893-1906.
695
Martin, S. R., & Schilstra, M. J. (2008). Circular Dichroism and Its Application to the Study of Biomolecules. 84, 263-293.
RI PT
694
Miwa, N., Yokoyama, K., Nio, N., & Sonomoto, K. (2013). Effect of enzymatic deamidation on the heat-induced
697
conformational changes in whey protein isolate and its relation to gel properties. Journal of Agricultural and Food
698
Chemistry, 61(9), 2205-2212.
700 701 702
Morales, R., Martínez, K. D., Ruiz-Henestrosa, V. M. P., & Pilosof, A. M. (2015). Modification of foaming properties of soy protein isolate by high ultrasound intensity: particle size effect. Ultrasonics Sonochemistry, 26, 48-55. Munishkina, L. A., & Fink, A. L. (2007). Fluorescence as a method to reveal structures and membrane-interactions of
M AN U
699
SC
696
amyloidogenic proteins. Biochimica et Biophysica Acta (BBA)-Biomembranes, 1768(8), 1862-1885.
703
Perreault, V., Hénaux, L., Bazinet, L., & Doyen, A. (2017). Pretreatment of flaxseed protein isolate by high hydrostatic
704
pressure: Impacts on protein structure, enzymatic hydrolysis and final hydrolysate antioxidant capacities. Food
705
Chemistry, 221, 1805-1812.
709 710 711 712 713
TE D
708
solid state. LWT-Food Science and Technology, 74, 331-337. Qiu, C., Xia, W., & Jiang, Q. (2014). Pressure-induced changes of silver carp (Hypophthalmichthys molitrix) myofibrillar protein structure. European Food Research and Technology, 238(5), 753-761.
EP
707
Qian, J.-Y., Ma, L.-J., Wang, L.-J., & Jiang, W. (2016). Effect of pulsed electric field on structural properties of protein in
Rahaman, T., Vasiljevic, T., & Ramchandran, L. (2016). Effect of processing on conformational changes of food proteins related to allergenicity. Trends in Food Science & Technology, 49, 24-34.
AC C
706
Raikos, V. (2010). Effect of heat treatment on milk protein functionality at emulsion interfaces. A review. Food Hydrocolloids, 24(4), 259-265.
714
Randall, L., Manta, B., Nelson, K. J., Santos, J., Poole, L. B., & Denicola, A. (2016). Structural changes upon
715
peroxynitrite-mediated nitration of peroxiredoxin 2; nitrated Prx2 resembles its disulfide-oxidized form. Archives of
716
Biochemistry and Biophysics, 590, 101-108.
717 718 719
Rinke, G., Rauschenbach, S., Harnau, L., Albarghash, A., Pauly, M., & Kern, K. (2014). Active conformation control of unfolded proteins by hyperthermal collision with a metal surface. Nano Letters, 14 (10), 5609-5615. Ruffin, E., Schmit, T., Lafitte, G., Dollat, J.-M., & Chambin, O. (2014). The impact of whey protein preheating on the 28
ACCEPTED MANUSCRIPT
722 723 724 725 726 727 728 729 730
Savadkoohi, S., Bannikova, A., Mantri, N., & Kasapis, S. (2016). Structural modification in condensed soy glycinin systems following application of high pressure. Food Hydrocolloids, 53, 115-124. Shen, L., & Tang, C.-H. (2012). Microfluidization as a potential technique to modify surface properties of soy protein isolate. Food Research International, 48(1), 108-118. Sheng, L., Wang, J., Huang, M., Xu, Q., & Ma, M. (2016). The changes of secondary structures and properties of
RI PT
721
properties of emulsion gel bead. Food Chemistry, 151, 324-332.
lysozyme along with the egg storage. International Journal of Biological Macromolecules, 92, 600-606.
Shilpashree, B., Arora, S., Chawla, P., & Tomar, S. (2015). Effect of succinylation on physicochemical and functional properties of milk protein concentrate. Food Research International, 72, 223-230.
SC
720
Song, C.-L., & Zhao, X.-H. (2014). Structure and property modification of an oligochitosan-glycosylated and crosslinked soybean protein generated by microbial transglutaminase. Food Chemistry, 163, 114-119. Spotti, M. J., Martinez, M. J., Pilosof, A. M., Candioti, M., Rubiolo, A. C., & Carrara, C. R. (2014). Influence of Maillard
732
conjugation on structural characteristics and rheological properties of whey protein/dextran systems. Food
733
Hydrocolloids, 39, 223-230.
M AN U
731
Stănciuc, N., Aprodu, I., Râpeanu, G., van der Plancken, I., Bahrim, G., & Hendrickx, M. (2013). Analysis of the
735
thermally induced structural changes of bovine lactoferrin. Journal of Agricultural and Food Chemistry, 61(9),
736
2234-2243.
TE D
734
Tabilo-Munizaga, G., Gordon, T. A., Villalobos-Carvajal, R., Moreno-Osorio, L., Salazar, F. N., Pérez-Won, M., & Acuña,
738
S. (2014). Effects of high hydrostatic pressure (HHP) on the protein structure and thermal stability of Sauvignon
739
blanc wine. Food Chemistry, 155, 214-220.
EP
737
Tong, P., Gao, J., Chen, H., Li, X., Zhang, Y., Jian, S., Wichers, H., Wu, Z., Yang, A., & Liu, F. (2012). Effect of heat
741
treatment on the potential allergenicity and conformational structure of egg allergen ovotransferrin. Food Chemistry,
742
131(2), 603-610.
AC C
740
743
Ustunol, Z. (Ed.). (2014). Applied food protein chemistry. John Wiley & Sons.
744
Wallace, B. A., & Janes, R. W. (2010). Synchrotron radiation circular dichroism (SRCD) spectroscopy: an enhanced
745
method for examining protein conformations and protein interactions. Biochemical Society Transactions, 38(4),
746
861-873.
747 748 749
Wang, K. Q., Luo, S. Z., Zhong, X. Y., Cai, K. Z., Cai, J., Jiang, S. T., & Zheng, Z. (2016a). Effect of Modified Wheat Gluten on Boiling Resistance Capacity of Pork Meatballs. Journal of Food Science, 81(2), E430-E437. Wang, K., Li, C., Wang, B., Yang, W., Luo, S., Zhao, Y., Jiang, S., Mu, D., & Zheng, Z. (2017a). Formation of 29
ACCEPTED MANUSCRIPT 750
macromolecules in wheat gluten/starch mixtures during twin‐screw extrusion: effect of different additives. Journal
751
of the Science of Food and Agriculture, DOI: 10.1002/jsfa.8392.
753 754 755 756 757
Wang, K., Luo, S., Cai, J., Sun, Q., Zhao, Y., Zhong, X., Jiang, S., & Zheng, Z. (2016b). Effects of partial hydrolysis and subsequent cross-linking on wheat gluten physicochemical properties and structure. Food Chemistry, 197, 168-174. Wang, K.-Q., Luo, S.-Z., Zhong, X.-Y., Cai, J., Jiang, S.-T., & Zheng, Z. (2017b). Changes in chemical interactions and protein conformation during heat-induced wheat gluten gel formation. Food Chemistry, 214, 393-399.
RI PT
752
Wang, W.-q., Bao, Y.-h., & Chen, Y. (2013). Characteristics and antioxidant activity of water-soluble Maillard reaction products from interactions in a whey protein isolate and sugars system. Food Chemistry, 139(1), 355-361. Whitford, D. (2013). Proteins: structure and function. John Wiley & Sons.
759
Whitmore, L., & Wallace, B. A. (2008). Protein secondary structure analyses from circular dichroism spectroscopy:
761 762
methods and reference databases. Biopolymers, 89(5), 392-400.
M AN U
760
SC
758
Wu, W., Zhang, C., Kong, X., & Hua, Y. (2009). Oxidative modification of soy protein by peroxyl radicals. Food Chemistry, 116(1), 295-301.
763
Xiong, G., Han, M., Kang, Z., Zhao, Y., Xu, X., & Zhu, Y. (2016). Evaluation of protein structural changes and water
764
mobility in chicken liver paste batters prepared with plant oil substituting pork back-fat combined with
765
pre-emulsification. Food Chemistry, 196, 388-395.
769 770 771
TE D
768
acidification with d-gluconic acid δ-lactone. Food Chemistry, 134(2), 1005-1010. Yan, W., Xu, B., Jia, F., Dai, R., & Li, X. (2016). The Effect of High-Pressure Carbon Dioxide on the Skeletal Muscle Myoglobin. Food and Bioprocess Technology, 1-8.
EP
767
Xu, Y., Xia, W., & Jiang, Q. (2012). Aggregation and structural changes of silver carp actomyosin as affected by mild
Yang, M., Cui, N., Fang, Y., Shi, Y., Yang, J., & Wang, J. (2015). Influence of succinylation on the conformation of yak casein micelles. Food Chemistry, 179, 246-252.
AC C
766
772
Yin, S. W., Tang, C. H., Wen, Q. B., Yang, X. Q., & Li, L. (2008). Functional properties and in vitro trypsin digestibility
773
of red kidney bean (Phaseolus vulgaris L.) protein isolate: effect of high-pressure treatment. Food Chemistry, 110(4),
774
938-945.
775
Yin, S.-W., Tang, C.-H., Wen, Q.-B., Yang, X.-Q., & Yuan, D.-B. (2010). The relationships between physicochemical
776
properties and conformational features of succinylated and acetylated kidney bean (Phaseolusvulgaris L.) protein
777
isolates. Food Research International, 43(3), 730-738.
778
Zhang, B., Chi, Y. J., & Li, B. (2014). Effect of ultrasound treatment on the wet heating Maillard reaction between
779
β-conglycinin and maltodextrin and on the emulsifying properties of conjugates. European Food Research and 30
ACCEPTED MANUSCRIPT 780 781 782
Technology, 238(1), 129-138. Zhang, W., Waghmare, P. R., Chen, L., Xu, Z., & Mitra, S. K. (2015). Interfacial rheological and wetting properties of deamidated barley proteins. Food Hydrocolloids, 43, 400-409. Zhang, X., Qi, J.-R., Li, K.-K., Yin, S.-W., Wang, J.-M., Zhu, J.-H., & Yang, X.-Q. (2012). Characterization of soy
784
β-conglycinin–dextran conjugate prepared by Maillard reaction in crowded liquid system. Food Research
785
International, 49(2), 648-654.
788 789 790 791 792 793
Agricultural and Food Chemistry, 60(30), 7526-7531.
Zhang, Z., Yang, Y., Zhou, P., Zhang, X., & Wang, J. (2017). Effects of high pressure modification on conformation and
SC
787
Zhang, Y., & Zhong, Q. (2012). Effects of thermal denaturation on binding between bixin and whey protein. Journal of
gelation properties of myofibrillar protein. Food Chemistry, 217, 678-686.
Zhao, L., Li, L., Liu, G. Q., Liu, X. X., & Li, B. (2012). Effect of frozen storage on molecular weight, size distribution and conformation of gluten by SAXS and SEC-MALLS. Molecules, 17(6), 7169-7182.
M AN U
786
RI PT
783
Zhao, W., & Yang, R. (2008). The effect of pulsed electric fields on the inactivation and structure of lysozyme. Food Chemistry, 110(2), 334-343.
Zhong, J., Liu, W., Liu, C., Wang, Q., Li, T., Tu, Z., Luo, S., Cai, X., & Xu, Y. (2012). Aggregation and conformational
795
changes of bovine β-lactoglobulin subjected to dynamic high-pressure microfluidization in relation to antigenicity.
796
Journal of Dairy Science, 95(8), 4237-4245.
TE D
794
Zhou, M., Liu, J., Zhou, Y., Huang, X., Liu, F., Pan, S., & Hu, H. (2016). Effect of high intensity ultrasound on
798
physicochemical and functional properties of soybean glycinin at different ionic strengths. Innovative Food Science
799
& Emerging Technologies, 34, 205-213.
EP
797
Zhuo, X. Y., Qi, J. R., Yin, S. W., Yang, X. Q., Zhu, J. H., & Huang, L. X. (2013). Formation of soy protein isolate–
801
dextran conjugates by moderate Maillard reaction in macromolecular crowding conditions. Journal of the Science of
802
Food and Agriculture, 93(2), 316-32.
AC C
800
31
ACCEPTED MANUSCRIPT
Table 1. Recent advances in application of spectroscopy techniques for monitoring proteins conformational changes during modification. Processing methods
Spectroscopic techniques
References
α-Lactalbumin and β-Lactoglobulin
High intensity ultrasound
CD
Chandrapala et al. (2012)
β‑conglycinin
Ultrasound + Maillard reaction
CD
Zhang et al. (2014)
Barley proteins
Deamidation
FTIR
Beer protein Z
Mashing, boiling
CD, FTIR
Black bean protein isolates
Ultrasound
CD, fluorescence
Bovine β-Lactoglobulin
High pressure microfluidization
Fluorescence, UV, CD
Bovine haemoglobin
Heat
UV, CD
Bovine lactoferrin
Heat
Fluorescence
Buckwheat protein
Maillard reaction
Chicken breast meat
High intensity ultrasound
Egg allergen ovotransferrin
Heat
Egg white protein
Pulsed electric field
Flaxseed protein isolate
High hydrostatic pressure
Gelatin
Phosphorylation
Immunoglobulins
High hydrostatic pressure
Milk protein concentrate
Acylation
Myofibrillar protein
High pressure
Myofibrillar protein
High pressure
Myofibrillar protein
High pressure
Peanut protein isolate
Maillard reaction
Phytohemagglutinin
RI PT
Samples
M AN U
SC
Zhang et al. (2015) Han et al. (2015) Jiang et al. (2014) Zhong et al. (2012) Coleman & Mitchell (2016) Stănciuc et al. (2013) Guo & Xiong (2013)
Raman
Li et al. (2014)
Fluorescence, CD, UV
Tong et al. (2012)
UV, FTIR
Qian et al. (2016)
Fluorescence
Perreault et al. (2017)
FTIR
Kaewruang et al. (2014)
FTIR
George et al. (2013)
Fluorescence
Shilpashree et al. (2015)
UV
Barrios-Peralta et al. (2012)
Fluorescence, UV, Raman, CD
Qiu et al. (2014)
Raman
Zhang et al. (2017)
Fluorescence, CD
Liu et al. (2012)
High hydrostatic pressure
FTIR
Liu et al. (2013)
Rapeseed protein isolate
High pressure, heat
CD
He et al. (2014)
Sauvignon blanc wine proteins
High hydrostatic pressure
FTIR
Tabilo-Munizaga et al. (2014)
Silver carp actomyosin
Acidification
CD, UV, fluorescence
Xu et al. (2012)
Skeletal muscle myoglobin
High pressure carbon dioxide
UV, CD, fluorescence
Yan et al. (2016)
Soy glycinin
High pressure
FTIR
Savadkoohi et al. (2016)
EP
TE D
UV, fluorescence
AC C
803
32
ACCEPTED MANUSCRIPT
FTIR, CD, Raman
Hu et al. (2015a)
Soybean glycinin
High intensity ultrasound
CD, fluorescence
Zhou et al. (2016)
Soybean protein
Glycosylation, enzymatic cross-linking
CD
Song & Zhao (2014)
Soybean protein isolate
High hydrostatic pressure
Fluorescence, CD
Li et al. (2012)
Soy protein
Ultrasound + enzymatic cross-linking
Raman
Soy protein isolate
Microfluidization
Fluorescence
Soy protein isolate
Low-frequency ultrasonication
CD
Soy protein isolate
Maillard reaction
Fluorescence, CD
Wheat gluten
Deamidation
FTIR, fluorescence
Liao et al. (2016)
Wheat gluten
Extrusion
FTIR
Wang et al. (2017a)
Wheat gluten
Heat
UV, fluorescence, FTIR
Wang et al. (2017b)
Whey protein
Heat
Raman
Blanpain-Avet et al. (2012)
Whey protein
Heat
Fluorescence
Ruffin et al. (2014)
Whey protein
Heat
Fluorescence, CD, FTIR
Zhang and Zhong (2012)
Whey protein
Maillard reaction
Fluorescence
Spotti et al. (2014)
Whey protein isolate
Maillard reaction
FTIR, CD
Wang et al. (2013)
Whey protein isolate
Enzymatic deamidation
CD, fluorescence
Miwa et al. (2013)
Yak casein micelles
Acylation
Fluorescence, FTIR
Yang et al. (2015)
RI PT
High intensity ultrasound
Hu et al. (2015a)
M AN U
SC
Hu et al. (2013) Zhuo et al. (2013)
EP
TE D
Shen & Tang (2012)
AC C
804
Soybean β-conglycinin and glycinin
33
ACCEPTED MANUSCRIPT
Table 2. Comparison of different spectroscopic techniques for detecting proteins conformational changes Protein structure
Protein state
Advantages
Limitations
FTIR
Secondary structure
Liquid or solid
Fast and Convenient Sensitive to conformational changes under various conditions Lack of dependence on the physical state of samples
High cost Strong IR absorbance of H2O H-D substitution affect protein properties
Raman
Secondary and tertiary structures
Liquid or solid
Non-destructive Convenient On-line and in situ Weaker H2O interference
High cost Inherently weaker Raman scattering Stronger biological fluorescence interference Thermal effect generated by the laser
CD
Secondary and tertiary structures
Liquid
Fast Low protein concentration
Fluorescence
Tertiary structure
Liquid
Economic and simple The data acquisition is quite fast Useful for in-depth thermodynamic studies
UV
Tertiary structure
Liquid
Fast Economic and simple
Time-consuming sample preparation procedures Samples should be highly clear Accurate sample concentrations and reference databases being essential for determining secondary structure content Not suitable for direct measurements of solid-state and high concentration samples Time-consuming sample preparation procedures Not suitable for direct measurements of samples in solid-state Unable to determine secondary structures Time-consuming sample preparation procedures Not suitable for direct measurements of samples in solid-state Unable to determine secondary structures Overlapping Tyr, Trp, and Phe spectra
EP
TE D
M AN U
SC
RI PT
Spectroscopies
AC C
805
34
ACCEPTED MANUSCRIPT 806 807 808
Figure captions Figure 1. Handling process of effect of pulsed electric field on IR spectra in amide I region of egg white protein powder; (a)-(f) represent peak-fitting of the secondary derived curves from IR
810
spectra for samples treated at 0, 5, 10, 15, 20, and 25 kV, respectively. Six peaks are observed
811
for four components, which are attributed to α-helices (1657 cm-1), β-sheets (1611 and 1626
812
cm-1), β-turns (1673 and 1688 cm-1) and random coils (1642 cm-1) (Qian et al., 2016).
RI PT
809
Figure 2. Raman spectra of freeze-dried high intensity ultrasound (20 kHz at 400 W) treated
814
β-conglycinin at 0, 5, 20 and 40 min (Hu et al., 2015a). 500-550 cm-1: S-S stretching
815
vibrational bands; 760, 880 and 1361 cm-1: Trp vibrational bands; 830 and 850 cm-1: Tyr
816
vibrational bands; 1200-1340 cm-1: amide III bands; 1600-1700 cm-1: amide I band; 1440-1465
817
cm-1: C-H bending; 2800-3000 cm-1: C-H stretching.
M AN U
SC
813
Figure 3. (a) Far-UV CD spectra and (b) near-UV CD spectra of myoglobin in solution at pH 7,
819
absorbed at tricaprin oil/water interface and at hexadecane oil/water interface (Day et al.,
820
2014).
823 824
different ultrasonic treatment (Jiang et al., 2014). Figure 5. (a) the zero-order and (b) second-derivative UV spectra of lysozyme before and after
EP
822
Figure 4. Intrinsic fluorescence emission spectra for 0.15 mg/mL black bean protein dispersion of
pulsed electric field (PEF) treated at 35 kV/cm for 1200µs (Zhao & Yang, 2008).
AC C
821
TE D
818
35
ACCEPTED MANUSCRIPT Figure 1
AC C
EP
TE D
M AN U
SC
RI PT
825
826
36
ACCEPTED MANUSCRIPT Figure 2
M AN U
SC
RI PT
827
AC C
EP
TE D
828
37
ACCEPTED MANUSCRIPT Figure 3
SC
RI PT
829
AC C
EP
TE D
M AN U
830
38
ACCEPTED MANUSCRIPT Figure 4
SC
RI PT
831
AC C
EP
TE D
M AN U
832
39
ACCEPTED MANUSCRIPT Figure 5
M AN U
SC
RI PT
833
AC C
EP
TE D
834
40
ACCEPTED MANUSCRIPT Highlights
Protein modification is essential for obtaining optimal functionalities. Protein functionalities are closely related to its conformational characteristics.
RI PT
Spectroscopies for evaluating protein conformational changes are reviewed. Advantages and limitations of each spectroscopic technique are discussed
AC C
EP
TE D
M AN U
SC
Spectroscopic techniques should be used complementary.