Macromolecular Interactions in Food Technology - American Chemical

Chapter 8. Laser-Light-Scattering Properties of Heat-Induced Ovalbumin Gels. Yoshinori Mine. Department of Food Science, University of Guelph, Guelph,...
0 downloads 0 Views 1MB Size
Chapter 8

Laser-Light-Scattering Properties of Heat-Induced Ovalbumin Gels

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

Yoshinori Mine Department of Food Science, University of Guelph, Guelph, Ontario N1G 2W1, Canada

Multi-angle laser photometer and dynamic light scattering spectrophotometer were used for characterizing the aggregation behaviour and gel network structure of elastic gels and less elastic ones of ovalbumin. Ovalbumin formed opaque and less elastic gels in 50 mM phosphate buffer (pH 7.0). However, succinylated ovalbumin formed transparent and elastic gels. Ovalbumin protein formed high molecular and polydisperse aggregates upon heating. Succinylation of the protein contributed to formation of lower molecular weight polymer with a narrow molecular size distribution during subsequent heat gelation. The root mean square and hydrodynamic radii of the protein polymers indicated that the elastic gel consisted of a more extended protein polymer network compared to that of the less elastic gel. The aggregation behaviour of denatured ovalbumin appeared to play a crucial role in the subsequent gelling process of the protein.

Elucidation of the protein-structure function relationship in food proteins can result in useful practical information in food and protein chemistry. Egg white proteins display multiple functional properties such as foaming and gelling properties. Although numerous methods exist to study protein structure, only limited information is available concerning the relationship between protein structure and functionalities of food proteins. In general, a protein gel network is formed via noncovalent crosslinkages such as hydrophobic interactions, or electrostatic and hydrogen bonds interactions and less frequently, by covalent interactions such as disulfide bonds to form a three-dimensional network (/). It is believed that the gelling properties of proteins could be affected by factors determining the state of aggregates found during the subsequent heat for gelation. The mechanism, however, underlying the formation of the three-dimensional network of protein gels is not fully understood. Ovalbumin forms heat-induced opaque gels, whereas transparent gels are formed by succinylation of the amino groups of the protein (2). Transparent gels obtained by two-step heating method were firm and elastic with high-water holding capacity, while the opaque gel obtained by one-step heating was soft and less elastic (5). Absolute characterization of these protein aggregates is neccessary to understand these phenomena. There is

0097-6156/96/0650-0104$15.00/0 © 1996 American Chemical Society

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

8. MINE

Laser-Light-Scattering Properties of Ovalbumin Gels 105

little information regarding the size of aggregated protein polymers until the recentstudies of Kato et at. (4). Thus, it is of great interest to elucidate the relationship between the heat-induced gelling properties of ovalbumin and the aggregation behaviour of these proteins. The development of laser light scattering techniques have enabled an advance in polymer science for characterizing the aggregation behaviour and networkstructure of polymer gels. Light scattering is an ideal means to determine the actual molecular weight and aggregate formation (5). In this study, the author demonstrate the technique for characterizing protein aggregation behaviour and gel structure upon heating in situ. Multi-angle laser scattering coupled with size exclusion HPLC and dynamic light spectrophotometry allows us to obtain not only an absolute molecular weight and the root mean square (RMS) radius of protein polymers but also the hydrodynamic radius of the protein gel network in situ. Materials and Methods Materials Ovalbumin was purchased from Sigma Chemicals, St. Louis, MO. Other chemicals used were ACS grade from Sigma. Ovalbumin was succinylated according to the procedure described by Groninger (6). Succinic anhydride was added at a 50:1 protein : anhydride weight ratio. The extent of succinylation was determined from the free amino contents by the method of Concon (7) using dinitrobenzene sulfonate (DNBS). Preparation of Gels A protein solution (8.0%) in 50mM phosphate buffer (pH 7.0) was injected into an aluminum cup (5mm inner diameter, 2.5mm height). The cup was filled, completely degassed, and sealed with a silicone sheet on which a steel plate wastightlyfixed with clips. The cups containing the protein solutions were heated at 80° C for 20 min., followed by rapid cooling to room temperature by immersion in tap water. Creep Analysis of Gels Creep behaviour under compression was analyzed with a Rheoner RE-3305 (Yamaden, Co Ltd., Tokyo) interfaced with a computer (NEC PC 9801 VM, NEC, Tokyo). The creep curves were analyzed according to the procedure described by Kamata et al. (8). Light Scattering Measurements Light scattering was measured on a DAWN DSP-F multi-angle laser light scattering photometer (MALLS) (Wyatt technology, Santa Barbara, CA) using a 632.8 nm laser and dynamic light scattering spectrophotometer (DLS-7000, Otsuka Electronics, Tokyo). The MALLS was coupled with HPLC system consisting of a Shodex KW-804 and KW-803 columns at a room temperature. A specificrefractionindex increment (dn/dc) of 0.186 was obtained for the dialyzed protein solution using a Wyatt OPTILAB DSP interferometric differential refractometer. Equations The MALLS photometer measures the intensity of the scattered light with the use of photodiodes placed at specific angles (6)relativeto the laser beam. This light intensity is converted to the Rayleighratioas follows: Rayleigh Ratio: R =I r /(I V) (eq.l) Where: R = Rayleigh ratio I = Scattered intensity 2

e

e

0

e

e

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

106

MACROMOLECULAR INTERACTIONS IN FOOD TECHNOLOGY

I = Incident beam intensity V = Scattering volume r = distance between scattering volume and detector From the Rayleigh ratio, we can use the Zimm equation to directly determine molecular weight provided that a number of physical properties are known. Zimm equation: R /K*c=MP(9)+2A cM P (0) (eq.2) Where: R^ = the Rayleigh ratio K* = an optical constant, 4fr nfi(dn/dcfX ~ N \ where n is the refractive index of the solvent at the incident radiation wavelength (nm), N is Avogadro's number and dn/dc is the differentialrefractiveindex increment solvent-solute solution with respect to a change in solute concentration. c = the concentration of solute molecules in the solvent P(9) = the theoretically derived form factor. P(0) is a function of the molecule's z-average size, shape, and structure. Aj = the second virial coefficient M = the weight average molecular weight Since we are using multi-angle detection here, we can extrapolate the value of q to 0. In the limit of 9- >0, P(0)->1 and equation (2) becomes: ReVK'c = M+2A cM (eq.3) If A2=0, then M=R /K*c (eq.4) Using the Debye Plot: Re/K*c vs. Sin (9/2) (eq.5) We can see that at an angle of 9=0, we can read the molecular weight (M) directly off of the y-axis of the plot. It depends on the RMS radius independent of molecular conformation (5). 0

2

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

e

2

2

2

4

0

A

0

A

2

2

e

2

2

Using a continuum model (9) that the wavelength of light is usually much larger than the average distance between neighboring cross-links, self-correlation function is derived as the following linear equation: 2

C(x)~a(kT/K)exp(-Dq T) where: C(x)= self correlation function k= Boltzmann constant T= absolute temperature K= modulus of elasticity D= diffusion coefficient We produce the Stokes-Einstein relation for the diffusion coefficient of polymers in solution. D=kT/67rr| a where: T|= liquid viscosity a= hydrodynamic radius of the polymer 0

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

Laser-Light-Scattering Properties of Ovalbumin Gels 107

8. MINE

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

Results and Discussion Creep-Analysis of Ovalbumin Gels By using succinic anhydride, 28% of the lysine of ovalbumin was modified. The gel formation of the succinylated ovalbumin (Su-ovalbumin) protein was compared to that of ovalbumin. Ovalbumin formed an opaque gel at pH 7.0 upon heating at 80° C for 20 min. On the other hand, the Suovalbumin gel gave a transparent gel under the same condition (The inset of Fig. 1). The typical creep and creep-recovery curves of these two gels were obtained to clarify the difference in the gel characteristics of a transparent gel and an opaque one. The creep compliance of the transparent gel was more than twice that of the opaque gel, indicating that the transparent gel was more soft and more deformable. Differences in the rheological properties of these gels were also observed clearly by the residual strain after the removal of the stress. The Su-ovalbumin gels showed lower residual strain than ovalbumin gels. These results suggested that the Su-ovalbumin gels were more elastic, while those of ovalbumin were likely to be more plastic or less elastic. Creep curves obtained were further analyzed with a four-element mechanical model as shown in Table 1. Each parametenP, stress, E , instantaneous modulus, E retarded elastic modulus: r\ , retarded viscous modulus, and r| , Newtonian viscous modulus was calculated. The P and E value of ovalbumin gels was over twice than that of Suovalbumin gels, while E r| and rj values were smaller in the opaque gels. These results occured the transparent gels have a small instaneous deformation but a larger retarded deformation with time up to 300 sec. than the opaque gels. The results suggested that the Su-ovalbumin gels were more elastic, while the opaque gels were more viscous. 0

x

p

n

0

p

x

n

Laser Light Scattering It is well known that the gelling properties may be affected by the cross Unking of unfolded molecules as a result of hydrogen bonding and ionic and hydrophobic interactions (10). The turbidity of the ovalbumin gels depends on the pH and ionic strength of the heated protein solution (11). In the past, low-angle laser light scattering was used to estimate the molecular weight distribution of heatinduced ovalbumin and dry-heated egg white protein aggregates (4). However, this system has been hampered severely by solvent noise due to impurities. The MALLS system dramatically reduces the effect of noisy background, resulting in cleaner data and highly reproducible results (5). The molecular weight and RMS of heat-induced soluble aggregates were analyzed by the MALLS system. Fig. 2 shows the relationship between retention volume (HPLC) and molecular weight of heat-induced protein aggregates. Ovalbumin aggregates have molecular weights between 1.7 x 10 to 1.6 x 10 , while the range for Su-ovalbumin was lower (2.1 x 10 to 4.2 x 10 ). In general, the retention volume by HPLC gel filtration reflects the size of and conformation of proteins eluted at a particular retention volume. Figure 3 shows the relationship between the RMS radius and the retention volume (HPLC) of the heatinduced aggregates of the proteins. Interestingly, there was little difference in the RMS radius for both the aggregates except large retention volume values in spite of the considerable differences of absolute molecular weight. Table 2 summarizes the laser light scattering characteristics of heat-induced ovalbumin and Su-ovalbumin aggregates. The data indicated that the molecular weight distribution of the heat-induced ovalbumin aggregates was large and heterogeneous. By contrast, Su-ovalbumin was smaller and less heterogeneous. Interestingly, no significant differences on the RMS of the aggregates between ovalbumin and Su-ovalbumin were shown. These results suggest that succinylation of ovalbumin contributes to the formation of lower molecular weight and narrower molecular size distribution of the aggregates upon heating. The formation of extended structures of heat-induced Su-ovalbumin polymer was also predicted from the RMS 6

8

5

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

7

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

108

MACROMOLECULAR INTERACTIONS IN FOOD TECHNOLOGY

Figure 1. Creep and creep-recovery curves of (a) ovalbumin and (b) Suovalbumin gels.

Table 1. Viscoelastic Parameters of Ovalbumin and Su-ovalbumin Gels parameters 4

P(xl0 ) E (xl0 ) E^xlO ) ^(xlO ) n (*io ) 5

D

6

7

8

N

*1

ovalbumin

Su-ovalbumin 2.50 1.52 2.05 3.31 6.08 16.13

4.99 3.57 1.46 4.86 4.00 33.19

P, stress; E , instantaneous modulus; E retarded elastic modulus; rjp retarded viscous modulus; t| , Newtonian modulus 0

p

N

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

MINE

Laser-Light-Scattering Properties of Ovalbumin Gels

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

1.0E+09

1.0E+04 20.0

12.0 Volume (mL)

Figure 2. Relationship between absolute molecular weight and retention volume (HPLC) of (a) ovalbumin and (b) Su-ovalbumin soluble aggregates.

12.0

16.0

20.0

Volume (mL)

Figure 3. Relationship between RMS radius and retention volume (HPLC) of (a) ovalbumin and (b) Su-ovalbumin soluble aggregates.

Parris et al.; Macromolecular Interactions in Food Technology ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

110

MACROMOLECULAR INTERACTIONS IN FOOD TECHNOLOGY

Table 2. Laser Light Scattering Characteristics of Heat-Induced Ovalbumin and Su-ovalbumin Aggregates

Downloaded by UNIV OF MASSACHUSETTS AMHERST on May 31, 2018 | https://pubs.acs.org Publication Date: November 19, 1996 | doi: 10.1021/bk-1996-0650.ch008

8

d

samples

Mw(xl(ff

Mn (xl&f

Mw/Mn

RMS (w)

ovalbumin

322.60

9.24

34.91

25.2

59.35

10.01

5.93

23.9

Su-ovalbumin a

0.1% protein in 50mM phosphate buffer (pH 7.0) was heated at 80° C for 20 min. Weight-average molecular weight (g/mol). Number-average molecular weight (g/mol). RMS radius (nm).

b c

d

10