ArticlePDF Available

Sustainable and practical utilization of feather keratin by an innovative physicochemical pretreatment: High density steam flash-explosion

Authors:

Abstract and Figures

Currently, great attention is being paid to the utilization of biomass, such as feather keratins. It is imperative to extract and dissolve keratins from animal keratinous materials for exploitation of innovative biopolymers. However, most of the current processes are based on strong acid and alkali hydrolysis, chemical cleavage and other violent reactions, which are not eco-friendly and/or result in severe degradation and destruction of feather keratins. In this study, high density steam flash-explosion (HDSF) as an innovative pretreatment of biomass was firstly employed to treat feather waste. In HDSF treatment, steam with a powerful seepage force first penetrates into fibrous tissues and cells of feathers, and then quickly expands and breaks free of the structure upon an explosive decompression at supersonic speed (within 0.0875 s). HDSF effectively destabilized β-sheet crystals and intermolecular disulfide bonds without causing substantial damage to the keratin protein chain, dramatically increasing the extraction and dissolubility of feather keratins in polar solvents like water, salt solution and weak bases, as well as enzymatic accessibility. HDSF treatment could be a sustainable and practical pretreatment for extraction of feather keratin for exploitation of biomaterials and conversion of feathers to nutrient animal feed instead of the current chemical hydrolysis and hydrothermal treatment.
Content may be subject to copyright.
Green Chemistry Dynamic Article Links
Cite this: Green Chem., 2012, 14, 3352
www.rsc.org/greenchem PAPER
Sustainable and practical utilization of feather keratin by an innovative
physicochemical pretreatment: high density steam ash-explosion
Wei Zhao,* Ruijin Yang,* Yiqi Zhang and Li Wu
Received 5th August 2012, Accepted 27th September 2012
DOI: 10.1039/c2gc36243k
Currently, great attention is being paid to the utilization of biomass, such as feather keratins. It is
imperative to extract and dissolve keratins from animal keratinous materials for exploitation of innovative
biopolymers. However, most of the current processes are based on strong acid and alkali hydrolysis,
chemical cleavage and other violent reactions, which are not eco-friendly and/or result in severe
degradation and destruction of feather keratins. In this study, high density steam ash-explosion (HDSF)
as an innovative pretreatment of biomass was rstly employed to treat feather waste. In HDSF treatment,
steam with a powerful seepage force rst penetrates into brous tissues and cells of feathers, and then
quickly expands and breaks free of the structure upon an explosive decompression at supersonic speed
(within 0.0875 s). HDSF effectively destabilized β-sheet crystals and intermolecular disulde bonds
without causing substantial damage to the keratin protein chain, dramatically increasing the extraction and
dissolubility of feather keratins in polar solvents like water, salt solution and weak bases, as well as
enzymatic accessibility. HDSF treatment could be a sustainable and practical pretreatment for extraction
of feather keratin for exploitation of biomaterials and conversion of feathers to nutrient animal feed
instead of the current chemical hydrolysis and hydrothermal treatment.
1. Introduction
Poultry feathers are waste products generated in large quantities
from commercial poultry processing and down product manufac-
ture, which have become a crucial industrial waste over the
world. An estimated 1.52 million tonnes are produced annually
as a waste from the poultry industry in the United States and
over 65 million tones are produced worldwide.
1,2
Poultry feath-
ers contain about 90% protein (keratin) and are a cheap and
renewable source for protein bers. However, feather keratin is
tightly packed in β-sheets into a supercoiled polypeptide chain
with a high degree of disulde cross-linkages, hydrophobic inter-
actions, and hydrogen bonds. The peculiar structure of keratin
confers indissolubility, mechanical stability and resistance to
common proteolytic enzymes and chemicals to feathers.
3
Thus,
most feathers are currently disposed of as waste by incineration
or landlling.
4
Currently, greater attention has been paid to the utilization of
biomass.
5,6
In order to effectively exploit biomass, a pretreat-
ment is generally required to destroy the structure preventing
degradation of biomass or to release the constitutive com-
ponents, thereby increasing enzyme and solvent accessibility.
6
If preferred processing techniques can be found, the insoluble
and hard-to-degrade animal keratin could be converted into
biomaterials with innovative properties, such as highly efcient
absorbent materials,
7
bioplastics and biomedical materials for
tissue engineering and regenerative medicine.
8
However, keratins
are stabilized by a large amount of intra- and intermolecular
disulde cross-links plus other protein structural features, like
crystallinity and hydrogen-bonding. It gives keratin high strength
and stiffness, and insolubility in polar solvents, like water, weak
acids and bases, as well as in apolar solvents. Therefore, it has
become imperative to extract and dissolve keratins from animal
keratinous materials for the exploitation of innovative biopoly-
mers. Advances in the extraction and dissolution of keratins
(Table 1), led to the exponential growth of keratin materials and
their derivatives.
The chemical methods (Table 1) mostly employ strong acid
and alkali hydrolysis and chemical cleavage of the disulde
bonds to extract soluble keratins. However, these chemicals used
in chemical methods, such as sultes, thiols, DTT, or peroxides,
are harmful, often toxic, and difcult to handle. The physico-
chemical methods, such as hydrothermal, superheated water
treatments, are mostly based on violent reaction conditions.
Although conventional steam explosion and extrusion treatments
are feasible for pretreatment of lignocellulosic biomass, reduc-
tants such as sodium sulte are required to synergistically
combine with these physicochemical methods for pretreatment
of keratinous materials because keratins are more robust than
carbohydrates.
21
These treatments always result in severe degra-
dation of keratin with reduction of molecular weight and loss of
mechanical properties.
21
In addition, enzymatic hydrolysis and
microbial fermentation are recently receiving considerable
State Key Laboratory of Food Science & Technology and School of
Food Science and Technology, Jiangnan University, No. 1800 Lihu
Road, Wuxi 214122, China. E-mail: zhaow@jiangnan.edu.cn,
yrj@jiangnan.edu.cn; Fax: +86 510 85919150
3352 |Green Chem., 2012, 14, 33523360 This journal is ©The Royal Society of Chemistry 2012
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
/ Journal Homepage
/ Table of Contents for this issue
attention.
23
However, the higher cost of enzymes themselves,
with a long production cycle, has so far limited the development
of industrial processes. Moreover, most keratins are converted
into much shorter oligopeptides and amino acids. Thus, research-
ers have focused on nding simple and eco-friendly processing
methods to extract and dissolve feather keratins.
High density steam ash-explosion (HDSF) is an innovative
method for pretreatment of biomass, which could be performed
on a large scale. This method is based on pressurization and
forcing steam into brous tissues and cells of biomass, followed
by rapidly releasing the pressure in an explosive decompression
event. This process functions as an adiabatic expansion process
and a conversion process of thermal energy into mechanical
energy.
24
HDSF is developed from conventional steam
explosion, extrusion or steam spout or swollen technologies, but
it is dramatically different from these physicochemical methods.
The comparison in detail between HDSF and other conventional
explosion technologies has been clearly illustrated in a recent
study.
24
HDSF adopts a structure in catapult explosion mode that
is principally composed of a cylinder and piston, which could
complete the explosion within 0.0875 s. The completion of
release of high density energy in an extremely short time could
provide enough force to disrupt and unfold the compact structure
of brous proteins and avoid a long time of violent treatment
under high temperature and pressure. Conventional steam
explosion technologies adopt the classical structure in valve
blow mode, which need at least several tens of seconds or
minutes to nish the release of high pressure steam.
24
Steam
with a powerful seepage force will penetrate the entirety of an
organic tissue structure. Given a rapid enough decompression
(HDSF), most of the steam in the biomass will quickly expand
and break free of the structure. Thus, the internal structure of the
biomass is disrupted by a mechanical shearing force. In a slow
decompression (conventional steam explosion technologies), the
pressure will have time to equalize across the structure, resulting
in a much smaller shearing force.
24
Moreover, HDSF could reduce the time of violent treatment of
biomass under high temperature and pressure. The temperature
of biomass treated by HDSF could immediately decrease to
50 °C or lower, which is also very benecial for the pretreatment
of biomass. Yu et al.
24
employed HDSF and conventional steam
explosion with the same thermochemical effect to treat maize
stalk, demonstrating that the most prominent factor in the action
of HDSF is the kinetic energy produced by steam explosion
rather than thermochemical reaction. This is also conrmed in
our unpublished study, as HDSF treatment could effectively
improve the extraction of heat-sensitive procyanidins with high
antioxidant activity from plants. The aim of this study is to
develop a novel, simple and eco-friendly pretreatment method to
extract and dissolve feather keratin, which could reduce or
eliminates the use of hazardous substances in the utilization of
keratins.
2. Materials and methods
2.1. Materials
Dry and clean feathers from chickens were supplied by
Hangzhou Venus Biological Nutrition Co., Ltd., China. The
feathers (6.0% moisture content (w/w, wet basis), 90.2% crude
protein and 1.6% crude fat (w/w, dry basis)) were defatted by
Soxhlet extraction with petroleum ether, washed with distilled
Table 1 Methods to extract and dissolve keratins from different biomaterials
Method Biomaterial Treatment conditions Aim Result (dissolubility)
Chemical methods Wool
8
7 M urea, SDS (2%, w/v), and 2-mercaptoethanol
(5%, v/v) at 50 °C for 12 h
Biodegradational lm
preparation
48%
Wool, chicken
feathers
9
2.4 M thiourea, 15 M urea, in 25 mM Tris-HCl, pH
8.5 containing 5% (v/v) DTT at 50 °C for 13d
Extraction 6770%
Wool
10
8 M urea, 0.14 M DTT, 5 mM Tris, pH 8.6 at 65 °C
for 2 h
Extraction 50%
Feather
11
50 g L
1
Na
2
S solution at 30 °C for 25 h Bio-polymer About 60%
Feather
12
A two-stage alkaline-enzymatic hydrolysis Keratin hydrolysate
Stage 1 : 0.10.3% KOH solution at 70 °C, for 24 h 24.08% (0.1% KOH)
Stage 2: 15% proteinase at 5070 °C for 48 h 90.83% (0.3% KOH)
Wool bres
13
8 M urea, 0.5 M sodium bisulphate,
SDS g/wool g = 0.6, pH = 6.5 at 65 °C for 5 h
Film preparation 33%
Feathers
14
Hydrophobic ionic liquid at 80 °C for 4 h Extraction About 21%
Silk broin
15
Ionic liquids at 30 °C after cooling from 100 °C Dissolution 12.24%
Wool keratin
5
Ionic liquids at 100130 °C for over 10 h Dissolution 11%
Physicochemical
methods
Silk broin
16
Hydrothermal treatment at 220 °C for 60 min Protein and amino acid
preparation
23.21% of protein;
75.39% of amino acid
Silk broin
bers
17
Radiation with 1 MeV electron beam, 1000 kGy
irradiation
Extraction 20% at 121 °C
Wool ber
18
Steam explosion at 00.8 MPa Modication The dissolving time in
2.5% NaOH increased.
Wool
19
Steam explosion at 220 °C for 10 min Conversion 18.66%, amino acids were
seriously damaged.
Feather
20
Extrusion with sodium sulte Bioplastic preparation 30%
Feather
21
Extrusion at 120 °C for 30 min, using a
combination of glycerol, water, and sodium sulte
as processing aids
Extrusion Much shorter
oligopeptides
Feather
22
Superheated water at 220 °C for 120 min Hydrolysis Oligopeptides
This journal is ©The Royal Society of Chemistry 2012 Green Chem., 2012, 14, 33523360 | 3353
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
water and dried at 40 °C for 48 h (the degreasing procedure in
this study was for the elimination of interference in the analysis
of treated samples. It was not found to inuence the treatment
efciency and could be omitted in the practical application of
HDSF in pretreatment of feathers.). All chemicals were of
analytical grade.
2.2 High density steam ash-explosion (HDSF) treatment
The experiments were carried out in the QBS-200B HDSF
device with a 5 L chamber from Gentle Science & Technology
Co. Ltd., China. The diagram of the HDSF process is shown in
Fig. 1. The structure, working principle and calculation of para-
meters can be obtained from a recent study.
24
The difference
between the catapult explosion mode of HDSF and the valve
blow mode of conventional steam explosion has been clearly
elaborated.
24
The whole structure (Fig. 1A and B) in catapult
explosion mode is composed of a cylinder and piston. The
piston is driven by three tailored pneumatic linear actuators. The
cylinder and piston set is installed in the shadowed area of
Fig. 1A. In the steam pressurization phase, the two parts are
tightly coupled (Fig. 1C). During the explosion, the piston,
driven by three pneumatic linear actuators, bursts out of the
cylinder, a process that is equivalent to a vessel suddenly frac-
turing into two halves (Fig. 1D). In the process of opening the
piston, the piston rapidly accelerates due to the kinetic energy of
the steam and material, as well as the force from the devices
driving the process. When the stroke of the piston reaches one
quarter of the cylinder diameter, the effective gas deation
passage the exposed area between the piston and cylinder
reaches the same area as the cylinder cross section.
24
The steam
was quickly forced into the chamber through the ination inlets.
When the pressure in the chamber reached expected the value
(35 s), ination inlets were closed and then the explosion was
completed within 0.0875 s. We also prolonged the time of
pressurization to 160 s (keeping the sample in the chamber
under expected pressure) to investigate the effects of the thermo-
chemical reaction. All experiments were carried out on the same
amount (1 kg) of air-dried feathers. The stream pressures were
set at 02.0 MPa. After HDSF treatment, samples were received
and frozen at 20 °C until the analyses.
2.3 Morphological characterization
Scanning electron microscopy (SEM) analysis was performed
with a Quanta 200 microscope (PHILIPS, Netherlands).
2.4 Dissolubility in different solvents and enzymatic
accessibility (pepsin digestibility) of feather keratins
The samples were dried in vacuum and ground to 40 mesh.
Protein solubility in several solvents: deionized water (pH 5.9),
sodium phosphate buffer (0.01 M, pH 7.5), sodium phosphate
buffer (0.01 M, pH 7.5) with 2% urea, sodium phosphate buffer
(0.01 M, pH 9.0) and 0.2% potassium hydroxide solution were
determined after a 20 min incubation at room temperature as
described previously by Araba and Dale (1990).
25
The solubility
was determined by measuring protein in the ltrate by the
Kjeldahl method. Pepsin-digestible N of each treatment was
determined by the procedure of the AOAC Ofcial Method
971.09 (AOAC, 2000)
26
using 0.2% pepsin (activity 1 : 10 000)
in 0.075 mol L
1
hydrochloric acid. In this assay, 1.000 g
(±0.010 g) of sample was incubated in 150 mL of pepsin solu-
tion at 45 °C for 16 h.
2.5 Amino acid analysis
The samples were hydrolyzed with hydrochloric acid
(6 mol L
1
) at 110 °C for 24 h in sealed tubes. An Alliance
(Agilent 1100) high performance liquid chromatograph (HPLC)
was used, and the eluate was detected at 338 nm. Tryptophan
was determined separately, as described by Molnar-Perl and
Pinter-Szakacs (1989).
27
2.6 Disulde-sulfhydryl analysis
Free thiol groups and disulde bonds were determined by a
solid-phase assay according to Chan and Wasserman (1993).
28
Colorimetric reaction with 2-nitro-5-thiosulfobenzoate (NTSB)
was conducted under the conditions described by Thannhauser
et al. (1987)
29
with the aim of disulde bond content determi-
nation. An extinction coefcient of 13 600 M
1
cm
1
was used
to calculate the number of sulfur containing groups.
2.7 Molecular weight distribution determination
The molecular weight distribution of feather keratins dissolved
in phosphate buffer (pH 7.5) with 2% urea and phosphate buffer
(pH 9.0) was assessed by gel permeation chromatography (GPC)
with a Sephadex G-50 column (1 × 100 cm). Feather keratins
Fig. 1 The diagram of the HDSF process. The whole structure (A and
B) in catapult explosion mode is composed of a cylinder and piston. The
cylinder and piston set is installed in the shadowed area of (A). In the
steam pressurization phase, the two parts are tightly coupled (C). During
the explosion, the piston, driven by three pneumatic linear actuators and
the kinetic energy of the steam and material, bursts out of the cylinder,
a process that is equivalent to a vessel suddenly fracturing into two
halves (D).
3354 |Green Chem., 2012, 14, 33523360 This journal is ©The Royal Society of Chemistry 2012
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
(30 mg) extracted using phosphate buffer ( pH 7.5) with 2% urea
and phosphate buffer (pH 9.0) were loaded onto a Sephadex
G-50 gel ltration column, pre-equilibrated and eluted with the
corresponding dissolved solutions. Separation was obtained at a
ow rate of 20 mL h
1
and eluted fractions (5.0 mL) were
pooled after spectrophotometric measurements at 280 nm. A
molecular weight calibration curve was prepared from the
average retention time of the following standards: bovine serum
albumin (66.2 kDa), ovalbumin (45.0 kDa), lactate dehydrogen-
ase (35.0 kDa), restriction endonuclease Bsp981 (25.0 kDa),
lysozyme (14.4 kDa), cytochrome C (12.5 kDa), aprotinin
(6.5 kDa), bacitracin (1450 Da).
2.8 X-ray diffraction analysis
The samples were dispersed onto a stub and placed within the
chamber of the analytical X-ray powder diffractometer (Bruker
AXS, Rheinfelden, Germany). Generator intensity was 40 kV,
generator current was 50 mA. The sample was then scanned
from 2θ=570°, in steps of 0.02°. The resultant graphs were
printed out on the Origin graph plotting package.
2.9 Fourier transform infrared (FTIR) spectroscopy
Infrared spectroscopy was performed using a FTIR 510 Nicolet
spectrophotometer using the diffuse reectance technique, spec-
tral resolution 4 cm
1
and 32 scans. The KBr pellet sampling
method was used to prepare the thin lm for testing. Spectra
were baseline corrected, smoothed with a nine-point Savitzky
Golay function. All spectra were shown after band-narrowing by
Fourier self-deconvolution (FSD) to resolve overlapping IR
bands using a half-bandwidth of 30.0 cm
1
and enhancement
factor k= 3.0. Second derivative spectra were obtained to
support the initial identication of band positions by deconvolu-
tion. These were carried out using the OMNIC software (Nicolet
Instrument Co., Madison, WI). The band positions obtained
from the above steps were then used as the initial guess for
curve-tting of the original spectra with Gaussian bands. For the
nal ts, the positions, heights, and widths of all bands were
varied simultaneously. The curving tting procedure was calcu-
lated on Peakt v4.12 software (SeaSolve Software Inc.).
2.10 Statistical analyses
Each experiment was carried out at least in triplicate. All statis-
tical analyses were conducted with SAS software (Version 8.0),
and P-0.05 was used to determine statistical signicance in
all tests.
3. Results and discussion
Extraction, dissolubility and enzymatic accessibility of feather
keratins after HDSF treatment
As is known, extraction and dissolubility of keratins from feath-
ers are the prerequisites for further utilization of feather keratins.
Many methods (Table 1) have been employed to directly extract
and dissolve keratins (chemical methods) or as pretreatments
(physicochemical methods) to improve the extraction and dis-
solubility of keratins. In this study, HDSF treatments were
employed to treat feathers. The exciting results could be seen
from Fig. 2. The dissolubility of feather keratins in all the sol-
vents tested in this study dramatically increased with the increase
of HDSF pressure. It was worth noting that the insoluble and
hard-to-degrade feather keratins could not dissolve in deionized
water (pH 5.9) and phosphate buffer (0.01 M, pH 7.5) at all, but
had a dissolubility of 12.9% and 22.0% after HDSF treatment at
2.0 MPa in the two solvents, respectively. This result is very
important and promising for the utilization of feather keratins,
because feathers are probably the most abundant and cheapest
keratinous materials in nature.
30
Although the dissolubility is
lower when compared with the current chemical methods in
Table 1, this makes possible the recycling of feathers using treat-
ments performed free of chemicals, only with water. As illus-
trated in Fig. 2, the dissolubility of feather keratins after 2.0 MPa
HDSF treatment dramatically increased to 65.0% from 2.0%
before HDSF treatment upon addition of 2.0% urea in phosphate
buffer (0.01 M, pH 7.5). Feather keratin primary chains are
10.210.4 kDa linear molecules containing a high number of
hydrophobic residues. The hydrophobic interaction is one of the
dominant stabilizers, which could be destroyed by urea. The
phosphate buffer (0.01 M, pH 9.0) also showed excellent dis-
solubility (68.0% at 2.0 MPa) for feather keratins treated by
HDSF. Compared with the severe chemical reactions in Table 1,
the phosphate buffer (0.01 M, pH 7.5) with 2% urea and phos-
phate buffer (0.01 M, pH 9.0) used in this study are milder, more
eco-friendly and efcient, indicating HDSF pretreatment could
efciently reduce or even eliminate the use of hazardous sub-
stances to improve the extraction and dissolubility of keratins.
The protein solubility in 0.2% potassium hydroxide solution is
always the common procedure to evaluate the pretreatment of
keratins. After HDSF pretreatment, the dissolubility of feather
keratins could reach 72.8% at 2.0 MPa from only 3.0% in the
original sample. It is much higher than those of the chemical
methods in Table 1 using harsh chemicals. From the point of
Fig. 2 The extraction, dissolubility and enzymatic accessibility of
feather keratins treated by HDSF treatment.
This journal is ©The Royal Society of Chemistry 2012 Green Chem., 2012, 14, 33523360 | 3355
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
view of eco-friendly conversion and utilization of extracted
keratins, it is not necessary to use 0.2% potassium hydroxide to
extract feather keratins.
Feathers are rich in crude protein (90%). However, they are
poorly susceptible to common proteolytic enzymes, such as
trypsin, pepsin, and papain, due to their intact and compact
structures.
3
Without pretreatment, only 10.5% of feather keratins
could be hydrolyzed by pepsin (Fig. 2). The pepsin digestibility
dramatically increased to 93.2% after HDSF treatment at
2.0 MPa, indicating that HDSF treatment could destroy the
tightly packed structures and increase the enzymatic accessibility.
Several researchers have employed chemical treatments to
increase the enzymatic accessibility of feather keratins. A recent
study
12
performed a two-stage alkaline-enzymatic hydrolysis
procedure using 0.10.3% potassium hydroxide solution at
70 °C for 24 h and 15% proteinase hydrolysis at 5070 °C for
48 h to obtain 24.08% and 90.83% of degradation in 0.1% and
0.3% potassium hydroxide, respectively. Moreover, many
researchers screened microorganisms that could secrete kerati-
nase for the degradation of feathers. This represents an environ-
mentally friendly approach, but requires a long production cycle
and tedious preparation. In a recent study,
23
twenty thermophilic
actinomycetes were selected that were able to degrade 8791%
of feather wastes after 72 h of cultivation. Obviously, the present
study could obtain higher enzymatic accessibility (93.2% pepsin
digestibility) after HDSF treatment without any harsh chemicals
for a shorter time. Nowadays, the growing demand for food,
materials, energy and reducing greenhouse gases has motivated
research on development of technologies to utilize renewable,
bio-based resources. Recycling of feathers is also a subject of
interest among animal nutritionists, because of their potential as
a cheap and alternative protein source. The present study demon-
strates that HDSF could convert more than 90% of feather
keratin to the pepsin-digestible form, which has extraordinary
appeal to the feed industry.
Morphological changes of feathers treated by HDSF
The effect of HDSF treatment (2.0 MPa) on the morphology of
feathers is illustrated in Fig. 3. SEM photos of untreated
(Fig. 3A and B) and treated feathers (Fig. 3CF) showed
obvious morphological changes induced by HDSF treatment. As
shown in Fig. 3A and B, feathers have a hierarchical structure
beginning with the level of the central barbs, which grow
directly from the quill. These central barbs are tiny quills
which also grow barbs. Quill fractions are composed of both the
inner and outer quill; the outer quill is more densely structured
than the inner quill, which is porous.
31
The hollow keratin bers
could increase the steamsolid contact area, facilitating the
penetration of steam with a powerful seepage force.
After HDSF treatment, most of the feathers were completely
destroyed and the structure changed from bers into an amor-
phous structure (Fig. 3C). The central axis of the feather dis-
appeared and the original structure of the feather could not be
identied. HDSF treatment destructured the bers and increased
the reactive area. It was benecial to enhance the extraction, dis-
solubility and enzymatic accessibility of feather keratins. In
addition, only a small proportion of feather bers still retained
their structure after HDSF treatment (Fig. 3DF). The untreated
feather has a smooth surface covered with a clutter of brils,
playing an important role in protecting the feather from damage.
The kinetic energy of the steam explosion fractured the feather
ber, along with debrillation of the ber, due to the removal of
the cementing materials (Fig. 3DF). After HDSF treatment,
the surface of the feather became damaged and rough (Fig. 3D
and F). It is interesting to note from Fig. 3DF that there are
many holes in the rough surface of feather bers, which indicates
the action of HDSF on the feather. During the HDSF process,
steam with a powerful seepage force will penetrate the interior
cells of the feather. Upon an explosive decompression (within
0.0875 s), most of the steam in the feather will quickly expand
and disrupt the internal structure of the feather, nally forming
the holes in the surface of feather when the steam rushes out.
This process is dramatically different from the conventional
steam explosion, steam spout or swollen or extrusion, which
mainly employ boiling at high temperature and pressure. In the
present study, we prolonged the time of the feather in the high
pressure to 160 s and then carried out steam explosion, but no
Fig. 3 SEM photos of untreated (A × 160 and B × 500) and treated
feathers (C × 600, D × 50, E × 500 and F × 1000) by HDSF pretreat-
ment (2.0 MPa).
3356 |Green Chem., 2012, 14, 33523360 This journal is ©The Royal Society of Chemistry 2012
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
additional effect was found on the treatment efciency, indicat-
ing that the kinetic energy produced in HDSF plays a major role
in the action of HDSF on feather keratins.
Effects of HDSF treatment on the molecular weight prole and
amino acids of feather keratins
The molecular weight prole of dissolved feather keratins is
related to the degree of degradation of keratin under HDSF treat-
ment. Most of the physicochemical methods in Table 1 caused
substantial damage to the protein chain due to the harsh con-
ditions, resulting in severe degradation of keratin and loss of
mechanical properties for application in eco-composites and
bioplastics. Whereas the chemical methods, such as the 2-mer-
captoethanol method, are a benchmark for good yield and un-
damaged keratin, but environmental harmfulness and high cost
mean they are not industrially viable. Based on the results of
extraction and dissolubility of feather keratins treated by HDSF
in this study, it is not necessary to use 0.2% potassium hydroxide
to dissolve feather keratins from the point of view of eco-friendly
processing and application of extracted keratins. The molecular
weight proles of dissolved feather keratins in phosphate buffer
(pH 7.5) with 2% urea and phosphate buffer (pH 9.0) were
assessed (Fig. 4). As shown in Fig. 4, the dissolved feather kera-
tins in phosphate buffer (pH 7.5) with 2% urea exhibited two
peaks of molecular weight distribution. One is around 10 kDa,
which represents the natural monomer of keratin with primary
chains of 10.210.4 kDa linear molecules.
32
Another is above
30 kDa, which represents soluble keratin of high molecular
weight. With the increase of pressure of HDSF treatment, the
peak of molecular weight of above 30 kDa increased. The dis-
solved feather keratins in phosphate buffer ( pH 9.0) displayed a
Table 2 Amino acids composition (g per 100 g crude protein) of the
samples from the steam explosion compared to reference feathers
a
Amino acids Control 1.4 MPa 1.6 MPa 1.8 MPa 2.0 MPa
ASP 5.42
a
5.37
a
5.64
a
5.32
a
4.88
a
GLU 8.41
a
8.59
a
8.72
a
8.41
a
8.25
a
SER 9.33
a
9.53
a
9.41
a
8.99
a
8.69
a
HIS 0.50
a
0.55
a
0.60
a
0.53
a
0.48
a
GLY 7.16
a
8.14
a
7.97
a
7.88
a
7.97
a
THR 4.04
a
3.99
a
4.01
a
3.86
a
3.75
a
ARG 4.96
a
5.15
a
5.11
a
5.00
a
4.94
a
ALA 3.31
a
3.90
a
3.80
a
3.73
a
3.84
a
TYR 4.15
a
4.11
a
4.10
a
3.99
a
3.93
a
CYS-S 5.07
a
2.16
b
2.21
b
1.66
c
1.16
d
VAL 6.06
a
6.49
a
6.48
a
6.41
a
6.32
a
MET 0.41
a
0.46
a
0.50
a
0.49
a
0.47
a
PHE 3.18
a
3.81
a
3.66
a
3.56
a
3.62
a
ILE 3.51
a
3.83
a
3.85
a
3.78
a
3.77
a
LEU 6.16
a
7.04
a
6.92
a
6.74
a
6.68
a
LYS 1.12
a
1.11
a
1.18
a
1.10
a
1.05
a
PRO 10.03
a
12.13
a
10.13
a
10.68
a
11.52
a
a
The value is the average of three replicates. The data marked with
different letters in the same row were statistically different at P= 0.05.
Fig. 4 The molecular weight distribution of feather keratins dissolved
in phosphate buffer (pH 7.5) with 2% urea and phosphate buffer (pH
9.0) assessed by gel permeation chromatography (GPC) with a Sephadex
G-50 column (1 × 100 cm).
Fig. 5 The changes of disulde bonds and free thiol groups of feather
keratins in HDSF treatment.
Fig. 6 The X-ray diffraction pattern of feather keratins before and after
HDSF treatment.
This journal is ©The Royal Society of Chemistry 2012 Green Chem., 2012, 14, 33523360 | 3357
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
similar molecular weight distribution (data not shown), indicat-
ing HDSF treatment did not cause substantial damage to the
protein chain of feather keratin.
The effects of HDSF treatment on the amino acids of feather
keratins were further evaluated in this study (Table 2). A signi-
cant reduction (p< 0.05) in cystine as a function of HDSF
pressure was found. Other amino acids in feathers were
unchanged during HDSF treatment ( p> 0.05). Feather keratin
contains a large number of disulde bonds, associated with the
amino acid cystine, which cross-link the protein chains.
Disulde bonds are mainly responsible for the high stability,
hardness and insolubility of feather keratin. Thus almost all the
chemical and physicochemical methods (Table 1) exert their
action by rst breaking the intermolecular disulde bonds. Fig. 5
illustrated the quantitative change of disulde bonds and free
sulfhydryl in feather keratins treated by HDSF. With the increase
of pressure of HDSF, the disulde bonds signicantly decreased
from 330 to 60 nmol mg
1
, whereas free sulfhydryls increased
from 5 to 27 nmol mg
1
, indicating that disulde bonds of
feather keratins were destroyed by HDSF. From the results of
Fig. 5 and Table 2, it could be deduced that breaking of disulde
bonds as a result of cystine destruction caused a considerable
amount of keratins to be soluble and susceptible to enzyme.
Generally, HDSF treatment is not a severe process, which could
cleave the intermolecular disulde bonds without causing sub-
stantial damage to the protein chain. Different from the HDSF
treatment, serious losses of lysine, cystine, serine and threonine
were found in other treatments of feathers, such as heat treat-
ment,
33,34
steam extrusion
35
and high temperature and pressure
treatment in combination with lime treatment.
36
Fig. 7 FTIR spectra (A) and resolved amide III of feather keratins before (B) and after HDSF treatment at different pressures (C: 1.4 MPa; D:
1.6 MPa; E: 1.8 MPa; F: 2.0 MPa).
3358 |Green Chem., 2012, 14, 33523360 This journal is ©The Royal Society of Chemistry 2012
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
Effects of HDSF treatment on the structure of feather keratins
Besides disulde bonds, the crystallinity also takes an important
role in the high strength and stiffness of feather keratins. The
original and HDSF treated feather samples were taken to
examine the crystal structure by wide angle X-ray diffraction
(Fig. 6). As shown in Fig. 6, the original feather sample dis-
played a typical diffraction pattern of β-keratins (feather keratins)
with a prominent 2θpeak at 9.2° corresponding to the α-helix
conguration of the feather keratin, and a more intense band at
19.6° indexed as its beta strand secondary structure.
7
After
HDSF treatment, the samples at different pressures of HDSF
presented the identied X-ray diffraction patterns, characterized
by decrease in the intensity of the peaks, indicating that crystals
were signicantly destroyed under HDSF treatment. Moreover, it
was observed that the peak at 9.2° became more intense and
dominant than the band at 19.6°, which was ascribed to the
decomposition of the β-sheet of feather keratins. As is known,
feathersmechanical stability, insolubility and resistance to pro-
teolytic digestion are consequences of the tight packing of the
protein chain in β-sheets (β-keratin) into a supercoiled polypep-
tide chain.
37
The decrease of crystallinity and decomposition of
the β-sheet structure could denitely contribute to the improve-
ment of extraction, dissolubility and enzymatic accessibility of
feather keratins. Interestingly, no signicant changes even
increase of the crystallinity of feather keratins were found in
chemical treatments, such as alkali treatment (0.75 mol L
1
sodium hydroxide)
38
and physicochemical treatments, such as
conventional steam explosion
8
or extrusion,
21
demonstrating the
different mechanisms and actions of these means on the keratins.
The FTIR spectra in the region of 4000400 cm
1
of untreated
and HDSF treated feather are shown in Fig. 7A. It can be seen
from Fig. 7A that the samples before and after HDSF treatments
at 1.42.0 MPa exhibited similar absorption bands at 3300 cm
1
(amide A: NH and OH), 2850 cm
1
(CH
2
), 1627 cm
1
(amide I: CvO), 1520 cm
1
(amide II: NH and CH) and
1233 cm
1
(amide III: CH, NH, CC and CvO). This indi-
cates that neither production of new functional groups and
chemical bonds nor substantial damage to the protein chain of
feather keratin was induced by HDSF treatment. Methods cur-
rently being used to extract information on protein secondary
structure from infrared spectra are based on empirical correlation
between the frequencies of certain vibrational modes and
types of secondary structure of polypeptide chains, such as
α-helix, β-sheet, β-turn, and random coil.
39
As is known,
the amide I region (17001600 cm
1
) and amide II region
(15801480 cm
1
) as the candidates for secondary structure
characterizations suffer from several limitations, such as strong
interference from water vibrational bands and overlap of
revolved bands corresponding to secondary structures.
39
In this
study, the easily resolved and better dened amide III region
(13501200 cm
1
) was employed for quantitative analysis of
protein secondary structure, due to the in-phase combination of
NH in-plane bending and CN stretching vibrations.
39
The
amide III mode was further resolved in Gaussian bands
(Fig. 7BF) corresponding to α-helix (13301295 cm
1
), β-turn
(12951270 cm
1
), random coil (12701250 cm
1
) and β-sheet
(12501220 cm
1
). The bands not related to the secondary struc-
ture are dened as other. From the resolved amide III of
feather keratin (Fig. 8), the β-sheet decreased and the unordered
structure (random coil) increased after HDSF treatment, which
was consistent with the observation of X-ray diffraction. The dis-
sociation of secondary structures, especially the decrease of
β-sheet is essential for the destabilization and rearrangement of
β-keratin.
Conclusions
Based on the present study, a diagram could be proposed to
describe the action of HDSF treatment on feather keratins
Fig. 8 The changes of secondary structures of feather keratins after
HDSF treatment.
Fig. 9 A diagram to describe the action of HDSF treatment on feather
keratins.
This journal is ©The Royal Society of Chemistry 2012 Green Chem., 2012, 14, 33523360 | 3359
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
(Fig. 9). In HDSF treatment, steam with a powerful seepage
force rst penetrates into the brous tissues and cells of feather,
and then quickly expands and breaks free of the structure
upon an explosive decompression at supersonic speed (within
0.0875 s).
This process effectively destroyed or destabilized β-sheet crys-
tals and intermolecular disulde bonds without causing substan-
tial damage to the keratin protein chain, dramatically increasing
the enzymatic accessibility, extraction, and dissolubility of
feather keratins in polar solvents like water, salt solution and
weak bases. HDSF is a physicochemical process, which includes
two phases: the pressurization phase (35 s) and the explosion
phase (within 0.0875 s). Denitely, the thermal factor plays a
role in the action of HDSF treatment of feathers. However, the
time of materials in high temperatures is very short compared
with the conventional physicochemical methods, including
steam explosion, steam spout or swollen or extrusion and the
temperature of materials could decrease to 50 °C or lower
immediately after HDSF treatment, indicating that HDSF treat-
ment is not a severe process.
HDSF treatment could be a sustainable and practical pretreat-
ment for the extraction of feather keratin for exploitation of bio-
materials and conversion of feathers to nutrient animal feed,
instead of the current chemical (strong acid, alkali or catalytic)
hydrolysis and hydrothermal treatment (high pressure and temp-
erature). Moreover, HDSF treatment could be performed on a
large scale now that HDSF equipment with a 5 m
3
chamber has
been developed.
24
In the catapult explosion mode of HDSF, the
valve opening speed transforms into the piston movement speed
in the cylinder. Thus the large diameter piston can get an equal
proportion of driving force to the small diameter piston, and
achieve the same speed breaking away from cylinder. Therefore,
the explosion speed and treatment efciency in scale-up appli-
cations of catapult explosion can remain consistent with the
small equipment.
Acknowledgements
The authors gratefully acknowledge the supported by Project of
the National Natural Science Foundation of PR China
(31271977). This study was also supported by 111 project-
B07029, Program for Changjiang Scholars and Innovative
Research Team in University and Project Funded by the Priority
Academic Program Development of Jiangsu Higher Education
Institutions.
References
1 J. P. Andrew, Biomacromolecules, 2009, 10,1.
2 A. Ullah, T. Vasanthan, D. Bressler, A. L. Elias and J. Wu, Biomacro-
molecules, 2011, 12, 3826.
3 D. A. D. Parry and A. C. T. North, J. Struct. Biol., 1998, 122, 67.
4 E. Deydier, R. Guilet, S. Sarda and P. Sharrock, J. Hazard. Mater., 2005,
121, 141.
5 H. Xie, S. Li and S. Zhang, Green Chem., 2005, 7, 606.
6 M. E. Vallejos, M. D. Zambon, M. C. Area and A. A. D. S. Curvelo,
Green Chem., 2012, 14, 1982.
7 I. A. Aguayo-Villarreal, A. Bonilla-Petriciolet, V. Hernández-Montoya,
M. A. Montes-Moránc and H. E. Reynel-Avila, J. Hazard. Mater., 2011,
167, 67.
8 K. Yamauchi, A. Yamauchi, T. Kusunoki, A. Kohda and Y. Konishi,
J. Biomed. Mater. Res., 1996, 31, 439.
9 A. Nakamura, M. Arimoto, K. Takeuchi and T. Fujii, Biol. Pharm. Bull.,
2002, 25, 569.
10 M. Zoccola, A. Aluigi and C. Tonin, J. Mol. Struct., 2009, 938, 35.
11 A. J. Poole, R. E. Lyons and J. S. Church, J. Polym. Environ., 2011, 19,
995.
12 P. Mokrejs, P. Svoboda, J. Hrncirik, D. Janacova and V. Vasek, Wa s t e
Manage. Res., 2010, 29, 260.
13 C. Tonin, A. Aluigi, C. Vineis, A. Varesano, A. Montarsolo and
F. Ferrero, J. Therm. Anal. Calorim., 2007, 89, 601.
14 Y. Wang and X. Cao, Process Biochem., 2012, 47, 896.
15 D. M. Phillips, L. F. Drummy, D. G. Conrady, D. M. Fox, R. R. Naik, M.
O. Stone, P. C. Trulove, H. C. De Long and R. A. Mantz, J. Am. Chem.
Soc., 2004, 126, 14350.
16 W. Lamoolphak, W. De-Eknamkul and A. Shotipruk, Bioresour. Technol.,
2008, 99, 7678.
17 H. Takeshita, K. Ishida, Y. Kamiishi, F. Yoshii and T. Kume, Macromol.
Mater. Eng., 2000, 283, 126.
18 W. Xu, G. Ke, J. Wu and X. Wang, Eur. Polym. J., 2006, 42, 2168.
19 C. Tonin, M. Zoccola, A. Aluigi, A. Varesano, A. Montarsolo, C. Vineis
and F. Zimbardi, Eur. Polym. J., 2006, 42, 2168.
20 A. Ullah, T. Vasanthan, D. Bressler, A. L. Elias and J. Wu, Biomacro-
molecules, 2011, 12, 3826.
21 J. R. Barone, W. F. Schmidt and N. T. Gregoire, J. Appl. Polym. Sci.,
2006, 100, 1432.
22 J. Yin, S. Rastogi, A. E. Ferry and C. Popescu, Biomacromolecules,
2007, 8, 800.
23 E. Vasileva-Tonkova, A. Gousterova and G. Neshev, Int. Biodeterior. Bio-
degrad., 2009, 63, 1008.
24 Z. Yu, B. Zhang, F. Yu, G. Xu and A. Song, Bioresour. Technol., 2012,
12, 335.
25 M. Araba and N. Dale, Poult. Sci., 1990, 69, 76.
26 AOAC, Ofcial Methods of Analysis, AOAC International, Gaithersburg,
USA, 17th edn, 2000.
27 I. Molnar-Perl and M. Pinter-Szakacs, Anal. Biochem., 1989, 177, 16.
28 K. Chan and B. P. Wasserman, Cereal Chem., 1993, 70, 22.
29 T. W. Thannhauser, Y. Konishi and H. A. Scheraga, Methods Enzymol.,
1987, 143, 115.
30 A. A. Onifade, N. A. Al-Sane, A. A. Al-Mussallam and S. Al-Zarbam,
Bioresour. Technol., 1998, 6,1.
31 N. Reddy and Y. Yang, J. Polym. Environ., 2007, 15, 81.
32 A. M. Woodin, Biochem. J., 1954, 57, 99.
33 M. C. Papadopoulos, Agric. Wastes, 1985, 14, 275.
34 J. S. Moritz and J. D. Latshaw, Poult. Sci., 2001, 80, 79.
35 X. Zhu, Z. Jin and D. Liu, J. Chin. Cereal Oils Assoc., 1998, 13, 16.
36 G. Coward-Kelly, V. S. Chang, F. K. Agbogbo and M. T. Holtzapple,
Bioresour. Technol., 2006, 97, 1337.
37 A. A. Onifade, N. A. Al-Sane, A. A. Al-Musallam and S. Al-Zarban,
Bioresour. Technol., 1998, 66,1.
38 P. Sun, Z. Liu and Z. Liu, Ind. Eng. Chem. Res., 2009, 48, 6882.
39 S. Cai and B. R. Singh, Biophys. Chem., 1999, 80,7.
3360 |Green Chem., 2012, 14, 33523360 This journal is ©The Royal Society of Chemistry 2012
Published on 01 October 2012. Downloaded by University of California - Los Angeles on 20/04/2016 05:35:56.
View Article Online
... It is estimated that approximately several million tons of feathers could be generated annually from the poultry industry globally [1], [2]. Due to inappropriate disposal, this oversupply has created several environmental problems as well as serious pollution and health dangers [3], [4], [5]. Feathers are mainly composed of keratin, a durable and resilient proteinaceous biopolymer [6]. ...
Article
Full-text available
Poultry waste management is a rising source of worry due to its detrimental effects on the environment and associated health issues. In response, recycling poultry manure using biological processes has become a potential strategy. Bioreactors are one of these techniques that have gained popularity. The ideal conditions for the microbial breakdown and production of organic fertilizers and biogas from poultry waste are provided by bioreactors. By using this approach, trash is managed effectively and greenhouse gas emissions are decreased, supporting sustainable agriculture practices. By dwelling on negative effects of chicken waste on the environment and human health, it has become an urgent need to manage such waste by developing a bioreactor for it. In this study, Stenotrophomonas maltophilia KARUNA5 strain isolated from poultry waste in our earlier studies was used and developed a lab-scale bioreactor which was optimized for poultry waste treatment. Shake flask studies at laboratory scale were conducted to optimize the conditions of bioreactor which helped us to identify key factors that contributed to the feather degradation to a large extent. These factors included a 2.0 % inoculum size of the culture (1O. D620nm), a feather concentration of 1.5 % in a whole feather medium with pH 8.0 kept on shaker (agitation rate of 140 rpm) at a 30°C for 6 days. The 78.5 % of feather degradation was observed. The lab-scale bioreactor was designed which could efficiently degrade the poultry waste, offering a sustainable method of waste management. Further, the foliar (liquid fertilizer) potential of the feather hydrolysate product was evaluated with chemical analysis (NPK analysis) showed the NPK Σ 2.68 % w/w (>1.15 minimum) fulfilling the specification of fertilizer.
... The peaks with this broadness showed that the α-helix and β-sheets that are present in raw feathers of chicken are deformed and broken due to the process of alkaline hydrolysis for keratin extraction. The two strong peaks at 9-11 0 and 15-31 0 were allocated to α-helix and β-sheet respectively [79][80][81]. It is evident from a study that the chicken feather and extracted keratin held small amount of α-helix, whereas, significant amount of β-sheet conformation [82]. ...
... It has many potential advantages, including lower cost, lower energy, and environmental friendliness . Zhao, Yang, Zhang, and Wu (2012) and Zhang, Zhao, and Yang (2015) treated feather waste with steam explosion, which significantly improved the solubility and enzyme accessibility of feathers. Similar to the structure of feathers, ESM also contains a large number of disulfide bonds and is expected to be dissolved by steam explosion. ...
Preprint
Full-text available
Waste wool, a by-product of sheep husbandry, is primarily composed of keratin, which has potential use in cosmetic products, biotechnology and medical applications. Traditional chemical methods for keratin extraction face limitations due to health and environmental concerns, and thus, green alternatives such as deep eutectic solvents (DESs) have been developed. The aim of the present study was to determine whether a natural DES composed of choline chloride and lactic acid can be used to extract keratin from sheep wool. The dissolution and keratin yield were investigated under different reaction times, temperatures, and wool-to-DES ratios. Fractions of water-soluble keratin and water-insoluble keratin aggregates were obtained by dialysis, centrifugation, and drying. The results showed that both the dissolution of wool and the yield of recovered keratin increased as the reaction time and temperature increased from 1 to 8 h and from 80 to 110◦C, respectively.With the longest reaction time of 8 hours and the highest temperature of 110◦C, the yield of water-soluble keratin increased to 23 %. Characterizations revealed that DES detached first the cuticular scales from the surface of wool fibres and subsequently degraded wool cortex layer. The extracted water-soluble keratin retained the characteristic amide structure of wool, but the secondary structure of keratin was converted to cross-β sheet form. Moreover, it had a narrow size distribution and a molecular weight of 10 kDa. The results of this study indicate that the proposed DES can be used as a green solvent for recovering keratin from sheep wool.
Article
Full-text available
Defined peptides with exclusive molecular functionalities from biomass streams provide an untapped treasure for innovative biogenic specialty chemicals and materials. In this context, feather keratin, a natural structural protein with high l-cysteine content, enables access to polythiol-containing peptides, which can be used as matrix compounds for new materials per se and be specifically modified via their amino and acid moieties. This study describes an innovative two-step approach for tailored feather keratin fragmentation involving selective enzymatic hydrolysis followed by optional chemical reduction. Several proteases were investigated to serve as a benchmark for the decomposition of chicken feather keratin, and we succeeded in the controlled decomposition of chicken feathers using trypsin and other specific proteases, producing polythiol-containing peptide fragments. We were able to implement a green hydrolysis process without the need for any denaturants or reducing agents and achieved yields of soluble protein up to 81% (w/w) and thiol concentrations up to 21 mmol L⁻¹. The obtained hydrolysates were used to produce peptide films, and the scalability of the newly developed hydrolysis process has been demonstrated in 25 L batch reactions.
Article
Full-text available
Two experiments were conducted to compare urease activity, orange G-binding capacity, trypsin inhibitor activity (TIA), and protein solubility (PS) in a potassium hydroxide solution as indicators of underprocessed soybean meal (SBM). Subsamples of raw dehulled, hexane-extracted SBM (RSBM) were autoclaved for varying lengths of time at 121 C. Orange G binding capacity of the meal did not vary substantially and did not clearly differentiate between raw and heated SBM. Urease activity, TIA, and PS all reflected degrees of heat treatment. However, PS values were more closely correlated with chick performance and paralleled with trypsin inhibitor concentration in meals for all autoclaving times used. Chick growth was optimized when RSBM was autoclaved for 15 min with a resulting PS of 75%. Diets containing this meal gave lowest pancreas weight (as percentage of body weight). It appears that PS is of value in the detection of underprocessed SBM. Protein solubility values in excess of 85% or less than 70% indicate under- or overprocessing of SBM, respectively.
Article
Full-text available
Keratin obtained from poultry feathers was extruded at 120°C using a combination of glycerol, water, and sodium sulfite as processing aids. Rheological proper-ties were assessed as a function of water, glycerol, and sodium sulfite content as well as extruder die temperature. The lowest viscosity blends at a constant feather keratin concentration of 60 wt % were found at glycerol concentra-tions that were higher than the water concentration and sodium sulfite concentrations of 3– 4 wt % of the feather keratin fraction. For the melt state, higher or lower sodium sulfite concentrations resulted in increased viscosity. In the solid state, it was observed that processing induced orien-tation increased the tensile properties of the extrudates. Raman spectroscopy and DSC showed that there was a transition from ␣-helix to ␤-sheet at sodium sulfite concen-trations of less than 4 wt %. At greater sodium sulfite con-centration, increased crystallinity was found, because kera-tin chains could be extended more during processing.
Article
Keratin was extracted from chicken feathers by using a hydrophobic ionic liquid (IL), 1-hydroxyethyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide ([HOEMIm][NTf2]). Extracted keratin has good solubility in water while the ionic liquid is immiscible with water, and therefore the extracted keratin could be easily separated from the reaction system by water. The effects of ionic liquid, NaHSO3, reaction temperature and time were investigated and extracting conditions were optimized. The maximum yield of keratin was up to about 21% with mass ratio of feathers to NaHSO31:1 and mass ratio of feathers to ionic liquid 1:40 at 80 °C for 4 h. Moreover, there was no obvious loss in the yield after ionic liquid was reused for five batches under optimized conditions. In addition, the recovery of ionic liquid was about 95% each time. The results indicated that [HOEMIm][NTf2] was very efficient as catalyst and solvent for dissolving feathers and could be easily recovered due to its hydrophobicity.
Article
1-Butyl-3-methylimidazolium chloride ionic liquid has been developed for the dissolution and regeneration of wool keratin fibers, which can be used to prepare wool keratin/cellulose blended materials directly.
Article
Wool fiber was modified by steam explosion in this study. SEM results show that some scales on the fiber surface were cleaved and tiny grooves generated during the explosion. FTIR results suggest no evident changes in the chemical composition of the fiber after the explosion treatment. However, the crystallinity of the fiber decreased slightly as the steam pressure increased based on the X-ray results. In the thermal analysis, DSC results show that the temperature corresponding to vaporization of absorbed water and cleavage of disulfide bonds respectively decreased as the steam pressure increased. The reduction in thermal decomposition energy of the treated fiber indicates that steam explosion might have destroyed some crystals and crosslinks of macromolecular chains in the fiber. The treatment also led to some alterations of the fiber properties, including reduction in strength, moisture regain and solubility in caustic solution.
Article
Cereal Chem. 70(1):22-26 A direct colorimetric method that simultaneously combines measure- material, absorbance at 412 nm was read. This assay was highly repro- ment of solubilized and insoluble thiol groups and disulfide bonds in ducible, and measurements agreed with direct amino acid analysis. Twin- corn meal-based materials is described. Ellman's reagent, 5,5'-dithiobis screw extrusion of corn meal at 1501C at moisture levels of 16 and 18% (2-nitrobenzoic acid), which reacts specifically with thiol groups, or had no significant effect on cysteine or disulfide bond levels. Other possible disodium 2-nitro-5-thiosulfobenzoate, which reacts with cysteine and thiol changes such as disulfide bond rearrangements could not be determined groups formed after reduction of disulfide bonds with sodium sulfite, by the mixed-phase assay. This method provides a rapid and convenient were reacted directly with corn meal in the presence of surfactants (urea means for screening thiol and disulfide levels in insoluble proteinaceous and/ or sodium dodecyl sulfate), releasing the soluble chromophore materials. 2-nitro-5-thiobenzoate. After a clarification step to remove suspended Disulfide bonds are thought to play an important role in the texture of cereal-based products. However, because of the hydro- phobic and insoluble nature of cereal proteins, quantification of thiol and disulfide bonds has proven difficult. Complete extraction of corn meal protein with sodium dodecyl sulfate (SDS), urea, and a reducing agent leads to cleavage of disulfide bonds. Further- more, since extraction with SDS and urea results in only partial solubilization of the protein, assays based on solubilization fol- lowed by measurement of thiol and disulfide content often yield highly variable results and underestimates of true thiol and disul- fide group content. A technique that avoids this initial protein solubilization step would largely eliminate these problems. This article describes a solid-phase assay that simultaneously combines quantification of soluble and particulate thiol and disul- fide groups in cereal-based proteins. The principle of this method is to suspend the entire sample in urea and to react it with a color reagent that will simultaneously react with both soluble and insoluble proteins, with release of a soluble chromophore. Ellman's reagent, 5,5'-dithiobis (2-nitrobenzoic acid) (DTNB), which reacts specifically with thiol groups (Ellman 1958, Riddles et al 1983), and disodium 2-nitro-5-thiosulfobenzoate (NTSB2-), which is used to quantify disulfide group content (Thannhauser et al 1987), are ideally suited for this purpose, since reaction with either results in the release of the 2-nitro-5-thiobenzoate anion (NTB2- ), which is soluble in aqueous solution. Following the removal of insoluble material by clarification steps, absorbance at 412 nm is then read. This method was used to assess the effects of twin-screw extrusion processing on thiol and disulfide levels in corn meal. MATERIALS AND METHODS Materials Corn meal (12% moisture, 7% protein, 0.7% oil, 0.5% fiber, 0.4% ash, and 79.4% N-free extract) was obtained from Lauhoff Grain Co., Danville, IL. DTNB and NTSB2- were obtained from Aldrich Chemical Co., Milwaukee, WI. Extrusion was conducted in a Brabender type 2003 single-screw extruder with an axially ground barrel; length-to-diameter ratio, 20; screw diameter, 1.9 cm (0.75 in.); screw length, 38.1 cm (15 in.); and compression ratio, 1:3.
Article
The chemical and structural characteristics of wool and horn–hoof were compared with the aim of better addressing possible exploitation of protein biomasses available as waste from textile industry and butchery. Amino acid analysis showed that wool has a higher amount of cystine and a lower amount of the amino acids that favour α-helix formation than horn–hoof. The difference in the α-helix content is confirmed by FTIR spectroscopy. Electrophoresis separation patterns showed two characteristic protein fractions related to low-sulphur proteins (between 60,000 and 45,000Da) in wool, while different low-sulphur proteins are present in horn–hoof. These data are partially confirmed by DSC analyses that showed different endothermic peaks at temperatures higher than 200°C in the horn–hoof thermograms, probably due to denaturation of α-keratins at different molecular weights. Moreover, wool keratin was more hygroscopic and showed a higher extractability with reducing agents than horn–hoof. On the basis of these results, waste wool is a more suitable source than horn–hoof for uses involving protein extraction, but application can be envisaged also in surfactant foams for fire extinguishers and slow-release nitrogen fertilizer.
Article
Low liquid–solid ratio (LSR) can be used to obtain high-content xylo-oligosaccharide (XOS) spend liquor by hot water pretreatment. Developing a technology based on low LSR results in more efficient water usage in the system and thus in lower capital and operating costs. Xylans from xylan rich agro-industrial waste are abundant hemicellulosic polymers with enormous potential for industrial applications. Currently, freeze-dried xylo-oligosaccharides are used as bio-based polymers and hydrolysates containing high xylose contents are converted to several chemical products. In this study, sugarcane bagasse was treated with water at low LSRs and mild temperatures in order to assess the effects of varying the pretreatment conditions on the xylo-oligosaccharide and xylose concentrations, and use a central composite experimental design to optimize the process parameters. The pretreatments were performed in the ranges temperature: 143.3–176.7 °C, time: 20–70 min and LSR: 1 : 1 to 11 : 1 (g g −1 ). The maximum concentrations of xylose and xylan were 13.76 and 36.18 g L −1 (equivalent to 48.29 g L −1 of xylan), respectively, which were achieved by treating bagasse at 170 °C for 60 min, with LSR of 3gg −1 . The amount of xylan removed under these conditions was almost 57%. The soluble xylan consisted mainly of xylo-oligosaccharides (74 wt% of the identified compound in the spent liquor).
Article
Advances in microbial enzyme technology, keratinolytic proteases in this case, offer considerable opportunities for a low-energy consuming technology for bioconversion of poultry feathers from a potent pollutant to a nutritionally upgraded protein-rich feedstuff for livestock. A compendium of recent information on microbial keratinolysis in nature and infection (dermatophytoses) has been provided as underscoring feasible harnessing of the biotechnology for nutritional improvement of feathers, and as an alternative to conventional hydrothermal processing. Supporting evidence of a nutritional (amino acid) upgrading sequel to diverse microbial treatments of feathers, and positive results obtained from growth studies in rats and chicks have been presented. The paper concludes with suggestions for avenues of application of biotechnology for nutritional improvement of feather (and other keratins) as feedstuffs for livestock
Article
To improve the sorption capacity of Cr(VI) ions, chicken feathers (CFs) were chemically treated by several methods: (1) To clarify the changes in the structure and morphology of CFs, CFs were treated with aqueous NaOH solutions of varying concentrations. The results suggest that reactions occur on the surface rather than in the interior of the CFs and that keratin fragments exfoliate from the CF surface layer by layer and then dissolve in aqueous solution. (2) Keratin fragments could be rejoined to the CF surface by cross-linking with epichlorohydrin (Epi) through a series of ring-opening, recyclization, and ring-opening reactions of Epi in aqueous NaOH solution, and Epi-modified CF (EpiCF) was obtained. (3) CFs were functionalized with ethylenediamine (EA) in aqueous NaOH solution by cross-linking with Epi (EAEpiCF). The structure and properties of chemically treated CFs were investigated by FT-IR spectroscopy, XRD, SEM, elemental analysis, and water contact-angle measurements. CFs treated with NaOH were found to be similar in structure and properties to each other and exhibited a relatively low sorption capacity for removal of Cr(VI) ions from water, but a higher capacity than raw CFs. EpiCF exhibited an excellent capacity for adsorbing relatively low concentrations (10 ppm) of Cr(VI) ions but a low sorption capacity for Cu(II) ions in water. EAEpiCF was found to be hydrophilic and to exhibit 90% efficiency for removing Cr(VI) ions in the concentration range of 10−80 ppm but to have a low sorption capacity for Cu(II) in water. Equilibrium sorption isotherms for the raw and chemically treated CFs fit the Freundlich and Dubinin−Radushkevich (D−R) isotherm models; however, the Langmuir model was not able to describe the sorption behavior. The values of the free sorption energy (E) were calculated. The E values of NaOH-treated CFs were found to be similar to each other and much larger than that of raw CFs for the sorption of Cr(VI), and the sorption processes were found to be physical. The process of EpiCF adsorbing Cr(VI) ions from water is chemisorption, and that for Cu(II) ions is physisorption. The free sorption energy of EAEpiCF sorption of Cr(VI) ions was close to 8 kJ·mol−1 (7.274 kJ·mol−1), so that chemical interactions might exist between EAEpiCF and Cr(VI), whereas for Cu(II), the sorption process is physical adsorption.