Changes to Collagen Structure during Leather Processing - Journal of

Feb 6, 2015 - Katie H. Sizeland†, Richard L. Edmonds‡, Melissa M. Basil-Jones†, Nigel ...... Brodsky , B.; Eikenberry , E. F.; Cassidy , K. An u...
9 downloads 0 Views 2MB Size
Subscriber access provided by MICHIGAN STATE UNIVERSITY | MSU LIBRARIES

Article

Changes to Collagen Structure During Leather Processing Katie H. Sizeland, Richard L. Edmonds, Melissa M. Basil-Jones, Nigel Kirby, Adrian Hawley, Stephen Mudie, and Richard G. Haverkamp J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/jf506357j • Publication Date (Web): 06 Feb 2015 Downloaded from http://pubs.acs.org on February 18, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

Journal of Agricultural and Food Chemistry

1

Changes to Collagen Structure During Leather

2

Processing

3

Katie H. Sizeland†, Richard L. Edmonds‡, Melissa M. Basil-Jones†, Nigel Kirby§, Adrian

4

Hawley§, Stephen Mudie§, Richard G. Haverkamp†*

5

†School of Engineering and Advanced Technology, Massey University, Private Bag 11222,

6

Palmerston North, New Zealand; ‡Leather and Shoe Research Association, Palmerston North

7

4442, New Zealand, §Australian Synchrotron, 800 Blackburn Road, Clayton, Melbourne,

8

Australia.

9

*[email protected]

10 11 12

RECEIVED DATE (to be automatically inserted after your manuscript is accepted if required according to the journal that you are submitting your paper to)

13

14

Abstract

15

As hides and skins are processed to produce leather, chemical and physical changes take

16

place that affect the strength and other physical properties of the material. The structural basis

17

of these changes at the level of the collagen fibrils is not fully understood and forms the basis

18

of this investigation. Synchrotron-based small-angle X-ray scattering (SAXS) is used to

19

quantify fibril orientation and D-spacing through eight stages of processing from fresh green

20

ovine skins to staked dry crust leather. Both the D-spacing and fibril orientation change with 1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

21

processing. The changes in thickness of the leather during processing impacts on the fibril

22

orientation index (OI) and accounts for much of the OI difference between process stages.

23

After thickness is accounted for the main difference in OI is due to the hydration state of the

24

material with dry materials being less oriented than wet. Similarly significant differences in

25

D-spacing are found at different process stages. These are due also to the moisture content,

26

with dry samples having a smaller D-spacing. This understanding is useful for relating

27

structural changes that occur during different stages of processing to the development of the

28

final physical characteristics of leather.

29 30

Keywords: Leather, Collagen, Small Angle X-ray Scattering; Orientation

31 32

2 ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

33

Journal of Agricultural and Food Chemistry

INTRODUCTION

34 35

The leather making process is a way of preserving skins to stop decomposition and to

36

provide a strong and flexible material. The process consists of a series of chemical treatments

37

and some mechanical processes. Each of the chemical treatments alters the composition of

38

the original skin, for example, extracting components from the native skin or adding

39

components such as cross–linking agents. In addition to changes in the chemistry of the

40

collagen, which are well known, it is possible that each of these processes may have an effect

41

on the collagen structure as well, although little information has been presented on this to

42

date.

43

The leather making process consists of eight main stages, which are referred to by a variety

44

of names in the industry, but here these are designated as fresh green, salted, pickled,

45

pretanned, wet blue, retanned, dry crust, and dry crust staked. The “fresh green” is the skin

46

after removal from the carcass. Skins are normally salted as a way of temporarily preserving

47

the skin before tanning. The salt acts to slow down bacterial growth by reducing water

48

activity. Salting causes some dehydration of the skins. The “salted” is the skin after salting

49

for preservation. The next stage of the leather making process is the removal of the salt by

50

soaking and washing the skins (where the skin becomes rehydrated) followed by an alkali

51

treatment carried out in the presence of sodium sulfide (‘liming’) combined with suitable

52

enzymes (‘bating’) which together break down and remove some of the non-fibrous proteins,

53

glycosaminoglycans and other undesirable components. It is also said to ‘open up’ the

54

structure of the leather to enable better penetration of tanning chemicals in subsequent stages.

55

After the alkaline treatment stage the skin is typically adjusted back to a lower pH and is

56

acidified in sulfuric acid and sodium chloride. After this stage the skin is referred to as

57

“pickled”. A synthetic, organic cross–linking agent and surfactant are often added at this 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

58

point which may assist with the subsequent chrome tanning stage. Stabilising agents are often

59

added to the pickle to raise the denaturation temperature of the collagen and thus enable skin

60

fat to be removed more efficiently at higher temperatures. The skins at this stage are called

61

“pretanned”. After the natural skin fats are removed the pretanned pelts are tanned using

62

chromium sulfate. The skin after chromium tanning is called “wet blue”. Chrome tanned

63

leather tends to be too rigid for most applications so there is normally a second tanning stage

64

using natural vegetable tannins or synthetic tannins to make the final leather feel softer and

65

‘fuller’. After this second tanning stage the skin is called “retanned”. At this stage dyes, fat

66

liquors, and modified fats or oils are added to complete the look and feel of the leather. After

67

this ‘fat liquoring’, the leather is dried and is called “dry crust”. The leather is then

68

mechanically softened or ‘staked’ and is referred to here as “dry crust staked”.

69

Previously it has been shown that small angle X–ray scattering (SAXS) can provide

70

detailed structural information on the amount of fibrous collagen, the microfibril orientation,

71

the D-spacing and the collagen fibril diameter in leather 1-3 and other tissues 4, 5.

72

The chemical treatments used to produce leather from skin and hide may result in changes

73

to the structure of the collagen fibrils and the arrangement of these fibrils. These chemical

74

processes include strong salt solutions, large changes in pH, enzymatic treatments to remove

75

cross–links, and new cross–links being formed. An overview of tanning chemistry has been

76

presented 6. Some aspects of chemical treatments of skins and their effect on structure have

77

been observed previously with liming of skins (increasing the pH with calcium hydroxide)

78

showing a decrease in the D-spacing of collagen 7. Dehydration of parchment has been shown

79

to result in a decrease in collagen D-spacing from 64.5 to 60.0 nm8. Pickling and retanning

80

agents have been shown to swell collagen fibres at low pH 9. At low ionic strength and non-

81

isoelectric pH charge dependent interactions (screening and selective ion adsorption) are

4 ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Journal of Agricultural and Food Chemistry

10

82

prevalent in maintaining the collagen architecture

which is also reflected in the greater

83

thermal stability of collagen at low pH 11 and in the elastic response of collagen 12.

84

In this work, we use SAXS to investigate the changes that take place in the microstructure

85

of leather through the different stages of processing from skin to leather. The structural basis

86

of these changes at the level of the collagen fibrils is not fully understood and forms the basis

87

of this investigation.

88 89

EXPERIMENTAL

90 91

An ovine pelt was obtained from 5-month-old, early season New Zealand Romney cross

92

lambs. Samples were removed from the same skin during several stages of processing at or

93

near the official sampling position13. These stages were termed fresh green, salted, pickled,

94

pretanned, wet blue, retanned, dry crust, and dry crust staked. Cross-sections were cut parallel

95

to the backbone and as close together as possible. The samples, except for dry crust and dry

96

crust staked, all had high moisture contents because these were taken during processing.

97

Conventional beamhouse and tanning processes were used to generate the leather. A pelt

98

was depilated using a caustic treatment then the residual keratinaceous material was removed.

99

The pelt was washed, pretanned and degreased using 4 % (w chemical/w limed leather for all

100

% chemical additions unless otherwise stated) Tetrapol LTN (Shamrock Group) and 2 %

101

Zoldine ZE (The Dow Company). Sodium formate was added followed by sodium

102

bicarbonate which was gradually added to increase the pH to 7.5-8. The pickled pelt was

103

then chrome tanned.

104

The resulting “wet blue” pelt was first neutralized, then washed and retanned using 2 %

105

w/w synthetic retanning agent (Tanicor PW, Clariant, Germany) and 3 % w/w vegetable

106

tanning material (mimosa, Tanac, Brazil). Fat liquor was added at a concentration of 6 % by 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 31

107

weight of wet leather and the leather processed at 50 °C for 45 min. Finally, the leather was

108

fixed by the addition of 0.5 % w/w formic acid and processed for 30 min followed by

109

draining and washing.

110

Tear strengths of the leathers were tested using standard methods 14, 15. Samples (strips 1 x

111

50 mm) were tested on an Instron 4467. “Dry” samples were conditioned at 23 °C and 50 %

112

relative humidity for 24 h before testing following the standard method

113

were in contact with free liquid up to the point at which they were tested.

16

. "Wet" samples

114

Diffraction patterns were recorded on the Australian Synchrotron SAXS/WAXS beamline,

115

utilizing a high-intensity undulator source. Energy resolution of 10-4 was obtained from a

116

cryo-cooled Si (111) double-crystal monochromator and the beam size (FWHM focused at

117

the sample) was 250 x 80 µm, with a total photon flux of about 2 x 1012 ph·s-1. All diffraction

118

patterns were recorded with an X-ray energy of 11 keV using a Pilatus 1M detector with an

119

active area of 170 x 170 mm and a sample-to-detector distance of 3371 mm. Exposure time

120

for diffraction patterns was 1 s and data processing was carried out using the SAXS15ID

121

software 17. The partially processed leather samples were sandwiched between kapton tape to

122

prevent drying and mounted for X-ray analysis.

123

A custom built stretching apparatus was built for in-situ SAXS measurements as described 18

124

elsewhere

. Strips of leather, 1 x 30 mm cut from the “official sampling position” (OSP),

125

were mounted horizontally without tension and without slack, with a measured separation

126

between the jaws, typically of 15 mm. SAXS patterns were taken through the full thickness

127

of the sample and the force and extension information recorded. The sample was stretched by

128

1 mm and was maintained at this extension for 1 minute before patterns were recorded. This

129

process was repeated with the sample stretched a further 1 mm each time until the sample

130

failed.

6 ACS Paragon Plus Environment

Page 7 of 31

Journal of Agricultural and Food Chemistry

131

Orientation index (OI) is defined as (90° – OA)/90° where OA is the azimuthal angle range

132

that contains 50% of the microfibrils centered at 180°. OI is used to give a measure of the

133

spread of microfibril orientation (an OI of 1 indicates the microfibrils are completely parallel

134

to each other; an OI of 0 indicates the microfibrils are completely randomly oriented). The OI

135

is calculated from the spread in azimuthal angle of the most intense D-spacing peak (at

136

around 0.059–0.060 Å–1) 19.

137 138

The D-spacing was determined from Bragg’s Law by taking the centre of a Gaussian curve fitted to the 6th order diffraction peak of an integrated intensity plot for each spectrum.

139 140

RESULTS

141 142

Orientation index changes. The OI of the corium region is higher than that of the grain

143

region throughout all the processing stages (Figure 1). While there is considerable variation

144

in the OI during processing, the OI of the final staked dry crust leather is fairly similar to that

145

of the fresh green skin so that overall there is not a large change in OI.

146

Orientation index and thickness. The thickness of the skin or leather varies by more than

147

a factor of two during processing. With this variation of thickness there is a change in OI

148

(Figure 2). The stages appear to be in two groups with the dry samples (salted, pretanned, dry

149

crust, and dry crust staked) in one group of lower OI which decreases with increasing

150

thickness, and the wet samples (pickled, wet blue, retanned, and fresh green) in another group

151

of higher OI that decreases with increasing thickness.

152

D-spacing with processing and pH. Significant changes are observed in the D-spacing of

153

the collagen fibrils with the application of each process step (Figure 3). Both the corium and

154

the grain regions of the leather have similar D-spacing and are similarly affected by the

155

processing. The most significant change in D-spacing occurs with pretanning when there is a 7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

156

contraction of the D-spacing of 1.4-1.5 nm (2%). The pH of the solutions in which each

157

process takes place does not appear to correlate with the D-spacing (Figure 4). For fresh

158

green and salted skins the pH was measured by a surface probe rather than the process liquor

159

as used in the later stages. Although there is not a correlation of D-spacing with pH, it

160

appears that there is a relationship between D-spacing and the moisture content. The three dry

161

samples have a lower D-spacing than the wet samples.

162

OI and pH. The OI appears to be correlated with pH (Figure 5) (Linear regression: r2 =

163

0.335, t = -1.74, P = 0.133). However, as discussed later in the paper we believe this

164

correlation is not a causal relationship.

165

D-spacing and OI. A correlation is observed between OI and D-spacing (Figure 6). The

166

correlation is statistically significant (Linear regression: r2 = 0.485, t = 2.4, P = 0.055).

167

However, as is discussed later in the paper we believe this correlation is not a causal

168

relationship.

169

Stress-strain. Stress-strain curves were measured in situ at the synchrotron during SAXS

170

data collection (Figure 7). The samples were not of a uniform width so an absolute

171

comparison of the modulus of elasticity from these curves is not possible. However, it is

172

possible in a very general way to compare the shapes of the curves. Most of the samples show

173

a toe region, with an initial lower elastic modulus, followed by a linear region.

174

OI and D-spacing with mechanical stretching. There is an increase in D-spacing as the

175

skin or leather samples are stretched (Figure 8). This increase in D-spacing is a measure of

176

the stress being transferred to the individual fibrils causing their length to increase. An

177

increase in OI on stretching is also found (Figure 9) as the collagen fibrils realign in the

178

direction of the applied force (values for the D-spacing and OI changes are tabulated in

179

Supplementary Information). At every stage of the leather processing, a strain applied to the

180

samples results in an increase in OI. 8 ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Journal of Agricultural and Food Chemistry

181 182

DISCUSSION

183 184

We have measured the changes to the structure and arrangement of collagen fibrils in skin

185

during several stages of the chemical processing to form leather. These build up a picture of

186

what happens in the process of transforming skin to leather.

187

Model of OI change with thickness. At the different stages of the processing of skin

188

through to leather the thickness of the skin varies, at times becoming thicker and at other

189

times thinner. The variation in thickness amounted to a factor of 2.25. If the collagen fibrils

190

are not lying flat then as the thickness of the leather increases the angle of these fibrils to the

191

plane of leather will increase (Figure 10). This will cause a change in OI. We have therefore

192

developed a model for the change in OI with thickness changes. This enables us to determine

193

how much of the OI change is due just to the change in thickness and how much is due to

194

other factors.

195

The range of fiber angles in a sample is given by the orientation index OI which is derived

196

from the orientation angle, OA1. OA is defined as the subtended angle that contains half the

197

fibrils, i.e. it satisfies eq 1.

198 θ =OA 2

199

0.5 =



θ Ndθ

θ = −OA 2 θ =90°



θ Ndθ

(1)

θ = − 90°

200

Where N is the number of fibrils, and θ is the fiber angle (relative to the plane of the

201

leather). In practice this is determined from the SAXS diffraction pattern by the integrated

202

intensity verses azimuthal angle for one of the D-spacing diffraction peaks, eq 2: 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

ϕ = OA 2

203

0.5 =



ϕ Idϕ



ϕ Idϕ

ϕ = −OA 2 ϕ =90°

(2)

ϕ = − 90°

204 205

where I is diffraction intensity of a selected D-spacing diffraction peak above a fitted background and φ is the azimuthal angle of the diffracted X-rays.

206

OI is derived from OA by eq 3.

207

OI =

208

Therefore, for perfect alignment in the reference direction, which is normally the plane of a

209

piece of leather or skin, OI = 1, for no alignment or completely isotropic fibrils OI = 0, and

210

for perfect alignment at right angles to the reference direction OI= –1.

1 − OA OA

(3)

211 212

If a sample of leather containing a fiber expands in thickness uniformly, and the fiber is at

213

an angle θ1 from the base (Figure 10), then the new angle of the fiber, θ2, depends on the

214

change in thickness by eq 4.

215

T2 tan θ 2 = T1 tan θ1

216

Where T1 is the original thickness, and T2 is the new thickness.

217

Rearranging eq 4 for θ2 gives the new angle of the fiber after the leather has increased in

218

(4)

thickness:

 T2  tan θ1   T1 

219

θ 2 = tan −1 

220

It is then possible to calculate a transformed OA after stretching. The OA defines the

221

subtended angle of 50% of the fibers. Therefore fibers which have an initial angle from the

(5)

10 ACS Paragon Plus Environment

Page 10 of 31

Page 11 of 31

Journal of Agricultural and Food Chemistry

222

reference angle of greater than half the OA will, after transformation, still have an angle

223

greater than the transformed OA; fibers which have an initial angle lower than half the OA

224

will, after transformation, still have an angle lower than half the transformed OA. In other

225

words the 50% inside the OA will remain inside the OA, the 50% outside the OA will remain

226

outside the OA. Therefore, it is only necessary to calculate the transformation by eq 5 of the

227

angle that is the OA to calculate the new OA that represents the new fiber distribution.

228

As an illustration of the extent of this change, for thickness changes similar to those

229

observed here, some values are listed in Table 1. Note that expanding the thickness

230

sufficiently results in an alignment of fibers vertically (negative OI).

231

We did also consider how the OI might be affected by the area change to the leather, with

232

wet stages generally of larger area than dry stages. This expansion of area would influence

233

the OI in the opposite sense to the thickness change. However the area change of the skin is

234

small, less than 5%, and can be neglected compared with the more than 100% change in the

235

thickness.

236

With the thickness taken into account the OI now reflects any underlying structural changes

237

that take place during the chemical and mechanical processing of skin through to leather. The

238

OI has therefore been recalculated for each sample using the thinnest sample, the salted skin,

239

as the reference and the adjusted results are presented in the following analysis.

240

OI and water. Once the effect of thickness on the OI is removed it can be seen that the

241

samples fall into two groups (Figure 11). There are the fresh green, pickled, wet blue, and

242

retanned materials. These all have a high OI (corrected for thickness relative to salted) of

243

around 0.8. These samples are all wet. The pretanned, dry crust, and dry crust staked all have

244

a corrected OI of around 0.67. These samples are all dry. The salted skin also has a low OI, of

245

0.62, and it is also a dry sample. Therefore the factor that separates the high OI from the low

246

OI materials is the wetness of the samples. When these collagen materials are dried, the OI 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 31

247

decreases. While it might be tempting to ascribe this to a crimping of the collagen fibrils as

248

the materials dry, which would result in a lower OI, crimping in not observed in skin (unlike

249

in tendon or pericardium

250

effect, we have been able to identify that there is a fundamental change between wet and dry

251

stages we have not determined the mechanism of this change. It is suggested that a possible

252

mechanism could be the space filling that results from hydration creates a hindrance to the

253

movement of fibrils and leads to a higher OI at higher moisture contents. At lower moisture

254

contents there is more space available for fibers to bend and adjust themselves into positions

255

resulting in lower OI.

20

for example). Therefore, although by removing the thickness

256

D-spacing and OI. Once the OI is corrected for thickness effects there is no longer a

257

statistically significant correlation between D-spacing and OI (Supporting information Figure

258

S1) (Linear regression: r2 = 0.109, t = 0.855, P = 0.426). It is known that the result of

259

straining leather is a straightening or alignment of the fibrils (an increase in OI) and that D-

260

spacing measures the stress applied to individual fibrils 18. A correlation between OI and D-

261

spacing during different stages of processing would have suggested that these were both due

262

to changes in internal strain in leather: when strain is released D-spacing decreases and OI

263

decreases. However, this was not observed here.

264

D-spacing and water. It was found that the D-spacing is lower in the dry samples by 0.38

265

nm; wet samples, which include fresh green, pickled, wet blue, and retanned, have an average

266

D-spacing of 64.90 (σ = 0.46) nm while dry samples, which include salted, pretanned, dry

267

crust, and dry crust staked, have an average D-spacing of 64.52 (σ = 0.53) nm. It has

268

previously been shown by X-ray diffraction and by atomic force microscopy that the D-

269

spacing is moisture dependant with a reduction in D-spacing on drying

270

believed to be due to the collapse of the gap and overlap regions, and the partial shearing of

271

unit cell contents within the gap region upon loss of water 21. 12 ACS Paragon Plus Environment

8, 21-24

. This is

Page 13 of 31

Journal of Agricultural and Food Chemistry

272

D-spacing and pH. It is perhaps a little surprising that D-spacing is not affected by pH.

273

The D-period is affected by both the length of the tropocollagen units, and by the way these

274

tropocollagen units are assembled into the collagen fibril. The length of these structures is

275

partially determined by the length of the hydrogen bonds between collagen molecules within

276

the tropocollagens and by hydrophobic interactions between tropocollagen units. pH can

277

affect hydrogen bonding which, in collagen and other proteins, relies on a polar interaction

278

between specific functional groups on amino acids 10, 11. Mechanisms for fibril elongation by

279

chemical treatments have previously been discussed 20, 25 and it appears that pH does not have

280

an impact here.

281

OI and pH. Once the OI is corrected for thickness it is apparent that there is no correlation

282

between OI and pH (Supporting information Figure S2) (linear regression: regression

283

coefficient = -0.0215, r2 = 0.488, P = 0.22). These instead fall into two groups: low OI for dry

284

samples and high OI for wet samples, irrespective of the pH conditions.

285

Response to strain. We found that when strain was applied to the samples both the OI and

286

D-spacing increased. It has previously been shown that upon stretching of leather collagen

287

fibrils first realign and then lengthen (an increase in the OI followed by an increase in the D-

288

spacing)

289

pretanned, dry crust, and dry crust staked samples, the OI increased first after the first few

290

increments in strain then the D-spacing increased with the further strain. These samples form

291

a cluster at an OI of approximately 0.67 following thickness correction. The D-spacing of the

292

wet samples (fresh green, pickled, wet blue, and retanned) increased with the first application

293

of strain. These wet samples form a cluster with an OI of about 0.80. It is possible for the dry

294

group to undergo more realignment with strain than the wet group because the dry samples

295

have a lower OI initially and therefore more scope for change. Therefore in the wet samples,

18

and that collagen fibrils are very resilient with a high Poisson’s ratio

13 ACS Paragon Plus Environment

26

. In

Journal of Agricultural and Food Chemistry

Page 14 of 31

296

because less realignment is possible, the strain was taken on by the individual collagen fibrils

297

sooner, as shown by the immediate increase in D-spacing.

298

Cross–linking. In the process of treating skin to produce leather, naturally present cross–

299

links are removed and new cross–links are formed. The natural glycosaminoglycan cross–

300

links are removed in the liming and bating stages so that the processed skin, known as

301

pickled, should have fewer cross–links. New cross–links are added at the pretanned and wet

302

blue stages. While there have been studies on the effects of cross–linking on fibril alignment,

303

there have been mixed conclusions drawn about the influence of cross–linking. It has been

304

suggested that cross–links may stabilize a network structure (with less aligned fibrils)

305

and that removal of cross–links may destabilize the network structure leading to fibrils

306

becoming more aligned

307

where some of the cross linking would be expected to be removed by the end of the pickled

308

stage and then cross links are added by the completion of the wet blue stage.

27, 28

28

. However such behavior is not observed here during processing,

309

We have found that although processing results in changes in OI, this is not a fundamental

310

redistribution of fibrils, but is rather a decrease in OI which occurs as the thickness increases

311

and as also (independently of thickness) as the water content increases. This understanding

312

may be useful for relating structural changes that occur during different stages of processing

313

to the development of the final physical characteristics of leather. For example, if fibril

314

orientation is critical to strength in certain leathers, the structural relationship that should

315

exist between the fresh skin and the final leather is now known and it is possible to seek to

316

maintain this characteristic during processing or modify it appropriately.

317 318

Acknowledgements

14 ACS Paragon Plus Environment

Page 15 of 31

Journal of Agricultural and Food Chemistry

319

This research was undertaken on the SAXS/WAXS beamline at the Australian

320

Synchrotron, Victoria, Australia. The New Zealand Synchrotron Group provided travel

321

funding. The work was supported by a grant from the Ministry of Business, Innovation, and

322

Employment. Sue Hallas of Nelson assisted with editing.

323 324

Supporting Information Available

325

Detailed tables of D-spacing changes in processing stages of leather with strain, OI changes

326

in processing stages of leather with strain OI, and the variation of D-spacing with OI and pH

327

are available as supporting information. This information is available free of charge via the

328

Internet at http: //pubs.acs.org.

329 330

REFERENCES

331

1.

Basil-Jones, M. M.; Edmonds, R. L.; Allsop, T. F.; Cooper, S. M.; Holmes, G.;

332

Norris, G. E.; Cookson, D. J.; Kirby, N.; Haverkamp, R. G., Leather structure determination

333

by small angle X-ray scattering (SAXS): Cross sections of ovine and bovine leather. J. Agric.

334

Food Chem. 2010, 58, 5286-5291.

335

2.

336

stretching of leather on fibre orientation and tensile modulus. J. Mater. Sci. 2004, 39, 2481-

337

2486.

338

3.

339

or X-ray-diffraction and quantitative relation to mechanical-properties. J. Am. Leather Chem.

340

As. 1986, 81, 221-230.

Sturrock, E. J.; Boote, C.; Attenburrow, G. E.; Meek, K. M., The effect of the biaxial

Kronick, P. L.; Buechler, P. R., Fiber orientation in calfskin by laser-light scattering

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

341

4.

Sasaki, N.; Odajima, S., Stress-strain curve and youngs modulus of a collagen

342

molecule as determined by the x-ray diffaction technique. J. Biomech. 1996, 29, 655-658.

343

5.

344

intact planar connective tissues under biaxial stretch. Acta Biomater. 2005, 1, 45-54.

345

6.

Covington, A. D., Modern tanning chemistry. Chem.Soc. Rev. 1997, 26, 111-126.

346

7.

Maxwell, C. A.; Wess, T. J.; Kennedy, C. J., X-Ray Diffraction Study into the Effects

347

of Liming on the Structure of Collagen. Biomacromolecules 2006, 7, 2321-2326.

348

8.

349

treatment and relation to long term deterioration. Thermochimica Acta 2000, 365, 119-128.

350

9.

351

fibrils: Interaction with pickling and retanning agents. Macromol. Biosci. 2007, 7, 234-240.

352

10.

353

interaction. Biopolymers 2008, 89, 700-709.

354

11.

355

Hydrogen Bonds of Collagen Triple Helical Structure: Influence of pH and Chaotropic

356

Nature of Three Anions. Matrix Biol. 2000, 19, 511-520.

357

12.

358

Modulus of Hydrated Collagen Fibrils. Biophys. J. 2009, 97, 2985-2992.

359

13.

360

location. In ISO: Switzerland, 2002; ISO 2418:2002.

361

14.

362

edge tear. J. Soc. Leather Tech. Ch. 2000, 84, 327-329.

363

15.

364

2: Double edge tear. In ISO: Switzerland, 2002; BS EN ISO 3377-2:2002(E).

Liao, J.; Yang, L.; Grashow, J.; Sacks, M. S., Molecular orientation of collagen in

Wess, T. J.; Orgel, J. P., Changes in collagen structure: drying, dehydrothermal

Bulo, R. E.; Siggel, L.; Molnar, F.; Weiss, H., Modeling of bovine Type-I collagen

Ciferri, A., Charge-dependent and charge-independent contributions to ion-protein

Zanaboni, G.; Rossi, A.; Onana, A. M. T.; Tenni, R., Stability and Networks of

Grant, C. A.; Brockwell, D. J.; Radford, S. E.; Thomson, N. H., Tuning the Elastic

IULTCS, Leather - Chemical, physical and mechanical and fastness tests - Sampling

Williams, J. M. V., IULTCS (IUP) test methods - Measurement of tear load-double

IULTCS, Leather - Physical and mechanical tests - Determination of tear load - Part

16 ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31

Journal of Agricultural and Food Chemistry

365

16.

IULTCS, Leather - Physical and mechanical tests - Sample preparation and

366

conditioning. In ISO: Switzerland, 2012; ISO 2419:2012.

367

17.

368

and calibration with a pinhole-geometry SAXS instrument on a synchrotron beamline. J.

369

Synchrotron Radiat. 2006, 13, 440-444.

370

18.

371

alignment and deformation during tensile strain of leather: A SAXS study. J. Agric. Food

372

Chem. 2012, 60, 1201-1208.

373

19.

374

fibril orientation in ovine and bovine leather affects strength: A small angle X-ray scattering

375

(SAXS) study. J. Agric. Food Chem. 2011, 59, 9972-9979.

376

20.

377

Haverkamp, R. G., Age Dependant Differences in Collagen Fibril Orientation of

378

Glutaraldehyde Fixed Bovine Pericardium. BioMed Res. Int. 2014, 2014, 189197.

379

21.

380

microfibrils: Effects of tissue extension on axial and lateral packing. J. Structural Biol. 1998,

381

122, 123-127.

382

22.

383

tissue hydration on nanoscale structural morphology and mechanics of individual Type I

384

collagen fibrils in the Brtl mouse model of Osteogenesis Imperfecta. J. Structural Biol. 2012,

385

180, 428-438.

386

23.

387

Biochim. Biophys. Acta 1980, 621, 162-166.

Cookson, D.; Kirby, N.; Knott, R.; Lee, M.; Schultz, D., Strategies for data collection

Basil-Jones, M. M.; Edmonds, R. L.; Norris, G. E.; Haverkamp, R. G., Collagen fibril

Basil-Jones, M. M.; Edmonds, R. L.; Cooper, S. M.; Haverkamp, R. G., Collagen

Sizeland, K. H.; Wells, H. C.; Higgins, J. J.; Cunanan, C. M.; Kirby, N.; Hawley, A.;

Wess, T. J.; Purslow, P. P.; Kielty, C. M., X-ray diffraction studies of fibrillin-rich

Kemp, A. D.; Harding, C. C.; Cabral, W. A.; Marini, J. C.; Wallace, J. M., Effects of

Brodsky, B.; Eikenberry, E. F.; Cassidy, K., An unusual collagen periodicity in skin.

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

388

24.

Price, R. I.; Lees, S.; Kirschner, D. A., X-ray diffraction analysis of tendon collagen

389

at ambient and cryogenic temperatures: Role of hydration. Int. J. Biol. Macromol. 1997, 20,

390

23-33.

391

25.

392

Mudie, S.; Haverkamp, R. G., Collagen D-spacing and the Effect of Fat Liquor Addition. J.

393

Am. Leather Chem. Assoc. 2015, 110.

394

26.

395

Haverkamp, R. G., Poisson's Ratio of Collagen Fibrils Measured by SAXS of Strained

396

Bovine Pericardium. J. Appl. Phys. 2015, 117, 044701.

397

27.

398

oxazolidine E on collagen fibril formation and stabilization of the collagen matrix. J. Agric.

399

Food Chem. 2007, 55, 6813-6822.

400

28.

401

G., Collagen Cross Linking and Fibril Alignment in Pericardium. RSC Adv. 2015, 5, 3611-

402

3618

Sizeland, K. H.; Wells, H. C.; Norris, G. E.; Edmonds, R. L.; Kirby, N.; Hawley, A.;

Wells, H. C.; Sizeland, K. H.; Kayed, H. R.; Kirby, N.; Hawley, A.; Mudie, S. T.;

Deb Choudhury, S.; Haverkamp, R. G.; DasGupta, S.; Norris, G. E., Effect of

Kayed, H. R.; Sizeland, K. H.; Kirby, N.; Hawley, A.; Mudie, S. T.; Haverkamp, R.

403 404

18 ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31

Journal of Agricultural and Food Chemistry

Figure 1. Variation in orientation index for all stages of processing: (▲, ─ ─ ─ ─, red) corium, (■, ·········, blue) grain, (●, ______, black) average. 74x76mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2. Fibril orientation index versus thickness. 56x43mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

Journal of Agricultural and Food Chemistry

Figure 3. Variation in D-spacing and pH between different stages of processing prior to stretching: (∆, ─ ─ ─ ─) corium, (□, ·········) grain, (●, ______, red ) pH. 66x61mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 4. Variation of D-spacing with pH (wet blue and retanned points are coincident). 54x41mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31

Journal of Agricultural and Food Chemistry

Figure 5. Correlation of OI with pH. 54x41mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 6. D-spacing versus OI for samples of different stages of processing when held without tension. 56x44mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31

Journal of Agricultural and Food Chemistry

Figure 7. Stress strain curves for each process stage, performed in situ concurrently with the SAXS measurement (not normalized for sample width or thickness): (●, ______ ) Fresh green, (●, ·······) salted, (▼, ─ ─ ─ ─) pickled, (▲, ─ ·· ─ ·· ─) pretanned, (■,― ― ) wet blue, (■, ─ · ─ · ─) retanned, (♦,― ― ―) dry crust, (♦,______) dry crust staked. 57x45mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 8. Changes in collagen D-spacing as samples of partially processed skin are stretched: (●, ______ ) Fresh green, (●, ·······) salted, (▼, ─ ─ ─ ─) pickled, (▲, ─ ·· ─ ·· ─) pretanned, (■,― ― ) wet blue, (■, ─ · ─ · ─) retanned, (♦,― ― ―) dry crust, (♦, ______ ) dry crust staked. 54x41mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

Journal of Agricultural and Food Chemistry

Figure 9. Changes in collagen fibril OI as samples of partially processed skin are stretched: (●, ______ ) Fresh green, (●, ·······) salted, (▼, ─ ─ ─ ─) pickled, (▲, ─ ·· ─ ·· ─) pretanned, (■,― ― ) wet blue, (■, ─ · ─ · ─) retanned, (♦,― ― ―) dry crust, (♦,______ ) dry crust staked. 57x46mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 10. Illustration of the change in collagen fibril angle θ1 to the plane of leather as the thickness T1 of the leather changes. 78x38mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31

Journal of Agricultural and Food Chemistry

Figure 11. Fibril orientation index versus thickness: (▲) measured OI, (■, red) calculated OI adjusted for thickness changes (relative to salted). 56x43mm (600 x 600 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table 1. Calculated change in orientation index of collagen fibrils for different thickness of material. 153x153mm (72 x 72 DPI)

ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31

Journal of Agricultural and Food Chemistry

TOC Graphic 352x129mm (72 x 72 DPI)

ACS Paragon Plus Environment