Effects of muscle-specific oxidative stress on ... - ACS Publications

Aug 10, 2017 - Effects of muscle-specific oxidative stress on cytochrome c release and ... of beef longissimus lumborum (LL) and psoas major (PM) musc...
0 downloads 0 Views 427KB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Effects of muscle-specific oxidative stress on cytochrome c release and oxidation-reduction potential properties Yiling Ke, Rachel M Mitacek, Anupam Abraham, Gretchen G. Mafi, Deborah L. VanOverbeke, Udaya DeSilva, and Ranjith Ramanathan J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b01735 • Publication Date (Web): 10 Aug 2017 Downloaded from http://pubs.acs.org on August 10, 2017

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 28

Journal of Agricultural and Food Chemistry

1 2 3

Effects of muscle-specific oxidative stress on cytochrome c release and oxidation-reduction

4

potential properties

5 6 7

Yiling Ke, Rachel M. Mitacek, Anupam Abraham, Gretchen. G. Mafi, Deborah L. VanOverbeke, Udaya DeSilva, Ranjith Ramanathan*

8 9 10 11 12 13 14 15 16



Department of Animal Science, Oklahoma State University, Stillwater, OK 74078, USA

17

* Corresponding author. TEL: +1 405-744-9260; Fax: +1 405-744-4716.

18

EMAIL: [email protected] (R. Ramanathan).

19

20

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 28

21

ABSTRACT

22

Mitochondria play a significant role in beef color. However, the role of oxidative stress in

23

cytochrome c release and mitochondrial degradation is not clear. The objective was to determine

24

the effects of display time on cytochrome c content and oxidation-reduction potential (ORP) of

25

beef longissimus lumborum (LL) and psoas major (PM) muscles. PM discolored by day 3

26

compared with LL. On day 0, mitochondrial content and mitochondrial oxygen consumption

27

were greater in PM than LL. However, mitochondrial content and oxygen consumption were

28

lower (P < 0.05) in PM than LL by day 7. Conversely, cytochrome c content in sarcoplasm was

29

greater on days 3 and 7 for PM than LL. There were no significant differences in ORP for LL

30

during display, but ORP increased for PM on day 3 when compared with day 0. The results

31

suggest that muscle-specific oxidative stress can affect cytochrome c release and ORP changes.

32

33

Keywords: Beef color, myoglobin, metmyoglobin reducing activity, mitochondria, oxidation-

34

reduction potential

35

36

37

2 ACS Paragon Plus Environment

Page 3 of 28

38

Journal of Agricultural and Food Chemistry

INTRODUCTION

39

Consumers associate the bright-red color of beef with freshness and wholesomeness.

40

Myoglobin is the primary sarcoplasmic protein responsible for beef color. Depending on the

41

redox state of the heme and the ligand bound, myoglobin exists in three forms.1 Ferrous heme

42

with no ligand attached results in purple color due to predominant deoxymyoglobin, while

43

oxymyoglobin gives the characteristic bright-red color. The oxidation of both oxymyoglobin and

44

deoxymyoglobin leads to metmyoglobin formation or discoloration. Approximately 15% of meat

45

is discounted in price due to discoloration.2 Hence, characterizing the factors associated with

46

myoglobin oxidation is important to minimize the losses due to discoloration.

47

Metmyoglobin accumulation is muscle-specific and depends on the inherent muscle

48

biochemistry. Longissimus lumborum (LL) is classified as a color-stable muscle while psoas

49

major (PM) is a color-labile muscle.3-5 Among the various factors that affect color stability, the

50

role of mitochondria has been investigated by numerous researchers.6-9 Mitochondrial activity

51

can influence myoglobin redox state via oxygen consumption and metmyoglobin reduction.6,7,9

52

However, oxidative processes that affect mitochondrial function can impact beef color.10

53

Immediately after animal harvest, blood flow ceases, and the muscle metabolism changes

54

from aerobic to anaerobic. This can significantly affect muscle homeostasis, increase oxidative

55

stress,11 and degeneration of mitochondria.12 In healthy cells, cytochrome c is the major electron

56

carrier between complex III and IV of the electron transport chain. The level of cytochrome c

57

content in sarcoplasm is an indicator of mitochondrial damage or outer membrane breakdown.13

58

The concentration of mitochondria is greater in type I (predominant red fibers) muscles

59

than in type II (white fibers). The proportion of red and white fibers varies between different

60

muscles in a beef carcass. PM muscle has greater red fiber content than LL muscle. Thus, it is

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 28

61

possible that the oxidative changes will be greater in PM than LL muscles. However, limited

62

studies have determined the effects of muscle type and display time on cytochrome c level in

63

relation to beef color. The oxidation-reduction potential (ORP) measurements can predict

64

oxidative stress in tissue.14 Hence, the quantification of ORP changes with display time will

65

provide a better understanding of the muscle-specific differences in oxidative stress. Therefore,

66

the objective of this study was to determine the effects of display time and muscle type (beef LL

67

and PM) on changes in mitochondria, cytochrome c content, and ORP.

68

MATERIALS AND METHODS

69

Raw materials and processing

70

Eight USDA paired Select beef LL (IMPS #180)15 and PM from market age animals were

71

collected within 48 h of harvest from a commercial packing plant. The muscles were transported

72

on ice to the Food and Agricultural Products Center at Oklahoma State University. Following

73

overnight storage, five 2.5-cm-thick steaks were cut from the anterior end of each muscle and

74

randomly assigned to days 0, 1, 3, 5, and 7. Steaks were placed (N = 16 steaks on each day; n = 8

75

from LL and n = 8 from PM) on foam trays with absorbent pads, over-wrapped with a PVC film

76

(oxygen-permeable polyvinyl chloride fresh meat film; 15,500–16,275 cm3 O2/m2/24 h at 23 °C,

77

E-Z Wrap Crystal Clear Polyvinyl Chloride Wrapping Film, Koch Supplies, Kansas City, MO).

78

After packaging, steaks were placed in a coffin-style open display case maintained at 2 °C ± 1

79

under continuous lighting (1612 to 2152 lx, Philips Deluxe Warm White Fluorescent lamps;

80

Andover, MA; color rendering index = 86; color temperature = 3000 K). All packages were

81

rotated daily to minimize the variances in light intensity or temperature caused by the location.

82 83

Following surface color measurements, steaks designated to days 0, 3, and 7 were cut in half. The first half was used to measure metmyoglobin reducing activity (MRA) and muscle 4 ACS Paragon Plus Environment

Page 5 of 28

Journal of Agricultural and Food Chemistry

84

oxygen consumption, and the second half was used to quantify ORP, cytochrome c content, pH,

85

lipid oxidation, and for mitochondrial studies. The steak half assigned to MRA and muscle

86

oxygen consumption was then bisected parallel to the oxygenated surface to expose the interior

87

of the steak (resulting in two interior pieces). The first piece was used to measure MRA and the

88

second piece was used to measure muscle oxygen consumption. Before taking samples from the

89

second half, a pH probe was used to measure muscle pH. Following pH measurements from the

90

second half, approximately 50 g of muscle samples visually devoid of fat and connective tissue

91

were vacuum packaged and assigned for mitochondrial studies. Another 10 g of the sample that

92

contained displayed surface and interior was used for lipid oxidation measurements. Two 5 g

93

samples were collected for cytochrome c and ORP measurements, vacuum packaged and kept at

94

-80 °C until use. Mitochondrial studies, cytochrome c, pH, MRA, muscle oxygen consumption,

95

and ORP measurements were conducted on days 0, 3, and 7. Fresh tissue samples were used for

96

oxygen consumption, mitochondrial quantification using protein content, and lipid oxidation

97

studies; however frozen tissue was used for mitochondrial quantification using DNA-method.

98

Proximate and pH analysis

99

Proximate analysis was conducted using an AOAC-approved (Official Method 2007.04)

100

near-infrared spectrophotometer (FOSS Food Scan™ 78800; Dedicated Analytical Solutions,

101

DK-3400 Hillerod, Denmark). Compositional values were reported on a percent (%) basis. pH

102

was determined by inserting the pH probes at four different locations within a steak using a

103

Mettler Toledo SevenGo pH meter.

104

Surface color

105

Surface color was measured on days 0, 1, 3, 5, and 7 at three locations using a

106

HunterLab MiniScan XE Plus spectrophotometer (HunterLab Associates, Reston, VA, USA) 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

107

with a 2.5-cm diameter aperture, Illuminant A, and 10° standard observer. Reflectance at

108

isobestic wavelengths from 400 – 700 nm was used to quantify myoglobin redox forms on the

109

surface of steaks. Reflectance at 474, 525, 572, and 610 nm was converted to K/S values using:

110

K/S = (1-R)2 / 2R. These values were then substituted into following equations1 to calculate the

111

percentage of oxy- and metmyoglobin.

112

% oxymyoglobin = [K/S610÷K/S525 for 100% metmyoglobin - K/S610÷K/S525 for sample]

113

114 115

116

Page 6 of 28

[K/S610÷K/S525 for 100% metmyoglobin - K/S610÷K/S525 for 100% oxymyoglobin]

% metmyoglobin = [K/S572÷K/S525 for 100% oxymyoglobin - K/S572÷K/S525 for sample] [K/S572÷K/S525 for 100% oxymyoglobin - K/S572÷K/S525 for 100% metmyoglobin]

K and S indicate absorption and scattering coefficient, respectively.

117

Myoglobin reference standards were prepared using the extra steaks not used in the study.

118

Percentage myoglobin values were also used to calculate muscle oxygen consumption and MRA.

119

Muscle oxygen consumption

120

Muscle oxygen consumption was determined indirectly as the changes in oxymyoglobin

121

level. In the current study, a modified procedure of Madhavi et al.16 was used to determine

122

muscle-specific oxygen consumption on days 0, 3, and 7. Following surface color measurements,

123

the steak was bisected parallel to the oxygenated surface to expose the interior of steak, resulting

124

in two interior pieces. The first interior (the section that was exposed to air) was used for oxygen

125

consumption studies and the second interior was used for MRA studies. Freshly cut LL or PM

126

muscle samples were allowed to oxygenate for 60 min at 4 ˚C, vacuum packaged, and scanned

127

twice on the bloomed surface to measure oxymyoglobin. Oxygen consumption (measured by

128

conversion of oxymyoglobin to deoxymyoglobin) was induced by incubating samples at 30 ˚C

6 ACS Paragon Plus Environment

Page 7 of 28

Journal of Agricultural and Food Chemistry

129

for 30 min. The surface color was scanned using a Hunter Lab Miniscan XE Plus

130

spectrophotometer for reflectance from 400 to 700 nm to determine the initial % oxymyoglobin.

131

Samples were rescanned immediately to determine remaining surface oxymyoglobin.1 To

132

calculate % muscle oxygen consumption, the following equation was used: (% pre-incubation

133

surface oxymyoglobin - % post-incubation surface oxymyoglobin).

134

Metmyoglobin reducing activity

135

Metmyoglobin reducing activity was determined on days 0, 3, and 7 of display.17

136

Samples from the interior of steak halves were submerged in a 0.3% solution of sodium nitrite

137

(Sigma, St. Louis, MO) for 20 min to facilitate metmyoglobin formation, and then removed,

138

blotted dry, vacuum packaged (Prime Source Vacuum Pouches, 4 mil, Koch Supplies Inc.,

139

Kansas City, MO), and scanned with a HunterLab MiniScan XE Plus spectrophotometer to

140

determine pre-incubation metmyoglobin values.1 Each sample was incubated at 30 °C for 2 h to

141

induce metmyoglobin reduction. Upon removal from the incubator, samples were rescanned to

142

determine the percentage of remaining surface metmyoglobin. The following equation was used

143

to calculate %MRA: [(% surface metmyoglobin pre-incubation - % surface metmyoglobin post-

144

incubation).

145

Mitochondrial content quantification

146

Mitochondrial content in LL and PM muscles during storage were quantified as total

147

yield and as mitochondrial DNA.

148

Total mitochondrial content in longissimus and psoas muscles

149 150

A differential centrifugation method was used to isolate mitochondria from LL and PM muscles.8 Mitochondrial protein content was determined using a bicinchoninic acid protein

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 28

151

assay, and reported as mg/g of tissue.18 Briefly, 10 g of minced tissue devoid of fat and

152

connective tissue was washed twice with 250 mM sucrose and suspended in 20 mL of

153

mitochondrial isolation buffer [250 mM sucrose, 10 mM HEPES, 1 mM EDTA, and 0.1% BSA,

154

pH 7.2]. The suspension was stirred slowly and hydrolyzed with protease (protease/tissue, 0.5

155

mg/g of tissue) for 20 min; the pH was maintained between 7.0 and 7.2. After proteolytic

156

digestion, the suspension was diluted to 300 mL with mitochondrial isolation buffer and

157

homogenized using a Kontes Duall grinder (Vineland, NJ) followed by a Wheaton Potter-

158

Elvehjem grinder (Millville, NJ). The homogenate was centrifuged at 900 x g for 20 min with a

159

Sorvall refrigerated RC-5B centrifuge (Thermo Fisher Scientific, Waltham, MA), and the

160

resulting supernatant was again centrifuged at 14,000 x g for 15 min. Mitochondrial pellets were

161

washed twice and suspended in mitochondrial suspension buffer (250 mM sucrose, 10 mM

162

HEPES, pH 7.2). All steps were performed between 0 - 4 °C.

163

Quantification of mitochondrial DNA

164

A 274 bp amplicon of bovine mitochondrial cytochrome b gene (gb|JX472274.1|;

165

forward primer – 5’GACCTCCCAGCTCCATCAAACATCTCATCTTGATGAAA3’, reverse

166

primer – 5’CTAGAAAAAGTGTAAGACCCGTAATATAAG3’) was used to evaluate

167

mitochondrial copy number19 and a 119 bp amplicon based on the 18S rRNA gene (forward

168

primer - 5’TTCGAACGTCTGCCCTATTAA3’, reverse primer –

169

5’GATGTGGTAGCCGTTTCTCAGG3’) was used as the normalizer. Total DNA was isolated

170

using a QIAamp DNA Mini Kit (Qiagen, Valencia, CA) and manufacturer’s instructions with

171

following modifications. Approximately, 0.3 g of the ground muscle in liquid nitrogen was used

172

for further processing. The PCR assay was validated first by assessing the primer efficiencies via

173

performing a standard curve analysis using 7-log dilutions of template DNA. The specificity of

8 ACS Paragon Plus Environment

Page 9 of 28

Journal of Agricultural and Food Chemistry

174

the assay was determined by ensuring a single amplicon and by sequence analyzing several

175

amplicons to confirm that they represent cytochrome b gene. Paired LL and PM muscles were

176

assayed for mitochondrial cytochrome b gene and 18S rRNA using an SYBR green1 reporter

177

assay (Roche Diagnostics, Indianapolis, IN). The reaction mixture contained forward and reverse

178

primer, master mix (Roche Diagnostics, Indianapolis, IN), and template DNA. Cycling

179

conditions were; 95 °C for 10 min followed by 50 cycles of 94 °C for 15 s, 58 °C for 15 s, and

180

72 °C for 20 s. The relative mitochondrial copy number was calculated using the comparative Ct

181

method as previously described.20 For the two muscle types, the mitochondrial cytochrome b Ct

182

was normalized to the 18S gene. The validity of the 18S rRNA gene as a normalizer was

183

ascertained by ensuring that the Ct values for each sample for the normalizer fell within ±1

184

amplification cycle.

185

Mitochondrial oxygen consumption

186

Mitochondrial oxygen uptake was measured using a Clark oxygen electrode6,7 (polarizing

187

voltage of 0.6 V and an 8 mL incubation chamber). The reaction components were added to the

188

incubation chamber, which was maintained at 25 °C by a water jacket and Lauda RE120

189

circulating water bath (Westbury, NY). The chamber was stirred with a 10 mm Teflon-covered

190

bar at 600 rpm. The electrode was attached to a Rank Brothers digital model 20 oxygen

191

controller (Cambridge, England) and connected to a computer and data logger. Oxygen

192

consumption was recorded over time at pH 5.6 by suspending mitochondria in incubation buffer

193

(250 mM sucrose, 5 mM KH2PO4, 5 mM MgCl2, 0.1 mM EDTA, 0.1% BSA, and 20 mM maleic

194

acid) with the addition of succinate (10 mM) as a substrate. The oxygen consumption rate was

195

calculated based on the method of Estabrook.21

196 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

197 198

Page 10 of 28

Quantification of cytochrome c The cytochrome c content in sarcoplasm was quantified using a sandwich-ELISA assay

199

per manufacturer’s instructions with following modifications (Abcam, Cambridge, UK). To

200

measure cytochrome c content in sarcoplasm, 100 mg of either LL or PM muscle was

201

homogenized in 900 µL of buffer (10 mM HEPES, pH 7.2) using a Wheaton Potter-Elvehjem

202

tissue grinder. The homogenates were centrifuged at 900 x g for 10 min. The supernatant was

203

then centrifuged again at 14,000 x g for 10 min. Mitochondria sediment at 14,000 x g, and the

204

supernatant was used to determine cytochrome c content in the sarcoplasm. The protein

205

concentration of the supernatant was determined using a BCA assay. The samples were

206

transferred to a 96-well microplate and were incubated for 3 h. Following incubation, the plate

207

was washed using TRIS washing buffer. Cytochrome c detector bovine specific antibody

208

solution was added to each well before another one-hour incubation time. The liquid was

209

discarded, and the plate was washed again. The horseradish peroxide solution was added, and the

210

microplate was incubated for one hour. The plate was washed five times before 200 µL

211

development solution was added to each well. The absorbance of each well in the microplate was

212

measured at 600 nm every 90 s for 30 min at room temperature using a kinetic program. A

213

standard curve was plotted with known cytochrome c and expressed as µg cytochrome c in per

214

gram tissue.

215

Oxidation-reduction potential

216

A RedoxSys analyzer was used to measure the ORP status in muscles. This machine

217

measures the balance of oxidants and reductants, providing a direct measurement of oxidative

218

stress using electrochemical principles.22 Briefly, 100 mg meat sample from each muscle type on

219

days 0, 3, and 7 after postmortem was homogenized in 50 mM phosphate buffer at pH 7.4 using

10 ACS Paragon Plus Environment

Page 11 of 28

Journal of Agricultural and Food Chemistry

220

a handheld homogenizer. The homogenate was centrifuged at 12,000 x g for 5 min. 20 µL of the

221

supernatant was applied on the sensor attached to RedoxSys analyzer reader (Aytu BioScience,

222

Inc, Englewood, CO) and ORP status was recorded as millivolt (mV). A lower number

223

represents a higher antioxidant status. This technique has been used to measure oxidative stress

224

status in serum and plasma.22,23

225

Lipid oxidation Thiobarbituric acid reactive substances values were utilized to indicate lipid oxidation.

226 227

24

228

blended with 25 mL trichloro acetic acid (TCA) solution (20%) and 20 mL distilled water.

229

Samples were homogenized using a Sorvell Omni tabletop mixer (Newton, CT) for 1 min and

230

filtered through a Whatman (#1) filter paper. One mL of filtrate was mixed with 1 mL of

231

thiobarbituric acid (TBA) solution (20 mM), incubated in a boiling water bath for 10 min. The

232

samples were cooled and the absorbance at 532 nm was measured using a Shimadzu UV-2600

233

PC spectrophotometer (Shimadzu Inc., Columbia, MD, USA). Blank samples for the

234

spectrophotometer consisted of 2 mL TCA/distilled water (1:1 v/v) and 2 mL TBA solution.

235

Statistical analysis

236

From each steak, 5 g of the sample that contained both interior and displayed surface was

The experimental design was a split-plot, where muscles (LL and PM) from each animal

237

served as a block in the whole plot (n = 8). In the subplot, a steak was assigned a day of the

238

display (0, 1, 3, 5, or 7 for surface color measurements and 0, 3, and 7 for biochemical studies).

239

For surface color, mitochondrial-, and biochemical studies, the fixed term included muscle type,

240

display time, and their interaction. The random terms included animal (Error A) and unspecified

241

residual error (Error B). Type-3 tests of fixed effects were performed using the Mixed Procedure

242

of SAS (version 9.3, SAS Institute Inc; Cary, NC). Least square means for protected F-tests (P < 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 28

243

0.05) were separated by using the pdiff option (least significant differences) and were considered

244

significant at P < 0.05 level.

245

RESULTS AND DISCUSSION

246

Effects of muscle type on surface color and biochemical properties

247

The relationship between oxidative stress, mitochondria, and cell death have been

248

investigated extensively in medical research.25,26 However, limited information is currently

249

available on muscle-specific oxidative stress on mitochondrial degeneration. There were no

250

significant differences in proximate compositions between two muscles (Table 1). In the current

251

study, PM had greater % metmyoglobin and lower redness (P < 0.05) compared with LL on day

252

3 (Table 2). LL had a stable red color than PM during display (changes in % metmyoglobin level

253

in LL and PM between day 0 and 7 were 28.4 and 78.3, respectively). A lower color stability in

254

PM was characterized by lesser (P < 0.05) MRA than LL on days 0, 3, and 7 (Table 3). However,

255

PM had greater (P < 0.05) muscle oxygen consumption on day 0, while LL had greater oxygen

256

consumption on days 3 and 7 than PM (Table 3). There were no differences (P > 0.05) between

257

muscle types for lipid oxidation on day 0 of the display, however, PM had greater lipid oxidation

258

(P < 0.05) on days 3 and 7. Various studies have reported color labile nature of PM during

259

display than LL.3-5 The authors speculated that differences in fiber type and greater oxygen

260

consumption attributed to rapid discoloration. In the present research, we focused on how

261

oxidative changes can affect mitochondrial degradation and cytochrome c level in sarcoplasm to

262

understand the lesser color stability of PM.

263 264

12 ACS Paragon Plus Environment

Page 13 of 28

265 266

Journal of Agricultural and Food Chemistry

Mitochondrial oxygen consumption Mitochondrial oxygen consumption was greater for PM on day 0 than LL (Table 4). This

267

suggests that there are inherent differences in oxygen consumption capabilities per mg

268

mitochondria between muscle types. Previous research also reported a greater oxygen

269

consumption in isolated bovine PM mitochondria27 and in steaks28 compared with LL. Greater

270

oxygen consumption ability of PM mitochondria and greater mitochondrial content may be

271

responsible for increased oxygen consumption in PM steaks. However, the overall change in

272

mitochondrial oxygen consumption from day 0 to day 7 for LL was less (P < 0.05) than PM

273

(changes in mitochondrial oxygen consumption in LL and PM between day 0 and 7 were 7.8 and

274

27.3, respectively).

275

Immediately after animal harvest, cells try to maintain homeostasis for energy production;

276

however, the lack of blood flow leads to failure of adaptive mechanisms. These changes can

277

affect the pro- and antioxidant balance, leading to accumulation of reactive oxygen species

278

(ROS). Tricarboxylic substrates such as citrate and malic acid are utilized faster in PM than that

279

in LL.29 Both citrate and malate are intermediates in the tricarboxylic acid cycle and can be used

280

for mitochondrial oxygen consumption. Postmortem bovine muscle can use succinate, pyruvate,

281

and malate for oxygen consumption.6,7 With a less efficient antioxidant system, mitochondrial

282

utilization of substrates for oxygen consumption can lead to greater ROS and increased oxidative

283

changes which can trigger lipid oxidation. Both complexes I and III are the potential sites of

284

ROS.30,31 Pre-incubation of HNE (a secondary lipid oxidation product) with bovine heart

285

mitochondria decreased mitochondrial oxygen consumption.10 The reaction between

286

mitochondrial protein and lipid oxidaiton products can be partially accounted for lower oxygen

287

consumption in PM than LL mitochondria on day 7. Although mitochondrial oxygen

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 28

288

consumption was greater on day 0, an increased oxidative stress combined with mitochondrial

289

degeneration may be responsible for greater changes in oxygen consumption between day 0 and

290

7 for PM than LL.

291

Oxidation-reduction potential

292

There was a significant muscle type and display time interaction for ORP (P < 0.001).

293

PM had a greater ORP on day 0 (indicative of electron donating proteins or compounds) than LL

294

(Figure 1). The ORP measurements using the RedoxSys implies the net balance of oxidants and

295

reductants, providing a direct measure of oxidative stress.22 Mitochondria contain several

296

antioxidant proteins and electron accepting compounds. Hence initial ORP values were higher in

297

PM than LL (a lower number indicates a greater antioxidant potential). Interestingly, ORP values

298

for LL did not change significantly during display. Although the mechanism for stable ORP in

299

LL is not clear, we speculate that a greater MRA and a lower muscle oxygen consumption may

300

have limited oxidative changes. ORP changes between days 0 and 3 were greater for PM than LL.

301

This suggests that the antioxidant defense mechanisms lose its efficacy more rapidly in PM than

302

in LL. In support, Joseph et al.4 reported that PM has a lower amount of antioxidant proteins than

303

LL in sarcoplasm, which makes it more susceptible to oxidative stress and buildup of free

304

radicals and secondary lipid oxidation products, leading to faster discoloration. Lipid oxidation

305

products can bind with lactic acid dehydrogenase and myoglobin to make muscles more prone to

306

discoloration.25,26 In support, redness and % metmyoglobin accumulation were greater in PM

307

than LL between days 1 and 3.

308

Mitochondrial and cytochrome c content changes

309 310

On day 0, PM muscle had a 2.55-fold greater mitochondrial DNA content than LL (Table 4). On day 0 of display, the average mitochondrial protein content in LL and PM were 0.21 and 14 ACS Paragon Plus Environment

Page 15 of 28

Journal of Agricultural and Food Chemistry

311

0.38 mg/g of muscle, respectively (P = 0.02). This represents approximately a 1.80-fold

312

difference in mitochondrial content between two muscles. PM has predominant red fibers32 than

313

LL. Mohan et al.3 reported approximately a 1.21-fold higher mitochondrial concentration in beef

314

PM compared with LL muscle in 10-day aged samples using cytochrome c oxidase activity.

315

Previous research also demonstrated that mitochondrial enzyme levels were lower in rabbit and

316

guinea pigs white muscles than in red muscles.33,34 However, in the present study mitochondrial

317

content in PM was lower than LL on days 3 and 7. This demonstrates that the mitochondrial

318

degradation was greater in PM than LL.

319

Mitochondria are the first organelle subjected to oxidative damage from free radicals,

320

which can trigger mitochondrial degradation.11 Although the mechanism of muscle-specific

321

differences in mitochondrial degradation in postmortem muscle is not clear; we speculate the

322

effects of oxidative stress on mitochondria from various disease conditions such as Alzheimer's

323

or Parkinson's diseases from the medical research. Both inner and outer mitochondrial

324

membranes are composed of phospholipids. Further, oxidative phosphorylation occurs within

325

the mitochondrial inner membrane which can be a source of ROS formation. Hence,

326

mitochondrial membranes are very prone to lipid peroxidation. Previous research demonstrated

327

that incubation of secondary lipid oxidation product such as HNE with bovine heart

328

mitochondria increased vacuolation and fragmentation.10 In normal mitochondria, the outer

329

membrane is more permeable due to porin proteins than the inner membrane; however, the

330

oxidative condition can increase permeability and fluidity of the inner membrane. 35,36 This can

331

lead to release of cytochrome c from the inner membrane to the outside. There were no

332

significant differences in cytochrome c content in sarcoplasm of LL muscle with increased

333

storage time, but the cytochrome c content increased for PM (Figure 2). In healthy cells,

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 28

334

cytochrome c is located within the mitochondrial inner membrane and also interact with

335

cardiolipin (a diphosphatidylglycerol lipid present in the mitochondrial membrane).37

336

Physiologically, greater levels of free cytochrome c in intermembrane of mitochondria indicate

337

mitochondrial damage by ROS.12 Several proapoptotic stimuli can trigger permeabilization of the

338

outer mitochondrial membrane and promote mobilization of cytochrome c from the inner

339

membrane and cardiolipin38,39 to sarcoplasm. Currently, limited knowledge is available on the

340

factors that trigger the release of cytochrome c to sarcoplasm in postmortem muscles. Since there

341

were significant differences between muscle types, a systemic analysis of changes in

342

mitochondrial- and cytochrome c content can be used as a potential tool to predict color stability.

343

As the results suggest, mitochondrial degradation caused by oxidative stress can

344

influence meat color stability. Various studies have demonstrated the role of mitochondria in

345

color stability6-7,9 and in bloom development.40-43 A greater oxygen consumption may be

346

detrimental to color stability and also for the formation of a bright-red color. Hence, controlling

347

mitochondrial oxygen consumption via postharvest techniques such as packaging or storage

348

temperature can minimize discoloration. A moderate mitochondrial activity might be beneficial

349

for a brighter-red steak that lasts for a longer time.

350

In summary, PM had a shorter color stability than LL. In addition to the role of oxygen

351

consumption and MRA, oxidative stress has muscle-specific effects on mitochondrial

352

degradation and cytochrome c release. PM had greater mitochondrial degradation and less

353

antioxidant capacity than LL with increased display time. Greater oxidative stress in PM may be

354

responsible for cytochrome c release and lower mitochondrial content in PM than LL. Therefore,

355

characterizing the factors responsible for mitochondrial degradation postmortem may help to

356

understand the muscle-specific differences in color stability.

16 ACS Paragon Plus Environment

Page 17 of 28

Journal of Agricultural and Food Chemistry

357

ABBREVIATIONS USED

358

LL, longissimus lumborum; PM, psoas major; ORP, oxidation-reduction potential; MRA,

359

metmyoglobin reducing activity; AOAC, Association of Official Analytical Chemists; IMPS,

360

Institutional Meat Purchase Specifications; USDA, United States Department of Agriculture;

361

PVC, polyvinyl chloride; HNE, 4-hydroxy-2-nonenal

362

ACKNOWLEDGMENT

363

Appreciation is expressed to the RedoxSys/Aytu Biosciences for providing the oxidation-

364

reduction potential meter.

365

REFERENCES

366 367

1. AMSA. Meat color measurement guidelines. American meat science association, Champaign. 2012.

368 369 370

2. Smith, G. C.; Belk, K. E.; Sofos, J. N.; Tatum, J. D.; Williams, S. N. Economic implications of improved color stability in beef. In Antioxidants in muscle foods: Nutritional strategies to improve quality. New York: Wiley-Interscience. 2000; pp. 397–426.

371 372 373

3. Mohan, A.; Hunt, M. C.; Muthukrishnan, S.; Barstow, T. J.; Houser, T. A. Myoglobin redox form stabilization by compartmentalized lactate and malate dehydrogenases. J. Agric. Food Chem. 2010, 58, 7021-7029.

374 375

4. Joseph, P.; Suman, S. P.; Rentfrow, G.; Li, S.; Beach, C. M. Proteomics of Muscle-Specific Beef Color Stability. J. Agric. Food Chem. 2012, 60, 3196-3203.

376 377 378

5. McKenna, D.; Mies, P.; Baird, B.; Pfeiffer, K.; Ellebracht, J.; Savell, J. Biochemical and physical factors affecting discoloration characteristics of 19 bovine muscles. Meat Sci. 2005, 70, 665-682.

379 380 381

6. Tang, J.; Faustman, C.; Hoagland, T. A.; Mancini, R. A.; Seyfert, M.; Hunt, M. C. Postmortem oxygen consumption by mitochondria and its effects on myoglobin form and stability. J. Agric. Food Chem. 2005, 53, 1223-1230.

382 383

7. Ramanathan, R.; Mancini, R. A. Effects of pyruvate on bovine heart mitochondria-mediated metmyoglobin reduction. Meat Sci. 2010, 86, 738-741.

384 385

8. Lanari, M.C.; Cassens, R.G. Mitochondrial activity and beef muscle color stability. J. Food Sci. 1991, 56, 1476-1479.

386 387

9. Mancini, R. A.; Ramanathan, R. Effects of postmortem storage time on color and mitochondria in beef. Meat Sci. 2014, 98, 65–70. 17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 28

388 389 390

10. Ramanathan, R.; Mancini, R. A.; Suman, S. P.; Cantino, M. E. Effects of 4-hydroxy-2nonenal on beef heart mitochondrial ultrastructure, oxygen consumption, and metmyoglobin reduction. Meat Sci. 2012, 90, 564–571.

391 392 393

11. Ouali, A.; Gagaoua, M.; Boudida, Y.; Becila, S.; Boudjellal, A.; Herrera-Mendez, C. H.; Sentandreu, M. A. Biomarkers of meat tenderness: Present knowledge and perspectives in regards to our current understanding of the mechanisms involved. Meat Sci. 2013, 95, 854-870.

394 395

12. Ott, M.; Gogvadze. V.; Orrenius. S.; Zhivotovsky. B. Mitochondria, oxidative stress, and cell death. Apoptosis. 2007, 12, 913–922.

396 397 398

13. Ding, W. X.; Shen, H. M.; Ong, C. N. Calpain activation after mitochondrial permeability transition in microcystin-induced cell death in rat hepatocytes. Biochemical and Biophysical Research Communications. 2002, 291(2), 321–331.

399 400 401

14. Bjugstad, K. B.; Fanale, C.; Wagner, J.; Jensen, J.; Salottolo, K.; Rael, L. T.; Bar-Or, D. A. 24h delay in the redox response distinguishes the most severe stroke patients from less severe stroke patients. J. Neurol. Neurophys. 2016, 7(5), 100395.

402 403

15. The North American Meat Processors Association. (NAMP). The meat buyers guide. Reston, VA: North American Meat Processors Association. 2002.

404 405

16. Madhavi, D. L.; Carpenter C. L. Aging and processing affect color, metmyoglobin reductase and oxygen consumption of beef muscles. J. Food Sci. 1993, 939-947.

406 407 408

17. Sammel, L. M.; Hunt, M. C.; Kropf, D. H.; Hachmeister, K. A.; Johnson, D. E. Comparison of assays for metmyoglobin reducing ability in beef inside and outside semimembranosus muscle. J. Food Sci. 2002, 67, 978-984.

409 410 411

18. Grubbs, J. K.; Fritchen, A. N.; Huff-Lonergan, E.; Dekkers, J. C. M.; Gabler, N. K.; Lonergan, S. M. Divergent genetic selection for residual feed intake impacts mitochondria reactive oxygen species production in pigs. J. Animal Sci. 2013, 91, 2133–2140.

412 413 414

19. Matsunaga, T.; Chikuni, K.; Tanabe, R.; Muroya, S.; Shibata, K.; Yamada, J.; Shinmura, Y. A quick and simple method for the identification of meat species and meat products by PCR assay. Meat Sci. 1999, 51, 143-148.

415 416

20. Schmittgen, T. D.; Livak, K. J. Analyzing real-time PCR data by the comparative C-T method. Nature Protocols. 2008, 3, 1101-1108.

417 418

21. Estabrook, R. W. Mitochondrial respiratory control and the polarographic measurement of ADP: O ratios. Meth. in Enzy. 1967, 10, 41–47.

419 420 421

22. Rael, L. T.; Bar-Or, R.; Kelly, M. T.; Carrick, M. M.; Bar-Or, D. Assessment of oxidative stress in patients with an isolated traumatic brain injury using disposable electrochemical test strips. Electroanalysis. 2015. doi: 10.1002/elan.201500178

422 423 424

23. Spanidis, Y.; Goutzourelas, N.; Stagos, D.; Kolyva, A. S.; Gogos, C. A.; Bar-Or, D.; Kouretas, D. Assessment of oxidative stress in septic and obese patients using markers of oxidation-reduction potential. In vivo. 2015, 29(5), 595-600.

425 426

24. Witte, V. C.; Krause, G. F.; Bailey, M. E. A new extraction method for determining 2thiobarbituric acid values of pork and beef during storage. J. Food Sci. 1970, 35, 582–585.

18 ACS Paragon Plus Environment

Page 19 of 28

Journal of Agricultural and Food Chemistry

427 428

25. Esterbauer, H.; Schaur, R. J.; Zollner, H. Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radi. Bio. Med. 1991, 11, 81–128.

429 430

26. Chen, J. J.; Yu, B. P. Alterations in mitochondrial membrane fluidity by lipid peroxidation products. Free Radical Bio. Med. 1994; 17, 411–418.

431 432 433

27. Belskie, K. M.; Ramanathan, R.; Suman, S. P.; Mancini, R. A. Effects of muscle type and display time on beef mitochondria. In The Proceedings of American Meat Science Association, Reciprocal Meat Conference. 2014; abstract 117.

434 435 436

28. Seyfert, M.; Mancini, R. A.; Hunt, M. C.; Tang, J.; Faustman, C.; Garcia, M. Color stability, reducing activity, and cytochrome c oxidase activity of five bovine muscles. J. Agric. Food Chem. 2006, 54, 8919-8925.

437 438 439

29. Abraham, A.; Dillwith, J. W.; Mafi, G. G.; VanOverbeke, D. L.; Ramanathan, R. Metabolite profile differences between beef longissimus and psoas muscles during display. Meat Mus. Bio. 2017. 1, 18-27.

440 441

30. Raha, S; Robinson. B.H. Mitochondria, oxygen free radicals, disease and ageing. Trends Biochem Sci. 2000, 25, 502–508.

442 443

31. Turrens, J. F. Mitochondrial formation of reactive oxygen species. J Physiol. 2003, 552, 335–344.

444 445

32. Hunt, M. C.; Hedrick, H. B. Profile of fiber types and related properties of five bovine muscles. J. Food Sci. 1977, 42, 513–517.

446 447

33. Glancy, B.; Balaban, R. S. Protein composition and function of red and white skeletal muscle mitochondria. Am J Physiol Cell Physiol. 2011, 300, C1280–C1290.

448 449 450

34. Peter. J. B.; Barnard, R. J.; Edgerton, V. R.; Gillespie, C. A.; Stempel, K. E. Metabolic profiles of three fiber types of skeletal muscle in guinea pig and rabbit. Biochem. 1972, 11, 627– 2633.

451 452

35. Guo, C.; Sun, L.; Chen, X.; Zhang. D. Oxidative stress, mitochondrial damage and neurodegenerative diseases. Neural Regener. Res. 2013, 8, 2003.

453 454 455

36. Chen, J.; Schenker, S.; Frosto, T. A.; Henderson, G. I. Inhibition of cytochrome c oxidase activity by 4-hydroxynonenal (HNE). Role of HNE adduct formation with the enzyme subunits. Biochim. et Biophy. Acta. 1998, 1380, 336–344.

456 457

37. Garrido, C.; Galluzzi L.; Brunet M.; Puig PE.; Didelot-Mirjolet C.; Kroemer G. Mechanisms of cytochrome c release from mitochondria. Cell Death Differ. 2006, 13, 1423–1433.

458 459 460

38. Kagan, V.E.; Tyurin, V.A.; Jiang, J.; Tyurina, Y.Y.; Ritov, V.B.; Amoscato, A.A.; sipov, A.N.; Belikova, N.A.; Kapralov, A.A.; Kini, V.; Vlasova. I.I. Cytochrome c acts as a cardiolipin oxygenase required for release of proapoptotic factors. Nat. Chem. Biol. 2005, 1, 223-232.

461 462

39. Sinha, K.; Das, J.; Pal, P. B.; Sil, P.C. Oxidative stress: the mitochondria-dependent and mitochondria-independent pathways of apoptosis. Arch Toxicol. 2013, 87, 1157–1180.

463 464

40. Ashmore, C. R.; Doerr, L.; Parker, W. Respiration of mitochondria isolated from darkcutting beef- postmortem changes. J. Anim. Sci. 1972, 34, 46–48.

465 466

41. Cornforth, D. P.; Egbert, W. R. Effect of rotenone and pH on the color of pre-rigor muscle. J. Food Sci. 1985, 50, 34–35. 19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 28

467 468

42. English, A. R.; Mafi, G. G.; VanOverbeke, D. L.; Ramanathan. R. Effects of extended aging on biochemical properties of dark cutting beef. J. Anim. Sci. 2016, 94, 4040–4048.

469 470 471

43. English, A. R.; G. G. Mafi.; D. L. VanOverbeke.; R. Ramanathan. Effects of extended aging and modified atmospheric packaging on beef top loin steak color. J. Anim. Sci. 2016. 94, 1727– 1737.

472

20 ACS Paragon Plus Environment

Page 21 of 28

473

Journal of Agricultural and Food Chemistry

Figure captions

474 475

Figure 1

476

Effects of muscle type and display time on oxidation-reduction potential properties

477 478 479 480

A lower value represents greater antioxidant capacity. a–c

Least squares means with a different letter differ (P < 0.05); standard error bars are indicated at each time point. LL = longissimus lumborum; PM = psoas major

481 482 483

Figure 2

484

Effects of muscle type and display time on cytochrome c content in sarcoplasm

485 486 487

a–d

Least squares means with a different letter differ (P < 0.05); standard error bars are indicated at each time point. LL = longissimus lumborum; PM = psoas major

488 489 490 491

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

492

Page 22 of 28

Figure 1

Oxidation reduction potential (mV)

295 290

LL

PM

a a

a

285 280

c

c

275 270

b

265 260 255 250 0

3

7

Display time (d) 493 494

22 ACS Paragon Plus Environment

Page 23 of 28

495

Journal of Agricultural and Food Chemistry

Figure 2 0.16 LL

PM

Cytochrome c content (µg/g tissue)

0.14

d c

0.12 0.10

b a a

0.08

a

0.06 0.04 0.02 0.00 0

3 Display time (d)

7

496 497 498

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

499

Table 1

500

Proximate composition and pH of LL and PM muscles Trait

LL

PM

SE

P-value

Moisture (%)

70.5

70.4

0.97

0.94

Protein (%)

21.7

20.9

0.4

0.28

Fat (%)

5.9

7.3

1.3

0.51

pH

5.57

5.61

0.02

0.13

Page 24 of 28

501 502

LL = longissimus lumborum; PM = psoas major

503

SE = standard error

504

24 ACS Paragon Plus Environment

Page 25 of 28

Journal of Agricultural and Food Chemistry

505

Table 2

506

Least square means for the effects of muscle type and display time on surface color Trait a) a* values (redness)

Muscle LL PM

0 30.6a,x 29.5a,y

Display time (day) 1 3 5 b,x a,x 33.4 30.1 28.4c,x 32.4b,y 20.3c,y 15.4d,y

b) % metmyoglobin

LL PM

7.2a,x 5.6a,x

7.2a,x 10.5b,x

12.5b,x 65.2c,y

21.4c,x 77.5d,y

7 26.4d,x 13.1e,y 35.6d,x 83.9e,y

507 508

LL = longissimus lumborum; PM = psoas major

509 510

For both a* value and % metmyoglobin: display time, P < 0.0001; muscle effect, P < 0.0001; display time x muscle, P < 0.0001

511 512

x-y

Means in a column with a different letter differ (P < 0.05)

513

a-e

Means in a row with a different letter differ (P < 0.05)

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

514

Table 3

515 516

Least square means for the effects of muscle type and display time on MRA, oxygen consumption, and lipid oxidation

Page 26 of 28

517

519 520

SE518 Display time (day) 0 3 7 x,a x,b 3.2 a) MRA (%) LL 65.4 60.4 54.4 x,c y,a y,b y,b PM 55.4 30.4 26.4 x,a x,b 3.1 b) Oxygen consumption (%) LL 57.4 50.4 42.4 x,c y,a y,b y,c PM 68.5 45.6 34.5 x,a x,b 0.009 c) Lipid oxidation LL 0.031 0.052 0.077 x,c PM 0.038 x,a 0.072 y,b 0.124 y,c LL = longissimus lumborum; PM = psoas major; MRA = metmyoglobin reducing activity; SE = standard error

521 522

For both MRA and oxygen consumption: display time, P < 0.0001; muscle effect, P < 0.001; display time x muscle, P < 0.001

523 524

Lipid oxidation is indicated as thiobarbituric acid reactive substances and measured as absorbance at 532 nm.

525

x-y

Means in a column with a different letter differ (P < 0.05)

526

a-c

Means in a row with a different letter differ (P < 0.05)

26 ACS Paragon Plus Environment

Page 27 of 28

Journal of Agricultural and Food Chemistry

527

Table 4

528

Comparison of mitochondrial content and oxygen consumption in beef longissimus and psoas muscles during display Parameters Mitochondrial content

Mitochondrial oxygen consumption (nanomoles of oxygen consumed/min per mg of mitochondria)

LL

PM

SE

P-value

Mitochondrial protein yield, d 0 Mitochondrial protein yield, d 3 Mitochondrial protein yield, d 7 Relative mitochondrial DNA content, d 0

0.21 x,a 0.19 x,ab 0.15 x,b 1.00

0.38 y,a 0.24 x,b 0.12 x,c 2.55

0.03

0.002

0.28

0.01

d0 d3

30.2 x,a 26.5 x,ab

42.5 y,a 32.4 x,b

3.2

0.001

d7

22.4 x,b

15.2 y,c

529 530 531 532

Mitochondrial protein yield represents total mitochondria mg of protein per gram of tissue isolated from paired LL and PM muscles by differential centrifugation; mitochondrial DNA content was reported as a relative fold change in comparison to paired PM and LL muscles.

533

For both mitochondrial content and mitochondrial oxygen consumption: display time, P < 0.0001; muscle effect, P < 0.001

534

SE = standard error; LL = longissimus lumborum; PM = psoas major

535

x-y

Means in a row with a different letter differ (P < 0.05)

536

a-c

Means in a column with a different letter differ (P < 0.05)

537 538

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

TOC for the manuscript 82x44mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 28