Pectoralis Muscle of Turkey Displays Divergent Function as ... · Pectoralis muscle of turkey...

71
Pectoralis Muscle of Turkey Displays Divergent Function as Correlated with Meat Quality Bly Addison Patterson Thesis submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements of degree of Master of Science In Animal and Poultry Sciences David E. Gerrard, Committee Chair Robert P. Rhoads Scott J. Eilert April 30, 2015 Blacksburg, VA Keywords: turkey quality, postmortem glycolysis, meat color Copyright 2015 (Bly Patterson)

Transcript of Pectoralis Muscle of Turkey Displays Divergent Function as ... · Pectoralis muscle of turkey...

Pectoralis Muscle of Turkey Displays Divergent Function as Correlated with Meat Quality

Bly Addison Patterson

Thesis submitted to the faculty of the Virginia Polytechnic Institute and State University in

partial fulfillment of the requirements of degree of

Master of Science

In

Animal and Poultry Sciences

David E. Gerrard, Committee Chair

Robert P. Rhoads

Scott J. Eilert

April 30, 2015

Blacksburg, VA

Keywords: turkey quality, postmortem glycolysis, meat color

Copyright 2015 (Bly Patterson)

Pectoralis muscle of turkey displays divergent function as correlated with meat quality

Bly Addison Patterson

ABSTRACT

Fresh turkey meat color is influenced by a myriad of biological factors which include

muscle fiber type composition and heme protein concentrations. These factors either contribute

to or are subject to the biochemical events involved in the conversion of muscle to meat. Subtle

deviations in the processing environment can also result in aberrant fresh meat quality

development and may ultimately alter the quality characteristics of cooked product. Our

objective was to describe the underlying cause and significance of two-toning in fresh turkey

breast. In the first experiment, pectoralis muscles were collected and subjected to image

processing software to describe color of fresh turkey. In the second experiment, shackling time

was tested as an aggravator of fresh turkey color. Results showed turkey breast possess two-

lobes that differ in pH, drip loss, energy metabolism and muscle fiber type composition. Results

also showed fresh turkey color was enhanced during the time from stun to exsanguination (P <

0.05). These results suggest inherent differences in breast muscle are responsible for variations

in fresh turkey color.

iii

Dedication

This thesis is dedicated to all of the farmers, agriculture teachers, and community leaders at

home that supported me as a 4-H and FFA member and have helped shape me into the

agriculturalist that I am.

iv

Acknowledgements

This experience has been one for the books. I would not have survived any of it without the

support of mentors, family, and friends.

Many thanks to

Jesus Christ, my savior: when I felt as though nothing was going my way, I poured over your

scriptures and prayed. It is there that I found peace and understanding that through hard work

and faith, I would accomplish this task. I only hope that I can continue to honor you with the

skills and gifts you have given me. Perhaps I might touch a few lives along the way.

Dr. Dave Gerrard, my advisor: for challenging me to reach my greatest potential. You taught

me to think outside the box and look at problems from a multitude of angles. It is because of you

I can now think. I truly appreciate “the push” and am a better scientist because of it.

Dr. Scott Eilert: for your insight on meat quality problems in commercial processing and how to

address them in academic research. I am grateful to have had the opportunity to work with

Cargill on this project and cannot thank you enough for introducing me to the wonderful people

on the Dayton team.

Dr. Robert Rhoads: for bringing a different perspective to my project and providing me with

direction when it was needed.

Eric Stewart and the statisticians: for writing the Matlab software scripts and algorithms for

meat image analysis.

Eric England: for answering countless questions and teaching me the wonders of protein gel

electrophoresis and western blot.

Sulaiman Matarneh: for helping me with all things statistics.

To all of my labmates: Eric England, Sulaiman Matarneh, Kristen Stufft, Richard Preisser,

Amanda Fabi, and Jordan Wicks. Many thanks for the great laughs and conversation. From our

social nights to pipetting and crunching numbers, I will cherish the memories for a lifetime.

My undergraduate volunteer, Mallory White: Without you, I probably would not be finished

with this degree. Thank you for the late night company in the lab running assays and grinding

hundreds of tissue samples.

v

Derick Shank, my fiancé: There are not enough words to describe how much I appreciate you

being there for me through my undergraduate and graduate education. From sharing in my daily

joys to my most defining and devastating moments, you have seen it all. I cannot wait to marry

you in a few short months. I love you.

My family: for all of the love and support during this experience. Mom and Dad: thank you for

always being there when I needed someone to talk to and reminding me that I can do anything I

set my mind to.

The Dayton team: Keith Carter, Randy Batson, Malfred Shaw, and Christian Rodriguez. For

being so accommodating when I needed to collect samples and data. I could not have asked for

better help. Thank you.

vi

Table of Contents

Abstract…………………………………………………………………………………………..ii

Dedication…………………………………………………………………………………..……iii

Acknowledgements……………………………………………………………………………...iv

Table of Contents………………………………………………………………………………..vi

List of Figures...………………………………………………………………………………..viii

List of Tables…………………………………………………………………………………….ix

Chapter 1. Introduction and Literature Review-Meat Quality Development……………….1

Introduction……………………………………………………………………………….1

Meat Quality………………………………………………………………………………1

Meat Color………………………………………………………………………………...2

Heme Pigments…………………………………………………………………….…2

Myoglobin…………………………………………………………………………2

Hemoglobin………………………………………………………………….…….6

Cytochrome c…………………………………………………………………...…7

Muscle Fiber Type……………………………………………………………………7

Conversion of Muscle to Meat…………………………………………………………...10

Aberrations Affecting Postmortem Metabolism and Meat Quality……………………...11

Stun and Blood Removal………………………………………………………...13

Summary……………………………………………………………………………….... 15

References………………………………………………………………………………..16

Chapter 2. Pectoralis muscle of turkey displays divergent function as correlated with meat

quality…………………………………………………………………………………………..19

Abstract…………………………………………………………………………………..19

Introduction………………………………………………………………………………19

Materials and Methods…………………………………………………………………...21

Results and Discussion…………………………………………………………………..27

Conclusions………………………………………………………………………………35

References………………………………………………………………………………..54

Chapter 3. Future directions…………………………………………………………………...56

Appendix A……………………………………………………………………………………...59

vii

List of Figures

Chapter 1

Figure 1. Structure of the heme group of myoglobin…………………………………...………....3

Figure 2. Myoglobin redox forms of raw meat…………………………………………...……….4

Chapter 2

Figure 1. Sampling location of the large and small lobe of the pectoralis major muscle in adult

tom turkeys for experiment one and two………………………………………………………...36

Figure 2. Lightness (L*) color values and standard deviation as an expression of variability of the

large and small lobes and single-toned and two-toned populations of pectoralis major muscle in

adult tom turkeys…………………………………………………………………………………37

Figure 3. Redness (a*) color values and standard deviation as an expression of variability for the

large and small lobes and single-toned and two-toned populations of pectoralis major muscle of

adult tom turkeys………………………………………………………………………………....38

Figure 4. Relative myoglobin mRNA abundance of the small and large lobes of the pectoralis

major muscle in adult tom turkeys……………………………………………………………….39

Figure 5. Hemoglobin content of small and large lobes of the pectoralis major muscle in adult

tom turkeys…………………………………………………………………………………….…40

Figure 6. Ultimate pH of the small and large lobes of the pectoralis major muscle in adult tom

turkeys………………………………………………………………………………….…….......41

Figure 7. Lactate dehydrogenase content of the large and small lobes of the pectoralis major

muscle in adult tom turkeys………………………………………………………………….......43

Figure 8. Driploss of the small and large lobes of the pectoralis major muscle in adult tom

turkeys……………………………………………………………………………………............44

Figure 9. Soluble and total protein concentration of the large and small lobe of the pectoralis

major muscle in adult tom turkeys……………………………………………………….………45

viii

Figure 10. Sarcomere length of the large and small lobes of the pectoralis major muscle in adult

tom turkeys……………………………………………………………….………………………46

Figure 11. Large and small lobes of the pectoralis major muscle of adult tom turkeys shackled

first and lass from a transport truck exhibit color changes and increased variation with time…..47

Figure 12. Redness (a*) color values and variability for the large and small lobes and of the

pectoralis major muscle of adult tom turkeys shackled first and last from a transport truck…....48

Figure 13. Hemoglobin content of the small lobe of the pectoralis major muscle in adult tom

turkeys sampled first and last from a transport truck…………………………………….………49

Figure 14. Ultimate pH of large and small lobes of the pectoralis major muscle and first and last

shackled adult tom turkeys from a transport truck …………………………...……………….....51

Figure 15. Metabolite concentrations and glycolytic potential of the large and small lobe of the

pectoralis major muscle and first and last shackled adult tom turkeys from a transport truck.....52

Figure 16. Effect of stun to stick time on body temperature of adult tom turkeys………………53

Chapter 3

Figure 1. Sampling method for second study to test the effect of bird location during transport on

meat quality……………………………………………………...……………………………….57

Appendix A

Figure 1. Software script converting red, green, blue (RGB) color values to CIELAB color

space……………………………………………………………………………………………...59

Figure 2. Software script producing statistical parameters for the raw color distribution……….60

Figure 3. Software script truncating the original color distributions to eliminate values

representing background, outliers, and light scattering………………………………………….61

ix

List of Tables

Chapter 1

Table 1. Estimated Percentages of Type I, Type IIA, and Type IIB fibers in the proximal and

distal portions of five turkey muscles…………………………………………………………..…9

Chapter 2

Table 1. Metabolite concentrations and glycolytic potential of the large and small lobe of the

pectoralis major muscle of adult tom turkeys……………………………………...……………42

Table 2. Frequency of single-toned and two-toned pectoralis major muscles in tom turkeys

sampled first and last from a transport truck………………………………………………..... ...50

1

Chapter 1. Introduction and Literature Review-Meat Quality Development

Introduction

Over the past twenty years, consumer preferences and perception of meat has changed.

The meat industry has witnessed a shift toward more healthy, cost effective, and convenient

options due to consumer demand. In a U.S. survey, 55% of respondents classified convenience as

a ‘very important’ determinant of food purchases (Senauer, 2001). With an industry desire to

provide consumers with a variety of convenient eating options, consumption of further processed

products has increased (Grunert, 2006). As a result, poultry has become the second most popular

meat consumed in the world at 13.6 kg/capita/year after pork at 15.8 kg/capita/year (FAOSTAT,

2014).

Due to an increase in demand for poultry as a protein source, it is estimated that turkey

production had increased by 104 percent since 1970 (National Turkey Federation, 2014). This

rise in poultry consumption has placed more emphasis on quality traits like appearance and

texture (Fletcher, 2002). To remain a top meat choice of American consumers, product quality

and consistency must be maximized.

Meat Quality

The parameters that define meat quality are influenced by consumer preferences and meat

purchasing decisions. Quality traits considered important for a fresh, wholesome food product

are color, texture, and water-holding capacity. Of these factors, meat color is the most influential

to meat purchasing decisions (Mancini, 2005). Heme pigment concentration and muscle fiber

type composition are primarily responsible for the color that consumers perceive.

2

Meat Color

Heme Pigments

Meat color development is largely controlled by three heme proteins: myoglobin,

hemoglobin, and cytochrome c. In the living muscle, myoglobin and hemoglobin coordinate to

store and deliver oxygen to the tissue, which is necessary for aerobic energy metabolism. This

activity, in partnership with cytochrome c in the electron transport chain, provides the muscle

with ATP. Postmortem, these proteins are better known for their role in color development of

beef, lamb, pork, and poultry. When an animal is harvested, most of the hemoglobin leaves the

body in blood. Clydesdale and Francis (1972) reported that iron associated with myoglobin only

10% of the total iron of an animal prior to slaughter. After exanguination, this source of iron

(myoglobin) accounts for 95% of the total iron. As a result, myoglobin concentration, redox

state, and thermal stability play a vital role in the meat color perception.

Myoglobin

Myoglobin is a monomeric protein containing a single heme prosthetic group. This group

is surrounded by a globin moiety that protects it from undesirable reactions. The heme portion

contains an iron atom at the center of its porphyrin ring (Fig. 1). Four of the coordination sites of

the iron atom are bonded to the ring, while the fifth and sixth sites are available for binding to

ligands. The ligand bound to the sixth coordination site is the primary determinant of its redox

state.

3

Figure 1. Structure of the heme group of myoglobin (Seideman, Cross, Smith, & Durland,

1984)

In fresh meat, myoglobin exists in three primary forms. The first state, deoxymyoglobin

is characterized by the absence of a bound ligand to the heme iron of the porphyrin ring (Mancini

& Hunt, 2005). As a result, the iron is in the ferrous (Fe2+

) state. This redox form is associated

with the purplish-red color of meat. When myoglobin is exposed to oxygen, the diatomic atom

binds to the ferrous iron, causing a change in meat color. The resulting oxygenation is evident by

a bright cherry-red color which it is most desired by consumers. The high rate at which this

reaction occurs is triggered by the high affinity of oxygen for myoglobin (Pirko & Ayres, 1957)

and generally takes less than thirty minutes in a process known as bloom (Johnson, 1974).

4

Heme Heme

Fe2+

Fe2+

H20 02

Heme

Fe3+

H20

Figure 2. Myoglobin redox forms of raw meat. Myoglobin in the ferrous state (Fe2+

) can be

oxygenated to create oxymyoglobin or oxidized to form metmyoglobin. Metmyoglobin can be

reduced to its ferrous (Fe2+

) in the presence of oxygen to form oxymyoglobin. (Adapted from

Pegg et al., 1997)

In contrast, discoloration of meat can result from the oxidation of the ferrous (Fe2+

) iron

to ferric (Fe3+

) iron due to loss of valence electrons (Bekhit et al., 2003). As a result,

oxymyoglobin is converted to metmyoglobin, which appears brown. When the globin moiety

surrounding the heme ring is denatured, it can no longer sufficiently protect it (Bekhit et al.,

2003). Consequently, the heme group is quickly oxidized. Brooks (1937) stated that this “loss of

bloom” is noticeable when 60% of the myoglobin pigments are in the metmyoglobin form.

Therefore, the color distinguished by the naked eye is dictated by a predominating pigment

rather than a single pigment (Fig. 1-2).

Deoxygenation

Oxygenation

Oxidation

Reduction

Myoglobin

Purplish Red

Oxymyoglobin

Bright Red

Metmyoglobin

Brown

5

The primary structure of myoglobin is responsible for its redox and thermostability.

While myoglobin function is conserved across species, the structure of the protein is not. In fact,

livestock and poultry species are <75% homologous (Suman & Joseph, 2013). Recently, the

primary structure of turkey myoglobin was sequenced and found to have 100% homology with

chicken (Poulson Joseph et al., 2011). Poultry myoglobin differs from livestock due to a

modification in the number of histidine residues in its sequence. Joseph et al. (2011) reported

that turkey and chicken myoglobin contained 9 histidines as compared with 9 in pork, 13 in beef,

as well as 12 in sheep and goats. Additionally, poultry myoglobin is 300-400 Daltons greater

than other livestock species (Suman & Joseph, 2013). Specifically, turkey myoglobin has a

molecular mass 346 Da greater than beef myoglobin (P Joseph, Suman, Li, Beach, & Claus,

2010). Subsequently, these amino acid differences in red meat myoglobin and turkey myoglobin

primary structure are linked to differences in thermostability. For example, turkey myoglobin has

a greater resistance to heat-induced denaturation (P Joseph et al., 2010). A study conducted by

Ueki et al. (2004) determined that a one or two amino acid substitution in the primary structure

of myoglobin is sufficient for altering its stability. The substituted amino acids residues affect

tertiary myoglobin structure and, in essence, thermal stability and meat color formation (Suman

& Joseph, 2013).

Meat color development is considerably variable due to concentration differences in

myoglobin between muscles. Kranen et al. (1999) determined that myoglobin concentration

remains undetectable in the pectoralis muscle of chickens, while only 0.12 mg/g are detectable in

the sartorius, or thigh muscle, via SDS-PAGE and western blot techniques. The pectoralis

myoglobin concentration supports previous findings by Nishida and Nishida (1985) that reported

no detectable level of myoglobin. It was further determined that myoglobin levels were

6

detectable at 0.56 mg/g in the adductor muscle and 1.08 mg/g in the chicken heart (Kranen et al.,

1999). The variation in myoglobin concentration among poultry muscles further demonstrates

its importance in meat color formation and variation.

Hemoglobin

Hemoglobin is a major protein and pigment found in blood. When an animal is harvested,

blood is removed from the circulatory system and with it, most of the hemoglobin. Due to the

very small concentration of the protein remaining in the system, it has very little effect on meat

color. However, if exsanguination protocols are ineffective, as much as 20-30% of the blood

volume may remain in the carcass tissues, with a maximum of 10% in carcass muscles (Fox Jr,

1966). Subsequently, hemoglobin can play a key role in meat color development.

Unlike myoglobin, hemoglobin is a more complex globular protein. It is differentiated

from myoglobin by its tetrameric quaternary structure. The protein is comprised of two α and

two β subunits. The polytpeptide chains forming each subunit consist of α helices and beta

pleated sheets arranged in an alternating fashion. The protein chains of each subunit fold to

create a hydrophobic pocket that protects the heme group. As a result of its complex structure,

its reactivity not only depends on its pocket environment, but its interactions with groups

belonging to other subunits of the protein (Antonini, 1965). Consequently, the ligands that bind

hemoglobin can create conformational changes that may alter its reactivity and affect meat color

development.

7

Cytochrome C

Cytochrome c is a heme pigment found in the mitochondria of skeletal muscle cells.

Unlike myoglobin and hemoglobin, it is very heat stable. Cytochrome c was reported to resist

heat denaturation to temperatures as high as 105◦C (Cornish & Froning, 1974). Consequently,

cytochrome c is most influential of cooked meat color. Since myoglobin and hemoglobin are

decomposed during cooking, native cytochrome c may be more important natural color retention

at high temperatures (Pikul, Niewiarowicz, & Kupijaj, 1986). This is attributed to its pink

appearance in the reduced state. Pikul et al. (1986) measured the cytochrome c content in various

poultry meats, including turkey. Breast meat of turkeys contains cytochrome c at a concentration

of 13.1 ± 0.49 μg g-1

of tissue, while thigh meat has a concentration of 47.4 ± 1.47 μg g-1

of

tissue. In breast meat, cytochrome c represents 2.36% of the total heme pigment concentration as

compared to 2.2% in the thigh (Pikul et al., 1986). With quantifiable concentrations of

cytochrome c in poultry meat, it is likely a contributor to cooked meat color. However, with

higher concentrations of hemoglobin and myoglobin in fresh meat, its impact on meat color is

minimal.

Muscle Fiber Type

Significant animal to animal variation in meat quality exists. This variation is partially

explained by differences in muscle fiber type composition both within muscles and between

animals. As a result of differences in muscle fiber type, postmortem biochemical processes are

affected along with meat quality (Klont, Brocks, & Eikelenboom, 1998). Specifically, meat color

is in part explained by underlying muscle fiber type composition (Henckel, Oksbjerg, Erlandsen,

8

Barton-Gade, & Bejerholm, 1997). This is a result of variation in myoglobin concentration

caused by the fiber types themselves. Due to the role of myoglobin as a heme pigment, the

existing fiber type profile will result in differences in meat color.

In general, four fiber types exist: slow oxidative (Type I), fast twitch oxidative-glycolytic

(Type IIA), fast twitch glycolytic (Type IIB), and fast twitch glycolytic (Type IIX). While the

role of Type IIX fibers have not been extensively determined in turkey muscles, most turkey

muscles are comprised of Type IIA and Type IIB fibers. The gastrocnemius and sartorius are the

most oxidative turkey muscles, with a proportion of 75% Type I fibers (Wiskus et al., 1976). The

number of each fiber type present in a given muscle is largely dependent upon the muscle

function and location. Slow oxidative muscle fibers often appear red and have a high capillary

density. These fibers are red due to a higher concentration of myoglobin. The high capillary

density allows for more oxygenated blood to flow to the tissue and oxygen itself to be transferred

to myoglobin. This oxygen delivery system provides the necessary aerobic environment to aid in

ATP synthesis. Furthermore, these Type I muscle fibers have a low capacity to participate in

anaerobic metabolism and thus are common in muscles used for long term endurance. Due to

low anaerobic capacity, the phosphocreatine pool and glycolytic enzyme activity of these red

fibers is low (Matthews & Fox, 1976). In contrast, fast glycolytic fibers function under anaerobic

conditions that last for a short, intense time period. These fibers utilize glycogen stores rather

than mitochondrial respiration to fuel a fast twitch (Lee, Joo, & Ryu, 2010). Consequently,

glycogen content, glycolytic enzyme activity, and phosphocreatine stores are high (Matthews &

Fox, 1976). Because oxygen is unnecessary for their function, these Type IIB fibers appear white

due to a low capillary density and myoglobin concentration.

9

A variety of fiber types profiles are found across species and muscle location. For

instance, the avian pectoralis muscle is responsible for providing the power necessary for wing

movement during flight (Hartman, 1961). Thus, it is considered a muscle of locomotion. Wiskus,

Addis, & Ma (1976) characterized six turkey muscles and showed 95% of the muscle fibers

found in the proximal and distal portions of the pectoralis muscle were type IIB. The

semitendinosus, biceps femoris, sartorius, and gastrocnemius, located in the turkey thigh and

were found to have a greater proportion of type I and IIA fibers with some type IIB fibers. The

proximal portion of the sartorius had 50% type I fibers, while the distal portion of the

gastrocnemius had 75% red fibers (Wiskus et al., 1976). Of the six muscles analyzed, the turkey

pectoralis was the most glycolytic in nature, followed by the semitendinosus and biceps femoris

(Table 1).

Table 1-1. Estimated Percentages of Type I, Type IIA, and Type IIB fibers in the proximal

and distal portions of five turkey muscles. (Adapted from Wiskus et al., 1976)

Of these muscles, Wiskus et al. (1976) concluded that the pectoralis muscle had significantly

higher lactate dehydrogenase and aldolase activities than other muscles; residual glycogen

content was also extremely low. Differences in inherent fiber type properties demonstrate that

Muscle Type I Type IIA Type II B Type I Type IIA Type IIB

Pectoralis 0 5 95 0 5 95

Semitendinosus 0 40 60 0 40 60

Biceps femoris Trace (0-1) 70 30 Trace (0-1) 45 55

Sartorius 50 20 30 50 20 30

Gastrocnemius 10 40 50 75 10 15

Proximal portion Distal portion

10

the rate of postmortem metabolism and pH decline are partially driven by enzyme and metabolite

concentration and activity.

The Conversion of Muscle to Meat

Typically, the conversion of muscle to meat is a well-orchestrated set of events in which

the metabolic pathways responsible for energy metabolism function with a limited oxygen

supply. This is due to a change in the physiological condition of the animal. During animal

harvest, the blood supply is rapidly removed from the circulatory system in the process known as

exsanguination. As a result, oxygen is slowly depleted from the muscle. In an attempt to

maintain cell homeostasis, ATP requirements must be met. The phosphagen system initially

maintains ATP levels by quickly re-phosphorylating ADP. During postmortem activation,

phosphoryl groups are liberated from the phosphocreatine pool that was created by the activity of

creatine kinase; this aids in replenishing ATP (Bendall, 1973). In chicken, once the

phosphocreatine reservoir is depleted by 80% when euthanasia is preceded by electrical

stunning, ATP levels begin to diminish (Briskey, CASSENS, & TEAUTMAN, 1966). The

buildup of adenosine diphosphate (ADP) is a direct result of ATP breakdown. ADP is removed

from the system by the adenylate kinase reaction in which ATP and adenosine monophosphate

(AMP) are generated. AMP is converted to inosine monophosphate (IMP) by AMP deaminase,

which accumulates in the dying muscle (Lawrie, 1985). Once the phosphagen system is

exhausted glycogenolysis and anaerobic glycolysis predominate. The glucose-6-phosphate

produced from glycogenolysis is used in anaerobic glycolysis to synthesize pyruvate. Pyruvate is

converted to lactate to generate NAD+, which drives the continuation of anaerobic glycolysis.

Lactate is unable to be removed postmortem and accumulates in the muscle.

11

Eventually, the rate of ATP breakdown exceeds ATP synthesis. As a result, the ATP

necessary to break actomyosin cross-bridges and maintain a relaxed muscle state is unavailable.

With less than 1 μmol of ATP/g remaining in the muscle (Pietrzak et al., 1997), actomyosin

cross-bridges become permanent, indicating rigor mortis onset (Lawrie, 1985).

Concurrently to postmortem glycolysis, pH declines. This drop in pH is a result of

the movement of substrate through the enzyme phosphofructokinase before its inactivation,

allowing for pyruvate to be converted to lactate more rapidly (England et al., 2014). The rate and

extent to which this happens are modulators in turkey meat quality development (Berri,

Wacrenier, Millet, & Le Bihan-Duval, 2001). Poultry muscle differs from other meat species in

that the rate of glycolysis is more rapid (Addis, 1986). Typically, pH moderately declines from

7.0 in the living muscle to approximately 6.0 within or in 1 h postmortem (Pietrzak, Greaser, &

Sosnicki, 1997) at a rate of 0.03units/min in turkey breast muscle (Sosnicki, Greaser, Pietrzak,

Pospiech, & Sante, 1998). An ultimate pH of 5.7-5.9 is reached (Briskey et al., 1966).

Aberrations Affecting Postmortem Metabolism and Meat Quality

Temperature

Temperature, processing techniques, and exsanguination contribute to quality variation in

a commercial processing environment. These factors often affect normal postmortem metabolism

by altering the rate and extent to which pH declines. Consequently, quality attributes such as

color, texture, and water-holding capacity are changed; the result is aberrant meat.

Meat pH and color are highly correlated. High ultimate pH is commonly associated with

dark meat and can be firm and dry (DFD); this is often the result of animals that have

12

experienced extensive pre-harvest stress. For poultry species, this defect is often correlated with

a pH greater than 6.1 (Swatland, 2008). Similarly, low muscle pH is linked to lighter meat and is

characterized as pale, soft, and exudative (PSE). Turkey meat with this quality defect has a final

pH of less than 5.8 (Sosnicki et al., 1998). Poor water-holding capacity combined with high

lightness (L*) color values are the general parameters describing PSE poultry (Barbut, 1997;

Woelfel, Owens, Hirschler, & Sams, 1998).

Rigor development at elevated muscle temperatures can induce the PSE condition in

turkey meat. FREMERY and POOL (1960) determined that birds exposed to temperatures

between 37 and 41◦C during processing exhibit rapid postmortem glycolysis and early rigor

onset. In agreement, it was reported that chicken pectoralis maintained at a temperature between

30 and 37◦C displayed an increased rate of postmortem glycolysis and ATP depletion (Khan,

1971). Poultry meat demonstrating and increased rate of postmortem glycolysis has been

characterized by high glycogen degradation and increased lactate accumulation (Zhu, Ruusunen,

Gusella, Zhou, & Puolanne, 2011).

Specifically, poor color, water-holding capacity, and texture resulting from PSE meat has

been attributed to loss of protein functionality/solubility and degradation due to rapid pH decline.

Myofibrillar and sarcoplasmic proteins play a key role the capacity of meat to retain moisture.

This greatly affects texture, cohesiveness, and light reflectance. Due to the dipolar nature of

water molecules, they interact, or bind to the residues of these charged proteins and are held

there. However, if muscle is exposed to moderate pH decline and high temperature for a long

period of time, protein denaturation can be significantly greater than if severe pH and

temperature conditions exist for a short period of time (Offer, 1991). This denaturation results in

13

a less birefringent myofibril. As a result, light back-scattering increases due to altered protein

structure, resulting in a paler meat color (Swatland, 2008).

Under these conditions, extractability of proteins is often reduced due to denaturation and

degradation. The inability to extract proteins due to loss of functionality is detrimental to

production of further processed products. For instance, pectoralis muscles exposed to

temperatures of 40◦C during postmortem metabolism, present a sarcoplasmic protein content

lower than groups exposed to 0 and 20◦C temperatures (Zhu et al., 2011). Similarly, the

myofibrillar fraction was higher in protein content (Zhu et al., 2011). Pietrzak et al. (1997)

showed that this is likely due to precipitation of the sarcomplasmic protein phosphorylase on the

myofilaments in turkey. On the contrary, Pietrzak et al. (1997) found through fluorescent anti-

phosphorylase staining that phosphorylase accumulates in specific locations on the sarcomere:

the Z lines and middle of the A band. This suggests that myosin is not coated by precipitating

sarcoplasmic proteins on the myofibril, but that it is denatured, resulting in lowered

extractability. Irving, Swatland, and Millman (1989) and Offer (1991) hypothesized that during

rigor, myosin head shrinkage and denaturation leads the thick and thin filaments becoming more

closely associated.

Stun Method and Blood Removal

Stun technique can have major effects on bleed out efficiency. For many slaughter

facilities, low voltage electrical stun is the primary method of humane slaughter.

Unconsciousness is achieved by a low voltage (~13 to 15 mA) current flowing through a water

bath (Alvarado, Richards, O’Keefe, & Wang, 2007). Although studies have shown that initial

14

blood loss is affected, total blood loss after a 90-120s exanguination period is no different from

high voltage stun methods popular in the European Union (Gregory, 1993; Papinaho & Fletcher,

1995). In the last decade, the use of a carbon dioxide gas mixture has become a widespread

alternative. The gas is generally comprised of 30-55% CO2 and an argon filler (Mohan &

Gregory, 1990). These gases work in unison to render the bird unconscious. Specifically, carbon

dioxide lowers the pH of the cerebrospinal fluid (Alvarado et al., 2007), while argon displaces air

(Eisele, Eger 2nd, & Muallem, 1966). The result is an anesthetic and anoxic response. The

differences in these stun methods have been investigated in regards to bleed out efficiency and

residual hemoglobin content in the breast muscle.

When comparing these two stun methods Alvarado et al. (2007) determined that the

hemoglobin concentration in the chicken breast was lower in birds stunned with CO2 gas (8.53

+/- 1.38 µmole/kg) as compared to low-voltage electrical stun (8.72 +/- 1.66 µmole/kg), although

not statistically significant. These data showed only a small reduction (13-17%) in hemoglobin

content from birds bled and not bled. This small difference in hemoglobin content may be

explained by the lack of force available to drive blood out of muscle capillary beds after a rapid

blood pressure drop due to exanguination (Alvarado et al., 2007). Mohan et al. (1990) reported

differences in bleed out time of approximately 35 s when comparing CO2 and electrical stun

methods; 155s bleed out time and 120 s were determined respectively. This correlated with a

marginally heavier carcass weight for the CO2 stunned birds. These data suggested that a bleed

out efficiency was slightly impaired by the CO2 stun, resulting in a higher muscle hemoglobin

content. However, Raj, Grey, Audsely, and Gregory (1990) demonstrated that a* (redness) color

values were greater in electrically stunned birds. Perhaps the elevated a* values were due to

hemorrhages in the muscle from a traumatic stun method. These studies have sufficiently shown

15

that variation in stun method affects bleed out efficiency and, in turn, the hemoglobin

concentration remaining in the muscle. Enough residual hemoglobin in the muscle will affect

color and ultimately meat quality.

Summary

Although meat color is known as the most important quality attribute to consumers, many

factors influence its development. Inherently, heme pigment content, color stability, and muscle

fiber type composition are contributors. Environmentally, temperature, processing techniques

and inefficient bleed out can result in aberrant quality. Recently, a visible two-toning has

become evident in the turkey pectoralis major muscle. With poultry consumption popular, the

direct cause of this aberration must be investigated.

16

References

Addis, P. B., 1986. Muscle as Food. P. J. Betchel, ed. Academic Press, New York, NY

Press, New York, NY.

Alvarado, C., Richards, M., O’Keefe, S., & Wang, H. (2007). The effect of blood removal on

oxidation and shelf life of broiler breast meat. Poultry science, 86(1), 156-161.

Antonini, E. (1965). Interrelationship between structure and function in hemoglobin and

myoglobin. Physiological reviews, 45(1), 123-170.

Barbut, S. (1997). Problem of pale soft exudative meat in broiler chickens. British Poultry

Science, 38(4), 355-358.

Bekhit, A. E.-D. A., Geesink, G. H., Ilian, M. A., Morton, J. D., Sedcole, R., & Bickerstaffe, R.

(2003). Particulate metmyoglobin reducing activity and its relationship with meat color.

Journal of agricultural and food chemistry, 51(20), 6026-6035.

Bendall, J. (1973). Postmortem changes in muscle. The structure and function of muscle, 2(Part

2), 244-309.

Berri, C., Wacrenier, N., Millet, N., & Le Bihan-Duval, E. (2001). Effect of selection for

improved body composition on muscle and meat characteristics of broilers from

experimental and commercial lines. Poultry Science, 80(7), 833-838.

Briskey, E. J., CASSENS, R. G., & TEAUTMAN, J. (1966). The physiology and biochemistry

of muscle as a food. The physiology and biochemistry of muscle as a food.

Brooks, J. (1937). Color of meat. Food industry, 9, 707.

Clydesdale, F., & Francis, F. (1972). Color measurement of foods. xXVII. the chemistry of meat

color. Food Prod Develop.

Cornish, D., & Froning, G. (1974). Isolation and purification of turkey heme proteins. Poultry

science, 53(1), 365-377.

Eisele, J., Eger 2nd, E., & Muallem, M. (1966). Narcotic properties of carbon dioxide in the dog.

Anesthesiology, 28(5), 856-865.

Food and Agriculture Organization of the United Nations. (2012). Statistics.

Fletcher, D. (2002). Poultry meat quality. World's poultry science journal, 58(02), 131-145.

Fox Jr, J. B. (1966). Chemistry of meat pigments. Journal of Agricultural and Food Chemistry,

14(3), 207-210.

FREMERY, D., & POOL, M. F. (1960). BIOCHEMISTRY OF CHICKEN MUSCLE AS

RELATED TO RIGOR MORTIS AND TENDERIZATIONa. Journal of Food Science,

25(1), 73-87.

Gregory, N. (1993). Causes of downgrading in chickens, turkeys and ducks. Broiler Ind, 56, 42-

45.

Grunert, K. G. (2006). Future trends and consumer lifestyles with regard to meat consumption.

Meat Science, 74(1), 149-160.

Hartman, F. A. (1961). Locomotor mechanisms of birds (Vol. 143): Smithsonian institution.

Henckel, P., Oksbjerg, N., Erlandsen, E., Barton-Gade, P., & Bejerholm, C. (1997). Histo-and

biochemical characteristics of the longissimus dorsi muscle in pigs and their relationships

to performance and meat quality. Meat Science, 47(3), 311-321.

Irving, T., Swatland, H., & Millman, B. (1989). X-ray diffraction measurements of myofilament

lattice spacing and optical measurements of reflectance and sarcomere length in

commercial pork loins. Journal of Animal Science, 67(1), 152-156.

Johnson, B. (1974). Chilled vacuum-packed beef. CSIRO Food Res. Quart, 34, 14.

17

Joseph, P., Suman, S., Li, S., Beach, C., & Claus, J. (2010). Mass spectrometric characterization

and thermostability of turkey myoglobin. LWT-Food Science and Technology, 43(2),

273-278.

Joseph, P., Suman, S. P., Li, S., Claus, J. R., Fontaine, M., & Steinke, L. (2011). Primary

structure of turkey myoglobin. Food Chemistry, 129(1), 175-178.

Khan, A. (1971). Effect of temperature during post‐mortem glycolysis and dephosphorylation of

high energy phosphates on poultry meat tenderness. Journal of Food Science, 36(1), 120-

121.

Klont, R., Brocks, L., & Eikelenboom, G. (1998). Muscle fibre type and meat quality. Meat

Science, 49, S219-S229.

Kranen, R., Van Kuppevelt, T., Goedhart, H., Veerkamp, C., Lambooy, E., & Veerkamp, J.

(1999). Hemoglobin and myoglobin content in muscles of broiler chickens. Poultry

science, 78(3), 467-476.

Lee, S., Joo, S., & Ryu, Y. (2010). Skeletal muscle fiber type and myofibrillar proteins in

relation to meat quality. Meat science, 86(1), 166-170.

Mancini, R., & Hunt, M. (2005). Current research in meat color. Meat science, 71(1), 100-121.

Matthews, D., & Fox, E. (1976). The physiological basis of physical education and athletics.

Philadelphia, WB Saunders, 135-149.

Mohan, R. A., & Gregory, N. (1990). Investigation into the batch stunning/killing of chickens

using carbon dioxide or argon-induced hypoxia. Research in Veterinary Science, 49(3),

364-366.

Nishida, J., & Nishida, T. (1985). Relationship between the concentration of myoglobin and

parvalbumin in various types of muscle tissues from chickens. British poultry science,

26(1), 105-115.

Offer, G. (1991). Modelling of the formation of pale, soft and exudative meat: Effects of chilling

regime and rate and extent of glycolysis. Meat Science, 30(2), 157-184.

Papinaho, P., & Fletcher, D. (1995). Effect of stunning amperage on broiler breast muscle rigor

development and meat quality. Poultry science, 74(9), 1527-1532.

Pietrzak, M., Greaser, M., & Sosnicki, A. (1997). Effect of rapid rigor mortis processes on

protein functionality in pectoralis major muscle of domestic turkeys. Journal of Animal

Science, 75(8), 2106-2116.

Pikul, J., Niewiarowicz, A., & Kupijaj, H. (1986). The cytochrome c content of various poultry

meats. Journal of the Science of Food and Agriculture, 37(12), 1236-1240.

Pirko, P. C., & Ayres, J. C. (1957). Pigment changes in packaged beef during storage. Paper

presented at the Food Technology.

Raj, A. M., Grey, T., Audsely, A., & Gregory, N. (1990). Effect of electrical and gaseous

stunning on the carcase and meat quality of broilers. British Poultry Science, 31(4), 725-

733.

Seideman, S., Cross, H., Smith, G., & Durland, P. (1984). Factors associated with fresh meat

color: a review. Journal of Food Quality, 6(3), 211-237.

Senauer, B. (2001). The food consumer in the 21st century: new research perspectives: Retail

Food Industry Center, University of Minnesota.

Sosnicki, A., Greaser, M., Pietrzak, M., Pospiech, E., & Sante, V. (1998). PSE‐LIKE

SYNDROME IN BREAST MUSCLE OF DOMESTIC TURKEYS: A REVIEW.

Journal of Muscle Foods, 9(1), 13-23.

18

Suman, S. P., & Joseph, P. (2013). Myoglobin chemistry and meat color. Annual review of food

science and technology, 4, 79-99.

Swatland, H. (2008). How pH causes paleness or darkness in chicken breast meat. Meat science,

80(2), 396-400.

Wiskus, K., Addis, P., & Ma, R. T. (1976). Distribution of βR, αR and αW fibers in turkey

muscles. Poultry Science, 55(2), 562-572.

Woelfel, R., Owens, C., Hirschler, E., & Sams, A. (1998). The incidence and characterization of

pale, soft and exudative chicken meat in a commercial plant. Poultry Sci, 77(Suppl 1), 62.

Zhu, X., Ruusunen, M., Gusella, M., Zhou, G., & Puolanne, E. (2011). High post-mortem

temperature combined with rapid glycolysis induces phosphorylase denaturation and

produces pale and exudative characteristics in broiler Pectoralis major muscles. Meat

Science, 89(2), 181-188.

19

Chapter 2. Pectoralis muscle of turkey displays divergent function as

correlated with meat quality

Abstract

Fresh turkey meat color is determined by many factors which include muscle fiber type

composition and heme protein concentrations. These factors either are affected by or influence

biochemical events occurring postmortem. Deviations in the processing environment can also

result in aberrant fresh meat quality and may ultimately change the quality characteristics of

further processed products. Our objective was to describe the underlying cause and significance

of the two-toning color defect in fresh turkey breast. In the first experiment, pectoralis muscles

were collected and analyzed using image processing to characterize fresh turkey color. In the

second experiment, time from stun to exsanguination was tested as a promoter of fresh turkey

color. Results showed that the turkey breast possesses two distinct lobes that differ in pH, drip

loss, energy metabolism and muscle fiber type composition. Results also showed that the time

from stun to exsanguination enhanced fresh turkey color (P < 0.05) by mitigating the differences

between the two lobes. These results show inherent differences in breast muscle and processing

conditions interact to establish variations in fresh turkey color.

Introduction

In recent decades, the poultry industry has witnessed an increase in product sales due to a

consumer desire for convenience. In 2013, the impact of turkey sales alone contributed $4.83

billion dollars to the United States meat industry from over 7 billion pounds of meat products

(USDA NASS, 2015). With the popularity of value-added products, numerous meat quality

problems have surfaced, however (Qiao, Fletcher, Smith, & Northcutt, 2001; Rathgeber, Boles,

& Shand, 1999) the problems have been attributed to undesirable color and poor water-holding

capacity (Sosnicki & Wilson, 1991). Two-toning is a meat quality defect often characterized by

20

an inconsistent or dramatic color change visualized within or between muscles. Recently,

industry partners have observed an inconsistent color pattern or two-toning in the face of fresh

turkey breasts that translate into inconsistencies in further processed products like turkey deli

rolls. Because color is the primary determinant of meat purchasing decisions (Mancini & Hunt,

2005; Mitsumoto, O’Grady, Kerry, & Buckley, 2005; Ramírez & Cava, 2007), inconsistencies in

the color of deli rolls makes them less desirable to consumers.

Underlying differences in muscle and changes in postmortem metabolism can contribute

to meat color development. Inherent muscle color differences commonly result from variation in

heme pigment content and muscle fiber type composition (Kranen et al., 1999). The primary

pigment that gives meat its red color is myoglobin though hemoglobin and cytochrome c

contribute as well (Mancini & Hunt, 2005). Muscle myoglobin abundance is dictated by the

fiber type profile found in the muscle (Choe et al., 2008; Dransfield & Sosnicki, 1999; Hunt &

Hedrick, 1977; Seideman, Cross, Smith, & Durland, 1984). Muscle fiber types have been

classified based upon contractile speed and primary metabolism (Ozawa et al., 2000; Peter,

Barnard, Edgerton, Gillespie, & Stempel, 1972; Ryu & Kim, 2005). While these underlying

muscle properties can be different, events of postmortem metabolism may affect color

development as well.

The current experiment was designed to characterize the two-toning nature of the

pectoralis major muscle. The objective was to determine if inherent muscle differences or

postmortem metabolic events are responsible for the development of the two-toning phenomenon

visualized by consumers. The second goal was to determine if the stun to exsanguination time

enhanced the two-toning quality defect.

21

Materials and Methods

Characterizing the two-toning defect of the pectoralis major muscle

Muscle sampling. The pectoralis major muscle of 48, 20-wk old tom turkeys were obtained

from a commercial processing facility. Breast muscles were chosen at random from immediately

after the de-boning. Muscle samples (~5g) were collected from the large and small lobes of the

pectoralis major as outlined in Figure 1. Samples were frozen in liquid nitrogen and stored at -

80◦C for further analysis.

Color measurement and image analysis. The epimysium of the pectoralis major muscle was

removed using a knife with approximately 4 mm of the muscle surface followed by a 10 min

bloom period. In attempt to quantify color contrasts existing in two-toned turkey breast meat, a

Konica-Minolta Chroma Meter CR-200 (Ramsey, NJ) was first used. Unfortunately, sampling

limitations of this technology surfaced when lightness (L*), redness (a*), and yellowness (b*)

color value differences could not be detected by its point measurement design (Kang, East, &

Trujillo, 2008). Due to the muscle’s non-homogenous distribution of color and the colorimeter’s

inability to account for localized variability (Antonelli et al., 2004), two-toning was masked.

To solve this problem, CIELAB color values for the entire breast were determined using

a image processing. Using this approach, breast images were collected and quantified in an RGB

color space. RGB values were then converted to values in CIELAB color space. A light stand

(Bencher, Antioch, IL) with Sylvania DVY 650W 120V halogen/tungsten bulbs (Interlight,

Hammond, IN) was used to capture images of the pectoralis major muscle with a Powershot

SX170 IS camera (Canon, Melville, NY). The camera was equipped with a 16 megapixel sensor

with a 28 mm wide-angle lens. All lighting selections were in accordance with guidelines

22

established by the American Meat Science Association Guidelines for meat color evaluation

(AMSA, 2012). Resulting images were cropped in Paint.NET (dotPDN, Kirkland, WA) and

analyzed using image processing software (MATLAB 2014a, MathWorks, Natick, MA). A

software script was written to convert red, green, and blue matrices (RGB), or an M x N x 3

image, to CIELAB color space (Ruszon, 1997). A second software script calculated the statistical

parameters for the resulting color distribution: mean, standard deviation, minimum, maximum,

first quartile, third quartile, and skew. To minimize outliers, thresholding was used to eliminate

values representative of fat, light scattering, and background. The script eliminated these values

and re-calculated the initial statistical parameters. Approximately four percent of the total pixels

were removed with this function.

To normalize the data to a known color value range, images of the Japanese Pork Color

Standards were captured under the same lighting conditions. Color standards were also measured

in triplicate with Konica-Minolta Chroma Meter CR-200 (Ramsey, NJ) to obtain color

measurements in the CIELAB color space. Color standard images were similarly converted from

RGB to the CIELAB color space. The delta change between the Konica-Minolta Chroma Meter

CR-200 and imaging color values were calculated and averaged for each standard. The delta

value was subtracted from all image values in order to normalize the MATLAB data output from

the pectoralis images to the colorimeter data.

Muscle classification. To evaluate the two-toning, images of the pectoralis major muscles

collected were visually categorized by an untrained academic panel of five individuals as single-

toned or two-toned. Images consistently classified by the panelists as single-toned or two-toned

were assigned to that category. Muscle images were analyzed for L* and a* color value

23

differences. Panel classifications were confirmed quantitatively by a delta change between the

large and small lobe.

pH analysis. Muscle pH was determined from the adapted method of Bendall (1973). Frozen

muscle tissue was homogenized in 5 mM sodium iodoacetate and 150 mM KCl (pH 7.0) at a 1:8

ratio (wt/vol) using a Tissue Lyser II (Qiagen, Valencia, CA). Tissue homogenates were heated

at 25◦C, centrifuged and measured using an Orion Ross Ultra pH electrode (Fisher Scientific,

Pittsburgh, PA).

Centrifugal drip collection. Exudates were collected from muscle samples with a modified

procedure of Bouton, Harris & Shorthose (1971). Muscle samples (~16 g) were collected in

triplicate, weighed and placed in a 50 mL conical tube and centrifuged at 1500 x g for 60 min.

Following centrifugation, samples were removed, lightly blotted and re-weighed. Drip loss was

calculated in triplicate as a percentage of the original weight of the sample 24 h postmortem.

Protein content. Frozen muscle samples were powered in liquid nitrogen for both soluble and

total protein content. Soluble proteins were extracted by homogenization from a muscle sample

using 0.025M potassium phosphate buffer (pH 7.2) according to Joo, Kauffman, Kim, and Park

(1999) and measured using the Pierce BCA Assay Kit (Thermo Scientific, Rockford, IL). Total

muscle proteins were solubilized using a buffer containing 8 M urea, 2 M thiourea, 3% SDS

(wt/vol), 75 mM diotheothreitol, 0.05 M Tis-HCl (pH 6.8) and heated at 95◦C for 5 min (Warren,

Krzesinski and Greaser, 2003). Total protein concentration was determined using the Bio-Rad

RC DC protein assay (Bio-Rad, Hercules, CA).

24

Sarcomere length determination. Myofibrils were purified from frozen tissue samples using

the procedure described by Weaver, Bowker, and Gerrard (2008). Tissue samples were

homogenized using a Polytron PT-MR 2100 (Kinematica Ag, Littau-Lucerne, Switzerland) in

rigor buffer containing 75 mM KCl, 10 mM imidazole, 2 mM MgCl2, 2 mM ethylene glycol

tetracetic acid, 1 mM NaN3 (pH 7.2). Samples were centrifuged for 1 min and the supernatant

was discarded. Fresh rigor buffer was added to the sample pellet and homogenized for 10 sec.

Steps were repeated three times. Samples were diluted 1:1 with glycerol and rigor buffer

solution. Samples were stored at -20◦C until analysis.

Samples were diluted 1:4 and applied to a microscope slide with coverslip. Images of ten

myofibrils per sample were captured using a Nikon Inverted Microscope Eclipse Ti-E/B

equipped with a DS-L3 camera (Chiyoda, Tokyo, Japan) and NIS Elements (version 4.13)

software. The length of five consecutive sarcomeres from each image was determined by

measuring from Z line to Z line using a 10 μm ruler. The average sarcomere length for each

sample was calculated based on randomly selected ten images.

Muscle metabolite analysis. To prepare samples for metabolite analysis, two Eppendorf tubes

with 100 mg of frozen muscle was powdered using liquid nitrogen. One sample was subjected to

acid hydrolysis at 95◦C for glycogen analysis. The second sample was homogenized in 0.5 M

perchloric acid and centrifuged. The supernatant was removed and neutralized with 2 M KOH.

Muscle glucose, glucose-6-phosphate, glycogen, and lactate concentrations were determined

using enzymatic methods (Bergmeyer, 1984) modified for a 96-well microplate assay

(Hammelman et al., 2003). Glycolytic potential was calculated according to Monin and Sellier

(1985): glycolytic potential [μmole/g] = 2(glucose+glucose-6-phosphate+glycogen) + lactate.

25

Gel electrophoresis and protein blotting. Total muscle proteins of a 50 mg sample were

solubilized in a buffer containing 8 M urea, 2 M thiourea, 3% SDS (wt/vol), 75 mM

diotheothreitol, 0.05 M Tis-HCl (pH 6.8) and heated at 95◦C for 5 min (Warren, Krzesinski and

Greaser, 2003). Total protein concentration was determined using the Bio-Rad RC DC protein

assay (Bio-Rad, Hercules, CA). Samples were diluted to contain 8 μg/μL protein using the

solubilization buffer (Warren, Krzesinski and Greaser, 2003). For electrophoresis, 240μg and 60

μg/lane of protein were loaded for hemoglobin and lactate dehydrogenase detection, respectively.

Proteins were electrophoresed at 60 V for 20 min and 120 V for 2 h 45 min using a 15%

polyacrylamide gel and transferred to nitrocellulose membrane at 50 V for 1 h. To prevent non-

specific antibody binding, membranes were blocked in PBS with 3% casein for 1 h. Primary

antibody incubations were performed in blocking buffer in a cold room overnight and were

washed with TBS-T containing 0.05% Tween three times after incubation. Membranes were

incubated overnight with the primary antibody as follows: anti-hemoglobin beta 1:200 (LS-

C294450; Lifespan Biosciences, Seattle, WA) and anti- lactate dehydrogenase A 1:2000 (NBP-

1-48336; Novus Biologicals, Littleton, CO). Secondary antibody donkey anti-rabbit IgG (926-

32213; Li-cor Biosciences, Lincoln, NE) was applied at 1:5000 dilution, incubated for 1 h at

room temperature, and washed three times. Antibody binding was determined by exposure to

fluorescence.

Following quantification, membranes were re-blocked for 1 h in PBS with 3% casein.

Beta actin abundance was used as a loading control on all membranes. Membranes were

incubated in primary and secondary membranes as follows: anti-beta actin 1:2000 (NB600-501;

Novus Biologicals, Littleton, CO) and goat anti- mouse 1:5000 (926-6818; Li-cor Biosciences,

26

Lincoln, NE). All membranes were re-scanned for fluorescent detection. Lactate dehydrogenase

membranes were stained Ponceau-S to confirm equal loading.

RNA extraction, reverse transcription, and PCR amplification. Total RNA was extracted

using the Direct-zol RNA Mini Prep Kit (Zymo Research, Irvine, CA). cDNA was generated

using the High Capacity cDNA Reverse Transcriptase Kit (Applied Biosystems, Waltham, MA)

and a DNA Engine PTC-200 Peltier Thermal Cycler (Biorad, Hercules, CA). For determination

of myoglobin mRNA abundance, a 7500 Fast Real-Time PCR System (Applied Biosystems,

Waltham, MA) was used with SYBR green detection. Myoglobin and 18S rRNA primers were

designed using the sequences provided by NCBI accession code XM_003202347.2 and

AJ419877, respectively. The myoglobin primer was (5’3

’) AGTGGAGGCCGATATTGCTG

(Forward) and TCACCTCCGGCTATAACGAC (Reverse) and was synthesized by Eurofins

MWG Operon (Huntsville, AL). Expression of the 18S rRNA housekeeping gene was detected

using primers (5’3

’) CCGTCGTAGTTCCGACCATAA (Forward) and

GAGGGTCATGGGAATAACG (Reverse). All primers were diluted to a 10 μM concentration

prior to use. For PCR quantification, the Ct endpoint was used. Gene expression was determined

by the 2-∆∆Ct

method.

Tom selection. A truck containing approximately 1000 birds were randomly selected and

stunned with carbon dioxide. Following stunning, 20 tom turkeys were randomly selected for

slaughter and identified as the ‘first’ treatment group coming off the truck. A second set of 20

toms were randomly selected from the ‘last’ 100 birds shackled off the truck and harvested.

Approximately 1.5 h post-chill, identified birds were removed from the processing line and one

27

pectoralis major muscle was removed for further analysis. This procedure was repeated on a

second day for a total of two replicates (trucks).

Statistical analysis. All muscle property parameters were subjected to an ANOVA test using

JMP Pro 11(SAS Institute, Cary, NC) statistical software as a completely randomized designed.

Lobe served as the main effect. Color measurements were analyzed as a two-by-two full factorial

design. Lobe and color condition served as main effects for characterization color measurements.

Time and lobe served as main effects for the study component evaluating time effect on color,

Truck/day was imputed in the model as a block. Significant differences of means were

determined using an alpha level of 0.05. Significant results were further compared using the LS

Means Student’s T Test.

Results and Discussion

Characterizing the two-toning defect of the pectoralis major muscle

Initially, there was no interaction between lobe and the color assignment of each

pectoralis major muscle classified by an untrained panel. Figure 2 shows L* color values and the

variability for the main effects of lobe and condition. Figure 2a indicates that the small lobe is

darker (P = 0.0007), while the two lobes are consistently variable in color (Fig. 2b). Analysis of

the single-toned and two-toned classifications made by the untrained panel suggests that the

single-toned group displays a higher degree of lightness (Fig. 2c) and similar color variation to

the two-toned group (Fig. 2d). Assessment of redness showed the small lobe (Fig. 3a) and two-

toned population to appear redder (Fig. 3c), while both were more variable (Fig. 3b) in color than

the large lobe (Fig. 3b) and single-toned population (Fig. 3d). These data show that an untrained

28

panel can detect differences in fresh turkey meat color and suggest the severity of two-toning in

turkey breast muscle is more a function of the smaller lobe being darker than the large lobe being

lighter.

In support of the aforementioned, redness (a*) values are shown in Fig. 3. Redness values

for those breasts muscles designated two-toned were greater (P = 0.0005) than that of the single-

toned group. The nature of redness, however, was more variable, as suggested by a greater

variation in two-toned and smaller lobes (Fig. 3b and 3d). These data confirm that the smaller

lobe of fresh turkey breast muscles plays a bigger role in color variation than would be

anticipated. To understand the cause of color variation in two-toning, a full characterization of

the large and small lobe muscle properties was performed.

Heme proteins such as myoglobin, hemoglobin, and cytochrome c give meat its color.

Postmortem, myoglobin is the major pigment responsible for meat color. However, other heme

pigments such as hemoglobin become more important in color formation when processing

abnormalities such as inefficient bleed-out occur (Alvarado et al., 2007; Bourbab & Idaomar,

2012; Warriss, 1977). With the darker (Fig. 2a) and redder (Fig. 3a) appearance of the small

lobe, both myoglobin and residual hemoglobin content were measured. Cytochrome c was not

investigated due to its low concentration in glycolytic muscles and its minor role in fresh meat

color development (Pikul et al., 1986)

Initial attempts to quantify myoglobin in the large and small lobe of the pectoralis major

muscle using western blot techniques were unsuccessful. This result is not surprising because

Kranen et al. (1999) was unsuccessful in detecting myoglobin in the pectoralis major muscle due

to the muscle’s predominating glycolytic nature. To circumvent this issue, myoglobin

abundance was evaluated in both lobes using real-time quantitative PCR. Results indicated that

29

myoglobin expression was indeed greater (P = 0.0121) in the small lobe than the large lobe (Fig.

4). These data suggest muscle fiber type in the small lobe may differ from the large lobe. To

eliminate the possibility of blood-based heme contamination, hemoglobin was determined.

Results indicated no differences (P = 0.99) in hemoglobin content between the large and small

lobe (Fig. 5). These data suggest differences between lobes are not due to incomplete

exsanguination of lobes but rather differences in muscle fiber type characterisitics.

Because myoglobin content is highly correlated with differences in muscle fiber type

composition and, thus, postmortem metabolism (Kranen et al., 1999), we determined the ultimate

pH and glycolytic metabolites between the large and small lobes. The small lobes had a higher (P

< 0.0001) ultimate pH (5.97 ± 0.02) compared to the large lobe (5.82 ± 0.03; Figure 6).

Moreover, the small lobe contained lower lactate (P < 0.0001) and glycogen (P = 0.0078) and as

such, a lower glycolytic potential (P < 0.0001; Table 1). Lower glycolytic potential is a measure

of the metabolites that may have come from the breakdown of glycogen. Glycolytic potential

would also be greater with in glycolytic, fast-contracting muscle fibers. Glycolytic potential

would also be inversely correlated with oxidative metabolism. Fewer glycogen stores are needed

because they rely less on glycolysis and more on the oxidative metabolism (Schwerzmann,

Hoppeler, Kayar, & Weibel, 1989). While the small lobe remains glycolytic, small shifts in

oxidative profile are evident in the small lobe. On the other hand, the resulting increased lactate

accumulation and low ultimate pH of the large lobe suggests an extended postmortem glycolysis.

A greater flux of substrate through the phosphofructokinase (PFK) before its inactivation allows

for pyruvate to be converted to lactate more rapidly (England, Matarneh, Scheffler, Wachet, &

Gerrard, 2014).

30

Differences in metabolite content and pH are likely due to differences muscle fiber type

content. For example, high glycogen stores, high lactate content and low myoglobin content all

contribute to the notion that the large lobe possesses a more glycolytic muscle fiber profile, while

lower glycogen stores and lower lactate accumulation and higher myoglobin content suggest a

more oxidative muscle fiber type composition. These fiber types characteristics strongly suggest

that the large lobe of the turkey pectoralis major is more glycolytic in nature than the small lobe,

which is more oxidative. To further evaluate this difference, lactate dehydrogenase (LDH)

abundance was measured. LDH is an enzyme commonly associated with glycolytic muscle fibers

and is responsible for converting pyruvate to lactate. As expected, LDH abundance of both lobes

was high because the pectoralis major is a glycolytic muscle known to be primarily comprised of

Type IIB muscle fibers (Wiskus et al., 1976). However, the large lobe contained higher (P =

0.012) LDH abundance than the small lobe (Fig. 7).

Poultry meat is known to experience rapid postmortem glycolysis (Rathgeber et al.,

1999). This is largely attributed to the glycolytic nature of their muscle fibers. Sustained elevated

temperatures in the muscle during postmortem glycolysis and pH decline can result in a loss of

protein functionality (El Rammouz, Babile, & Fernandez, 2004; Zhu et al., 2011). To determine

if protein denaturation occurred during processing, parameters such as percent drip-loss, and

protein content were measured.

Drip-loss of the large and small lobe (Fig. 8) indicated that the small lobe released 2.3%

± 0.13 of its total mass as purge, while the large lobe purged 4.9% ± 0.28 (P < 0.0001). This

difference was confirmed by the soluble protein concentration of the two groups (Fig. 9a). A

higher soluble protein content was detected in the small lobe group as compared to the large lobe

(P = 0.0027). Typically, sarcoplasmic proteins are readily soluble in buffers of low salt content.

31

If only low concentrations are detected in the soluble protein fraction, then some have likely lost

functionality. Consequently, they are likely partially lost in the form of purge (Swatland, 2008).

No differences were detected in total protein content (Fig. 9b) between the large and small lobe

groups (P = 0.8889). This was expected as total protein content is remains constant within or

between different skeletal muscles. Combined, the higher percent drip-loss and lower soluble

protein concentration of the large lobe group compared to the small lobe group suggests that a

greater number of proteins are becoming denatured postmortem. As a result, the large lobe is

potentially experiencing a greater loss of protein functionality. This protein denaturation may

partially explain two-toning phenomenon. Swatland (2008) reported that L* values are

significantly affected by denaturation of sarcoplasmic proteins, resulting in an increased amount

of extracellular water. Thus, light is greater due to an alteration in protein structure (Swatland,

2008). The increase in light scatter causes the meat to appear paler. Consequently, the lower

soluble protein content and higher percent driploss of the large lobe indicates that it may have

experienced more protein denaturation due to pH/temperature abuse during postmortem

processing. However, an evaluation of sarcomere length (Fig. 10) showed no statistical

difference between the lobes (P = 0.1423). While the sample size was small (n = 10), these data

suggest temperature abuse played no major role in the postmortem events resulting in the two-

toning effect and loss of protein functionality.

Characterization of the two-toning defect indicated significant differences in the inherent

muscle properties of the large and small lobe of the pectoralis major muscle. The large lobe had

higher metabolite concentrations, indicating a higher glycolytic potential and possible shuttling

of more substrate through glycolysis. This is correlated with the lower ultimate pH of the large

lobe. A greater flux of substrate through PFK to synthesize pyruvate downstream before its

32

inactivation would promote the accumulation of lactate postmortem. Lactate accumulation in the

muscle would result in a lower ultimate pH and ultimately meat quality. On the contrary, lower

metabolite concentrations and glycolytic potential in the small lobe suggests a shift toward

oxidative metabolism. These results combined with distinct differences in LDH abundance and

protein content suggests divergent function of the pectoralis muscle, resulting in the two-toning

phenomenon.

Stun to Exsanguination Study

Our industry partners use whole truck CO2 stunning. Due to lack of automation and the

number of birds on the truck, some birds are not exsanguinated for up to 30 min post-stunning.

Therefore, we determined the effect that the time to exsanguination played in the severity and

incidence of two-toning between the large and small lobes.

Twenty toms were randomly selected from the first 100 birds removed from the transport

truck and served as the first shackled treatment group (First). A second 20 were selected from

the last 100 of the same truck and represented the last shackled treatment group (Last). After

chilling, pectoralis major muscles was removed and subjected to color determination via image

processing. There was no interaction between lobe and time. Consequently, the main effects are

represented in Figure 11. L* color values indicated that the small lobe was darker (Fig. 11a) and

more variable in color than the large lobe (Fig. 11b). Additionally, turkey breast appeared to

darker with time (Fig. 11c), while also displaying more variation (Fig. 11d). Assessment of

degree of redness showed the small lobe being more red (Fig. 12a) and variable (Fig. 12b) in

color than the large lobe. The a* values of the turkey breast became redder (Fig. 11c) and more

varied with time (Fig. 12d). These data suggest the increased variation, as well as the darkening

33

and reddening of the turkey breast with time was a result of the inherently redder small lobe and

its variation.

One possible explanation for the inherently darker and redder small lobe was hemoglobin

content due to concerns of inefficient bled out as a result of 30 min between stunning and

exsanguination. However, hemoglobin content did not differ in the small lobe as evaluated by

time (Fig. 13). Therefore, the source of the reddening effect was not due to incomplete bleeding

caused by time.

Because stun to stick time was the primary interest of this study, breast muscle two-

toning was ignored. However, we chose to classify images of the pectoralis major muscles on

the basis of two-toning using the cut-off established from our initial studies. For example, those

samples lying within two standard deviations of the mean delta change for L* of the single-toned

group were classified as such. Once grouped, the incidence of two-toning was calculated based

on the frequency per treatment group (n = 20). Table 2 shows the frequency of single-toned and

two-toned pectoralis muscles. Curiously, the incidence of two-toned samples increases by nearly

50% in both replications of turkeys sampled last. These data show a greater frequency of two-

toning in CO2 stunned turkey breasts suggesting that the incidence of two-toning may be

partially responsible for darker and redder of the small lobe.

Our initial studies characterizing the two-toning effect suggested the underlying cause of

two toning was the result of differences in inherent muscle properties, which would not change

with time postmortem. To explore the cause of this increased incidence of two-toning further,

color, pH and metabolites were assessed. Since there was no interaction between lobe and time,

the main effects of lobe and time were evaluated independently (Fig. 14). The mean ultimate pH

values of the small lobes were higher than the large lobe (Fig. 14a). However, the ultimate pH

34

was higher in those turkeys last shackled than first shackled (Fig. 14b). This difference in pH

with time could be related to the inherent differences in glycolytic flux between the large and

small lobe. Indeed, the smaller lobe is more oxidative than the large lobe. Consequently, less

substrate was likely shuttled through PFK before shutdown postmortem. As a result, lower

concentrations of flux would occur before PFK inactivation (England et al., 2014). This means

that the ultimate pH remains higher in the small lobe. The lower (Fig. 15a) lactate concentration

of the small lobe irrespective of time (Fig. 15b), indicates that less pyruvate was converted to

lactate by LDH postmortem. If a lower flux of substrate through PFK occurred before its

inactivation, the glycolytic cascade would result in less production of pyruvate. Consequently,

less lactate would accumulate in the muscle postmortem, resulting in a higher ultimate pH.

Analysis of the residual glycogen content and glycolytic potential indicated no interaction

between time and lobe, while there was also no statistical significance to the main effect of time

(Fig. 15d and 15f). All effects were a result of inherent lobe differences. More residual

glycogen was detected in the large lobe than the small lobe (Fig. 15c). Glycolytic potential was

also higher in the large lobe than the small (Fig. 15e). These data are in agreement with earlier

characterizations.

Characterization suggested that the small lobe displayed a shift toward a slightly more

oxidative metabolism, while the large lobe remained more glycolytic. Recall, that it takes

approximately 30 min for a single truck of turkeys to be shackled. The higher ultimate pH of the

turkey breast with time (Fig. 14b) could be due to a higher ultimate pH (Fig. 14a) of the small

lobe and its lower lactate content (Fig. 15a). If so, the higher ultimate pH of the small lobe (Fig.

14) and lower lactate content (Fig. 15a) could be explained by a changed in body temperature of

the bird within the shackling time period. If the small lobe functions under a more oxidative

35

metabolism than the large lobe, then perhaps the lowering of the external body temperature of

turkeys by approximately 6◦C (Fig. 16) slows metabolism. This would results in a higher

ultimate pH and less lactate accumulation.

Conclusions

Taken together, data suggest that the time from stun to exsanguination enhances the two-

toning color defect and that differences in muscle metabolism between the large and small lobe

may explain the color development postmortem.

36

Figure 1. Sampling location of the large and small lobes of the pectoralis major muscle in

adult tom turkeys for experiment one and two. The large and small lobe were differentiated

by the fibrous raphe running medial through the muscle.

37

Figure 2. Lightness (L*) color values and standard deviation as an expression of variability

for the large and small lobes and single-toned and two-toned populations of pectoralis

major muscle of adult tom turkeys. (A) Large and small lobe lightness color values. (B)

Standard deviation as an expression of variability of L* lobe color values. (C) Lightness color

values of single-toned and two-toned populations. (D) Standard deviation as an expression of

variability of L* single-toned and two-toned population color values. Data are LS Means ± SE.

Different subscripts (a, b) indicate significant differences (P < 0.05) between treatments.

38

Figure 3. Redness (a*) color values and standard deviation as an expression of variability

for the large and small lobes and single-toned and two-toned populations of pectoralis

major muscle of adult tom turkeys. (A) Large and small lobe redness color values. (B)

Standard deviation as an expression of variability of a* lobe color values. (C) Redess color

values of single-toned and two-toned populations. (D) Standard deviation as an expression of

variability of a* single-toned and two-toned population color values. Data are LS Means ± SE.

Different subscripts (a, b) indicate significant differences (P < 0.05) between treatments.

39

Figure 4. Relative myoglobin mRNA abundance of the small and large lobes of the

pectoralis major muscle in adult tom turkeys. Data are LS Means ± SE. Different subscripts

(a,b) indicate significant differences (P < 0.05) between treatments.

40

Figure 5. Hemoglobin content of the small and large lobes of the pectoralis major muscle in

adult tom turkeys. Data are LS Means ± SE. Different subscripts (a,b) indicate significant

differences (P < 0.05) between treatments. Control = turkey blood; Large = large lobe; Small =

small lobe.

41

Figure 6. Ultimate pH of the small and large lobes of the pectoralis major muscle in adult

tom turkeys. Data are LS Means ± SE. Different subscripts (a,b) indicate significant differences

(P < 0.05) between treatments.

42

Large Lobe Small Lobe

P-

Value

Lactate (μmole/g) 104 ± 1.51 94.3 ± 1.82 0.0001

Residual Glycogen (μmole/g) 8.91 ± 0.41 7.25 ± 0.40 0.0078

Glycolytic Potential (μmole/g) 138 ± 1.95 118 ± 1.87 0.0001

Table 1. Metabolite concentrations and glycolytic potential of the large and small lobe of

the pectoralis major muscle of adult tom turkeys. Data are LS Means ± SE.

43

Figure 7. Lactate dehydrogenase content (LDH) of the large and small lobes of the

pectoralis major muscle in adult tom turkeys. Data are LS Means ± SE. Different subscripts

(a,b) indicate significant differences (P < 0.05) between treatments. Control = pooled mouse led

muscle; Large = large lobe; Small = small lobe.

44

Figure 8. Drip-loss of small and large lobes of pectoralis major muscle in adult tom turkeys.

Data are LS Means ± SE. Different subscripts (a,b) indicate significant differences (P < 0.05)

between treatments.

45

Figure 9. Soluble and total protein concentration of the large and small lobe of the

pectoralis major muscle in adult tom turkeys. (A) Soluble protein concentration of the large

and small lobe. (B) Total protein concentration of the large and small lobe. Data are LS Means ±

SE. Different subscripts (a,b) indicate significant differences (P < 0.05) between treatments.

46

Figure 10. Sarcomere length of the large and small lobes of the pectoralis major muscle in

adult tom turkeys. Data are LS Means ± SE. Different subscripts (a,b) indicate significant

differences (P < 0.05) between treatments.

47

Figure 11. Lightness (L*) color values and variability for the large and small lobes and of

the pectoralis major muscle of adult tom turkeys shackled first and last from a transport

truck. (A) Large and small lobe lightness color values. (B) Standard deviation as an expression

of variability of L* lobe color values. (C) Lightness color values of first and last shackled

treatment groups. (D) Standard deviation as an expression of variability of L* color values for

first and last shackled treatment groups. Data are LS Means ± SE. Different subscripts (a, b)

indicate significant differences (P < 0.05) between treatments.

48

Figure 12. Redness (a*) color values and variability for the large and small lobes and of the

pectoralis major muscle of adult tom turkeys shackled first and last from a transport truck.

(A) Large and small lobe lightness color values. (B) Standard deviation as an expression of

variability of a* lobe color values. (C) Redness color values of first and last shackled treatment

groups. (D) Standard deviation as an expression of variability of a* color values for first and last

shackled treatment groups. Data are LS Means ± SE. Different subscripts (a, b) indicate

significant differences (P < 0.05) between treatments.

49

Figure 13. Hemoglobin content of the small lobe of the pectoralis major muscle in adult tom

turkeys sampled first and last from a transport truck. Data are LS Means ± SE. Different

subscripts (a,b) indicate significant differences (P < 0.05) between treatments. Control = turkey

blood; First = small lobe from the first shackled treatment group; Last = small lobe from the last

shackled treatment group.

50

Single-Toned Two-Toned Replication

First 80% 20% Truck 1

Last 65% 35% Truck 1

First 75% 25% Truck 2

Last 55% 45% Truck 2

Table 2. Frequency of single-toned and two-toned pectoralis major muscles in tom turkeys

sampled first and last from a transport truck.

51

Figure 14. Ultimate pH of large and small lobes of the pectoralis major muscle and first and

last shackled adult tom turkeys from a transport truck. (A) Lobe pH. (B) Time effect ton pH.

Data are LS Means ± SE. Different subscripts (a,b) indicate significant differences (P < 0.05)

between treatments.

52

Figure 15. Metabolite concentrations and glycolytic potential of the large and small lobe of

the pectoralis major muscle and first and last shackled adult tom turkeys from a transport

truck. (A) Lobe lactate content. (B) Time effect on lactate content. (C) Lobe glycogen content.

(D) Time effect on residual glycogen content. (E) Lobe glycolytic potential. (F) Time effect on

glycolytic potential. Data are LS Means ± SE. Different subscripts (a,b) indicate significant

differences (P < 0.05) between treatments.

53

Figure 16. Effect of stun to stick time on body temperature of adult tom turkeys. Data are

LS Means ± SE. Different subscripts (a,b) indicate significant differences (P < 0.05) between

treatments

54

References

Alvarado, C., Richards, M., O’Keefe, S., & Wang, H. (2007). The effect of blood removal on

oxidation and shelf life of broiler breast meat. Poultry science, 86(1), 156-161.

American Meat Science Association (2012). Meat Color Measurement Guidelines.

Antonelli, A., Cocchi, M., Fava, P., Foca, G., Franchini, G. C., Manzini, D., & Ulrici, A. (2004).

Automated evaluation of food colour by means of multivariate image analysis coupled to

a wavelet-based classification algorithm. Analytica Chimica Acta, 515(1), 3-13.

Bendall, J. (1973). Postmortem changes in muscle. The structure and function of muscle, 2(Part

2), 244-309.

Bergmeyer, H. U. 1974. Methods of enzymatic analysis. Academic Press, New York.

Bouton, P., HARRIS, P. t., & Shorthose, W. (1971). Effect of ultimate pH upon the water‐holding capacity and tenderness of mutton. Journal of food science, 36(3), 435-439.

Choe, J., Choi, Y., Lee, S., Shin, H., Ryu, Y., Hong, K., & Kim, B. (2008). The relation between

glycogen, lactate content and muscle fiber type composition, and their influence on

postmortem glycolytic rate and pork quality. Meat Science, 80(2), 355-362.

Dransfield, E., & Sosnicki, A. (1999). Relationship between muscle growth and poultry meat

quality. Poultry Science, 78(5), 743-746.

El Rammouz, R., Babile, R., & Fernandez, X. (2004). Effect of ultimate pH on the

physicochemical and biochemical characteristics of turkey breast muscle showing normal

rate of postmortem pH fall. Poultry science, 83(10), 1750-1757.

England, E. M., Matarneh, S. K., Scheffler, T. L., Wachet, C., & Gerrard, D. E. (2014). pH

inactivation of phosphofructokinase arrests postmortem glycolysis. Meat science, 98(4),

850-857.

Hammelman, J., Bowker, B., Grant, A., Forrest, J., Schinckel, A., & Gerrard, D. (2003). Early

postmortem electrical stimulation simulates PSE pork development. Meat science, 63(1),

69-77.

Hunt, M., & Hedrick, H. (1977). Profile of fiber types and related properties of five bovine

muscles. Journal of Food Science, 42(2), 513-517.

Joo, S., Kauffman, R., Kim, B., & Park, G. (1999). The relationship of sarcoplasmic and

myofibrillar protein solubility to colour and water-holding capacity in porcine

longissimus muscle. Meat science, 52(3), 291-297.

Kang, S., East, A., & Trujillo, F. (2008). Colour vision system evaluation of bicolour fruit: A

case study with ‘B74’mango. Postharvest Biology and Technology, 49(1), 77-85.

Kranen, R., Van Kuppevelt, T., Goedhart, H., Veerkamp, C., Lambooy, E., & Veerkamp, J.

(1999). Hemoglobin and myoglobin content in muscles of broiler chickens. Poultry

science, 78(3), 467-476.

Mancini, R., & Hunt, M. (2005). Current research in meat color. Meat science, 71(1), 100-121.

Mitsumoto, M., O’Grady, M. N., Kerry, J. P., & Buckley, D. J. (2005). Addition of tea catechins

and vitamin C on sensory evaluation, colour and lipid stability during chilled storage in

cooked or raw beef and chicken patties. Meat Science, 69(4), 773-779.

Monin, G., and P. Sellier. 1985. Pork of low technological quality with a normal rate of muscle

pH fall in the immediate post-mortem period: the case of the Hampshire breed. Meat Sci.

13: 49-63.

Ozawa, S., Mitsuhashi, T., Mitsumoto, M., Matsumoto, S., Itoh, N., Itagaki, K., . . . Dohgo, T.

(2000). The characteristics of muscle fiber types of longissimus thoracis muscle and their

55

influences on the quantity and quality of meat from Japanese Black steers. Meat Science,

54(1), 65-70.

Peter, J. B., Barnard, R. J., Edgerton, V. R., Gillespie, C. A., & Stempel, K. E. (1972). Metabolic

profiles of three fiber types of skeletal muscle in guinea pigs and rabbits. Biochemistry,

11(14), 2627-2633.

Pikul, J., Niewiarowicz, A., & Kupijaj, H. (1986). The cytochrome c content of various poultry

meats. Journal of the Science of Food and Agriculture, 37(12), 1236-1240.

Qiao, M., Fletcher, D., Smith, D., & Northcutt, J. (2001). The effect of broiler breast meat color

on pH, moisture, water-holding capacity, and emulsification capacity. Poultry Science,

80(5), 676-680.

Ramírez, R., & Cava, R. (2007). The crossbreeding of different Duroc lines with the Iberian pig

affects colour and oxidative stability of meat during storage. Meat science, 77(3), 339-

347.

Rathgeber, B., Boles, J., & Shand, P. (1999). Rapid postmortem pH decline and delayed chilling

reduce quality of turkey breast meat. Poultry science, 78(3), 477-484.

Ruszon, M., (1997). RGB to CIELAB from c code by Yossi Rubner.

Schwerzmann, K., Hoppeler, H., Kayar, S. R., & Weibel, E. R. (1989). Oxidative capacity of

muscle and mitochondria: correlation of physiological, biochemical, and morphometric

characteristics. Proceedings of the National Academy of Sciences, 86(5), 1583-1587.

Seideman, S., Cross, H., Smith, G., & Durland, P. (1984). Factors associated with fresh meat

color: a review. Journal of Food Quality, 6(3), 211-237.

Sosnicki, A., & Wilson, B. (1991). Pathology of turkey skeletal muscle: implications for the

poultry industry. Food structure (USA).

Swatland, H. (2008). How pH causes paleness or darkness in chicken breast meat. Meat Science,

80(2), 396-400.

United States Department of Agriculture National Agriculture and Statistics Service. (2013).

Poultry Production and Value 2013 Summary.

Warren, C. M., Krzesinski, P. R., & Greaser, M. L. (2003). Vertical agarose gel electrophoresis

and electroblotting of high‐molecular‐weight proteins. Electrophoresis, 24(11), 1695-

1702.

Weaver, A., Bowker, B., & Gerrard, D. (2008). Sarcomere length influences postmortem

proteolysis of excised bovine semitendinosus muscle. Journal of animal science, 86(8),

1925-1932.

Wiskus, K., Addis, P., & Ma, R. T. (1976). Distribution of βR, αR and αW fibers in turkey

muscles. Poultry Science, 55(2), 562-572.

56

Chapter 3. Future Directions

A multitude of factors influence meat quality development. Loading, transport, wait time

prior to slaughter, and heat or cold exposure are all examples of antemortem stressors that can

affect quality attributes. Furthermore, temperature extremes and time exposures peri- and

postmortem are well documented to have the same effect. The experiments presented here offer

great insight into the inherent properties of the two lobes of the turkey pectoralis major.

Although it is not well understood, the time effect study demonstrated how color is developing

early postmortem. Despite this, there are still missing pieces to the puzzle. An understanding of

how the inherent lobe differences affect functionality as a cooked turkey deli roll is essential. To

add to this, knowing how antemortem factors, such as bird location on the truck, affect

postmortem color development is key. Lastly, mapping of postmortem metabolism in a single

bird at various processing points could provide useful information about how the muscle is

converted to meat and what equipment or processes are the most detrimental to color and protein

functionality as quality parameters.

Initially, a follow up study should be conducted where the specific functional differences

of the two lobes are tested in a cooked turkey deli roll. For instance, a sample size of

approximately twenty single-toned and two-toned breasts should be collected immediately

following de-bone. The two-toned breasts should be separated into large and small lobes along

the membrane running medial through the muscle. Since the single-toned breasts should exhibit

more consistent and desirable quality, they would serve as the control. With no visible color

variation in each individual single-toned breast, the large and small lobes would remain

unseparated. Following this, both the large and small lobe of the two-toned group would be

processed independently to create a formed deli roll product.

57

A determination of percent cook-loss would provide information about additional

proteins that are being denatured during the cook-step, attributing to reduce moisture retention

and perhaps color inconsistency. Despite separating the two lobes to make a more consistent

product, each deli roll could be sliced so that L* and a* color measurements could be determined

using imaging software. This would aid in understanding cooked meat color development.

A second study could focus on two-factors: bird location on the truck and postmortem

metabolic events. For instance, a transport truck could be divided into eight quadrants as

represented by Figure 1. Two birds from each quadrant would be wing-banded per truck,

sampling five trucks. Five trucks would be sampled from these locations in the summer, fall,

and winter seasons. By doing this, season could serve as a block in the statistical model. If there

Rear Front

Figure 1. Sampling method for second study to test the effect of bird location during

transport on meat quality.

is a major block effect, then antemortem temperature extremes play a role in meat quality and

postmortem metabolic events. After wing-banding, these birds would be followed through the

processing facility to de-feathering. Afterward, the right pectoralis major would be removed,

allowed to bloom for ten minutes, and photographed. Two samples from each lobe would be

obtained for driploss, metabolite, and pH determination. After removing the right pectoralis

muscle, birds would be re-shackled until post-chill. At that time, carcasses would be removed

A

A

E

B

F

C

G

D

H

58

from the processing line and the left pectoralis would be removed. Then, it would be

photographed and sampled as previously described. Using the photographs, the breasts could be

categorized as single-toned or two-toned. With these data, it could be determined whether the

bird’s location on the truck during transport is a stress factor contributing to the two-toning color

phenomenon. With the metabolite data, a better understanding of what is happen in the turkey

muscle during postmortem events would provide insight on how the muscle is functioning.

Antemortem, perimortem, and postmortem factors are complex and impact product color

and quality. To aid in production of more consistent and quality meat products, more information

concerning how meat color and protein functionality are affected by these factors should be

created.

59

Appendix A

if nargin == 1

B = double(R(:,:,3));

G = double(R(:,:,2));

R = double(R(:,:,1));

end

if max(max(R)) > 1.0 || max(max(G)) > 1.0 || max(max(B)) > 1.0

R = double(R) / 255;

G = double(G) / 255;

B = double(B) / 255;

end

% Set a threshold

T = 0.008856;

[M, N] = size(R);

s = M * N;

RGB = [reshape(R,1,s); reshape(G,1,s); reshape(B,1,s)];

% RGB to XYZ

MAT = [0.412453 0.357580 0.180423;

0.212671 0.715160 0.072169;

0.019334 0.119193 0.950227];

XYZ = MAT * RGB;

% Normalize for D65 white point

X = XYZ(1,:) / 0.950456;

Y = XYZ(2,:);

Z = XYZ(3,:) / 1.088754;

XT = X > T;

YT = Y > T;

ZT = Z > T;

Y3 = Y.^(1/3);

fX = XT .* X.^(1/3) + (~XT) .* (7.787 .* X + 16/116);

fY = YT .* Y3 + (~YT) .* (7.787 .* Y + 16/116);

fZ = ZT .* Z.^(1/3) + (~ZT) .* (7.787 .* Z + 16/116);

L = reshape(YT .* (116 * Y3 - 16.0) + (~YT) .* (903.3 * Y), M, N);

a = reshape(500 * (fX - fY), M, N);

b = reshape(200 * (fY - fZ), M, N);

if nargout < 2

L = cat(3,L,a,b);

end

Figure 1. Software script converting red, green, blue (RGB) color values to CIELAB color

space.

60

function [xmean, xstd, xmin, xmax, xskew, xquan25, xquan75] = RunStats(x)

xmean = mean(x);

xstd = std(x);

xmin = min(x);

xmax = max(x);

xskew = skewness(x);

xquan25 = quantile(x,0.25);

xquan75 = quantile(x,0.75);

Figure 2. Software script producing statistical parameters for the raw color distribution.

61

clc; clear; close all

[X] = imread('select image.jpg');

[L,a,b] = RGB2Lab(X);

[m,n] = size(L);

Lv =reshape(L,[1,m*n]);

hist(Lv,100);

[Lvmean, Lvstd, Lvmin, Lvmax, Lvskew, Lvquan25, Lvquan75] = RunStats(Lv)

Lv_new=Lv(Lv<79.499 & Lv>60.446);

hist(Lv_new,100)

[Lvmean_trunc, Lvstd_trunc, Lvmin_trunc, Lvmax_trunc, Lvskew_trunc, Lvquan25_trunc,

Lvquan75_trunc] = RunStats(Lv_new)

[ma,na] = size(a);

av =reshape(a,[1,ma*na]);

hist(av,100);

[avmean, avstd, avmin, avmax, avskew, avquan25, avquan75] = RunStats(av)

av_new=av(av<24.6013 & av>5.8017);

hist(av_new,100)

[avmean_trunc, avstd_trunc, avmin_trunc, avmax_trunc, avskew_trunc, avquan25_trunc,

avquan75_trunc] = RunStats(av_new)

[mb,nb] = size(b);

bv =reshape(b,[1,mb*nb]);

hist(bv,100);

[bvmean, bvstd, bvmin, bvmax, bvskew, bvquan25, bvquan75] = RunStats(bv)

bv_new=bv(bv<35.6164 & bv>16.363);

hist(bv_new,100)

[bmean_trunc, bvstd_trunc, bvmin_trunc, bvmax_trunc, bvskew_trunc, bvquan25_trunc,

bvquan75_trunc] = RunStats(bv_new)

Figure 3. Software script truncating the original color distributions to eliminate values

representing background, outliers, and light scattering.

62