Beverage emulsions: Recent developments in formulation, production, and applications

Beverage emulsions: Recent developments in formulation, production, and applications

Accepted Manuscript Beverage Emulsions: Recent Developments in Formulation, Production, and applications Daniel T. Piorkowski, David Julian McClements...

3MB Sizes 1 Downloads 140 Views

Accepted Manuscript Beverage Emulsions: Recent Developments in Formulation, Production, and applications Daniel T. Piorkowski, David Julian McClements PII:

S0268-005X(13)00211-7

DOI:

10.1016/j.foodhyd.2013.07.009

Reference:

FOOHYD 2311

To appear in:

Food Hydrocolloids

Received Date: 13 May 2013 Revised Date:

8 July 2013

Accepted Date: 9 July 2013

Please cite this article as: Piorkowski, D.T., McClements, D.J., Beverage Emulsions: Recent Developments in Formulation, Production, and applications, Food Hydrocolloids (2013), doi: 10.1016/ j.foodhyd.2013.07.009. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Graphical Abstract “BEVERAGE EMULSIONS: RECENT DEVELOPMENTS IN FORMULATION, PRODUCTION, AND APPLICATIONS” by D.T. Piorkowski and D.J. McClements

M AN U

SC

RI PT

Food Hydrocolloids

The article provides an overview of recent research on the formation, stability, and

AC C

EP

TE D

properties of beverage emulsions.

ACCEPTED MANUSCRIPT

1

BEVERAGE EMULSIONS: RECENT DEVELOPMENTS IN

2

FORMULATION, PRODUCTION, AND APPLICATIONS

3

Daniel T. Piorkowski and David Julian McClements1

RI PT

4 5 6

Department of Food Science, University of Massachusetts, Amherst, MA 01003

7

SC

8 9 10

M AN U

11 12 13 14 15

TE D

16 17 18 19

Journal: Food Hydrocolloids

21

Submitted: March 2013

22

Revised: July 2013

24

AC C

23

EP

20

1

Department of Food Science, University of Massachusetts, Amherst, MA 01003, [email protected], 413 545 1019

ACCEPTED MANUSCRIPT 2

25

27

Abstract Soft drinks are one of the most widely consumed and profitable beverages in the world.

RI PT

26

This review article focuses on the utilization of emulsion science and technology for the

29

fabrication of soft drinks by the beverage industry. A brief overview of the various high and low

30

energy methods available for preparing this type of beverage emulsions is given, as well as a

31

discussion of the functional ingredients used to formulate these systems, including oil phases,

32

emulsifiers, weighting agents, ripening inhibitors, and thickening agents. The influence of

33

droplet characteristics on the physicochemical and sensory properties of beverage emulsions is

34

reviewed, with special focus on their influence on product stability. Finally, we discuss recent

35

developments in the soft drinks area, including fortification with vitamins, reduced calorie

36

beverages, and “all-natural” products.

37

EP

TE D

Keywords: beverages; soft drinks; nutraceuticals; flavors; emulsions; nanoemulsions

AC C

38

M AN U

SC

28

ACCEPTED MANUSCRIPT 3

39

41

1. Introduction Globally, soft drinks are one of the most widely consumed and profitable beverages (Table

RI PT

40

1). Cola is the top soft drink flavor currently consumed in the United States, with lemon-lime

43

and orange being the second and third. All three of these soft drink flavors contain hydrophobic

44

citrus compounds extracted from fruit peels. Soft drinks may also contain a variety of other

45

hydrophobic components, such as clouding agents, weighting agents, nutraceuticals, oil-soluble

46

vitamins, and oil-soluble antimicrobials. The non-polar character of flavor oils and other

47

hydrophobic ingredients means that these ingredients cannot simply be dispersed directly into an

48

aqueous phase – they would rapidly coalesce and separate through gravitational forces leading to

49

a layer of oil on top of the product (Given, 2009). Instead they first have to be converted into a

50

colloidal dispersion consisting of flavor molecules encapsulated within small particles suspended

51

within an aqueous medium, e.g., a microemulsion, nanoemulsion, or emulsion (McClements,

52

2011; McClements & Li, 2010). These colloidal delivery systems must be carefully designed to

53

provide desirable physicochemical, sensory, and biological attributes to the final product. A

54

number of desirable attributes of colloidal delivery systems suitable for application in beverage

55

products are highlighted below (McClements, Decker, & Weiss, 2007; McClements & Li, 2010):

M AN U

57 58



62 63 64

Fabrication: Ideally, the delivery systems should be fabricated using robust, reliable and inexpensive manufacturing methods that are easily implemented.



Stability: The delivery systems should be designed to withstand all of the stresses

AC C

61

Composition: Ideally, the delivery systems should be fabricated entirely from “label friendly” food-grade ingredients that are economic and easy to handle.

59 60

TE D



EP

56

SC

42

that a product may experience during its production, storage, transport and utilization, such as temperature fluctuations, exposure to light and oxygen, exposure to mechanical forces (such as stirring, flow through a pipe, and vibrations), variations in aqueous phase composition (such as pH, ionic strength, buffer type,

65

ingredient interactions), and exposure to microorganisms (such as yeasts, molds or

66

bacteria).

ACCEPTED MANUSCRIPT 4 •

67

Physicochemical and sensory properties: The delivery system should not adversely

68

affect the optical properties, rheology, or flavor profile (aroma, taste, and mouthfeel)

69

of the beverage product into which it is incorporated. Biological activity: The delivery system should not adversely affect the biological

RI PT



70 71

activity of any encapsulated bioactive components, such as antimicrobials, vitamins,

72

or nutraceuticals.

73

This review article provides an overview of the current status of the design, formulation, and production of emulsion-based delivery systems suitable for utilization within the beverage

75

industry.

76

2. Emulsion Science and Technology in the Beverage Industry

M AN U

77

SC

74

Hydrophobic components (such as flavor oils, clouding agents, oil-soluble vitamins, and nutraceuticals) can be incorporated into a variety of different colloidal delivery systems suitable

79

for application within beverage products (McClements, 2012; McClements & Rao, 2011), with

80

the most common being microemulsions, nanoemulsions, and emulsions (Figure 1). Each of

81

these colloidal dispersions has particular benefits and limitations for the encapsulation of

82

hydrophobic compounds. Microemulsions are thermodynamically stable systems under a

83

specific set of environmental conditions (e.g., composition and temperature), and are therefore

84

easy to fabricate (often by simple mixing) and tend to have good long-term stability.

85

Microemulsions typically contain very small particles (r < 25 nm) and therefore tend to be

86

optically transparent, which is desirable for soft drinks that should be clear. On the other hand,

87

the formation of microemulsions usually requires relatively high levels of synthetic surfactants

88

and sometimes the use of cosurfactants/cosolvents, which can be undesirable for cost, taste, and

89

labeling reasons. Microemulsions may also become thermodynamically unstable if

90

environmental conditions are altered (such as temperature or composition).

EP

AC C

91

TE D

78

Conventional emulsions (r > 100 nm) and nanoemulsions (r < 100 nm) are both

92

thermodynamically unstable systems, and therefore tend to breakdown during storage through a

93

variety of instability mechanisms (Figure 2), such as gravitational separation, flocculation,

94

coalescence and Ostwald ripening (McClements, et al., 2007; McClements & Rao, 2011).

95

Emulsion systems must therefore be carefully designed to inhibit these instability mechanisms

ACCEPTED MANUSCRIPT 5

and provide sufficient kinetic stability throughout the lifetime of the product. Emulsions usually

97

contain larger droplets than microemulsions and therefore they scatter light more strongly and

98

appear more turbid or cloudy. This is an advantage for soft drinks that are required to have a

99

cloudy appearance, but a disadvantage for products where optical clarity is required.

RI PT

96

Nevertheless, recently it has been shown that emulsions with ultrafine droplets, often referred to

101

as “nanoemulsions”, can be prepared that are optically transparent (McClements, 2012;

102

McClements & Rao, 2011). A major advantage of emulsions and nanoemulsions is that the

103

emulsifier-to-oil ratio required to formulate them is often much less than that required for

104

microemulsions, and they can be formulated from all natural ingredients (such as proteins and

105

polysaccharides) rather than synthetic surfactants (such as Tweens). In this article, we focus

106

primarily on the utilization of emulsion systems (conventional emulsions and nanoemulsions) in

107

the preparation of soft drinks but much of the material is also relevant to the formulation of

108

microemulsions.

M AN U

109

SC

100

It should be noted that the emulsions used in the beverage industry are typically divided into two groups: flavor emulsions and cloud emulsions. Flavor emulsions contain lipophilic

111

compounds that are primarily present to provide taste and aroma to a beverage product (such as

112

lemon, lime, or orange oils). On the other hand, cloud emulsions are used to provide specific

113

optical properties to certain beverage products, i.e., to increase their turbidity (“cloudiness”).

114

Cloud emulsions are typically prepared using an oil phase that is highly water-insoluble and that

115

is not prone to chemical degradation, such as flavorless vegetable oils. In addition, the size of the

116

droplets within cloud emulsions is designed so that they have dimensions where strong light

117

scattering occurs, but are not too large to undergo gravitational separation (e.g., r = 100-200

118

nm). Cloud emulsions are often added to beverages that only contain a relatively low percentage

119

of juice and provide a desirable cloudy appearance that hides sedimentation and ringing.

AC C

EP

TE D

110

120

In this article, we will use the term “emulsion” to refer to both nanoemulsions and

121

conventional emulsions because they have similar structures and properties. Generally, an

122

emulsion consists of at least two immiscible liquids (usually oil and water), with one of the

123

liquids being dispersed as small spherical droplets in the other (Dickinson, 1992a; Dickinson &

124

Stainsby, 1982; Friberg, Larsson, & Sjoblom, 2004; McClements, 2005b). In general, emulsions

125

are classified according to the relative spatial organization of the oil and water phases. A system

126

that contains oil droplets dispersed within water is called an oil-in-water (O/W) emulsion,

ACCEPTED MANUSCRIPT 6

whereas a system that contains water droplets dispersed in oil is called a water-in-oil (W/O)

128

emulsion. It is possible to prepare more complex emulsion structures, e.g., oil-in-water-in-oil

129

(O/W/O), water-in-oil-in-water (W/O/W) or oil-in-water-in-water (O/W/W) emulsions

130

(Benichou, Aserin, & Garti, 2004; Garti & Bisperink, 1998; van der Graaf, Schroen, & Boom,

131

2005). Currently, almost all of the emulsions used in the beverage industry are of the O/W type,

132

although there may be certain advantages to using other emulsion types for certain applications.

133

For example, in principle it is possible to trap a hydrophilic bioactive component within the inner

134

water phase of a W/O/W emulsion to protect it from chemical degradation or for taste masking.

135

In practice, it is often difficult to formulate W/O/W emulsions that have sufficient stability for

136

commercial applications, although this is still an active area of research.

SC

RI PT

127

Emulsions are thermodynamically unfavorable systems that tend to break down over time

138

though a variety of physicochemical mechanisms, including gravitational separation (creaming

139

and sedimentation), droplet aggregation (flocculation and coalescence) and droplet growth

140

(Ostwald ripening) (Dickinson, 1992a; Friberg, et al., 2004; McClements, 2005b). It is possible

141

to form emulsions that are kinetically stable for a reasonable period of time by including

142

substances known as stabilizers, e.g., emulsifiers, weighting agents, ripening inhibitors, or

143

texture modifiers. It is important to clearly distinguish the different physicochemical

144

mechanisms involved in promoting emulsion stability for these different categories of stabilizers.

145

Emulsifiers are surface-active molecules that adsorb to the surface of freshly formed droplets

146

during homogenization, forming a protective layer that prevents the droplets from aggregating.

147

Weighting agents are dense hydrophobic components added to low-density oils to prevent

148

gravitational separation. Ripening inhibitors are water-insoluble components added to polar oils

149

to prevent Ostwald ripening. Texture modifiers are substances used to increase the viscosity or

150

gel aqueous solutions, thereby retarding or preventing droplet movement. A more detailed

151

description of different types of stabilizers that can be used in beverage emulsions is given in a

152

later section. Selecting the most appropriate stabilizer(s) for a particular application is one of the

153

most important factors determining the shelf-life and physicochemical properties of beverage

154

emulsions.

AC C

EP

TE D

M AN U

137

ACCEPTED MANUSCRIPT 7

155

3. Controlling Droplet Characteristics The bulk physicochemical properties of beverage emulsions (such as optical properties,

157

stability, rheology, molecular partitioning, and release characteristics) are largely determined by

158

the properties of the droplets they contain (McClements, 2005b), such as composition,

159

concentration, size, and charge (Figure 3). In this section, we discuss some of the most

160

important droplet characteristics that can be controlled by beverage manufacturers in order to

161

create products with specific desirable functional properties.

162

3.1. Droplet composition

SC

163

RI PT

156

The composition of the oil phase has a major influence on the formation and stability of beverage emulsions, which has often been overlooked in academic research. Beverage

165

emulsions may contain a variety of different hydrophobic components, including flavor oils,

166

essential oils, triacylglycerol oils, oil-soluble vitamins, nutraceuticals, weighting agents, and

167

ripening inhibitors. These components vary in their molecular characteristics (such as molecular

168

weight, molecular conformation, and functional groups), which leads to changes in their

169

physicochemical properties (such as polarity, water-solubility, density, viscosity, refractive

170

index, physical state, and melting point). Many of these molecular and physicochemical

171

properties have a major influence on the formation, stability, and functionality of emulsions. For

172

example, oil viscosity influences the efficiency of droplet disruption during high energy

173

homogenization – the closer the ratio of dispersed phase viscosity to continuous phase viscosity

174

(ηD/ ηC) is to unity, the more efficient is droplet disruption and the smaller is the particle size

175

produced (Walstra, 1993, 2003). Oil density determines the rate of particle creaming or

176

sedimentation within emulsions – the greater the density contrast between the droplets and

177

surrounding fluid, the faster the rate of gravitational separation (McClements, 2005c). Oil

178

refractive index determines the efficiency of light scattering by droplets in emulsions – the

179

greater the refractive index contrast between the droplets and surrounding fluid, the stronger the

180

degree of light scattering and the more turbid the appearance (Chanamai & McClements, 2002b).

181

The water-solubility of an oil phase determines the stability of an emulsion to Ostwald ripening

182

due to diffusion of oil molecules through the aqueous phase (Kabalnov, 2001; McClements,

183

Henson, Popplewell, Decker, & Choi, 2012). Oil interfacial tension plays a number of important

184

roles in determining emulsion formation and stability. First, the ease of droplet disruption during

AC C

EP

TE D

M AN U

164

ACCEPTED MANUSCRIPT 8

high energy homogenization decreases as the interfacial tension decreases (Walstra, 1993).

186

Second, the rate of droplet coalescence increases as the interfacial tension decreases (Kabalnov

187

& Wennerstrom, 1996). Third, the ability of emulsifiers to adhere to droplet surfaces decreases

188

as the bare oil-water interfacial tension decreases (Chanamai, Horn, & McClements, 2002).

189

Finally, the rate of droplet growth due to Ostwald ripening depends on the interfacial tension at

190

the oil-water interface (Kabalnov, 2001).

RI PT

185

For flavor emulsions, it is important to control the type and concentration of the flavor

192

molecules initially present in the oil phase. It is also important to be aware that the location of

193

the flavor molecules within an emulsion is governed by their oil-water partition, which depends

194

on carrier oil type (Choi, Decker, Henson, Popplewell, & McClements, 2009; Choi, Decker,

195

Henson, Popplewell, & McClements, 2010b). The flavor profile of an emulsion may therefore

196

change if the carrier oil type is altered, if the physical state of the carrier oil changes, or if an

197

emulsion is diluted, since this will change the distribution of the flavor molecules in the oil,

198

water and air (Choi, et al., 2009; Choi, et al., 2010b; Mei, et al., 2010).

M AN U

SC

191

It is important for beverage manufacturers to understand the composition of the oil phases

200

used to formulate commercial products, and to understand how specific lipophilic components

201

influence the formation, stability, and properties of final products.

202

3.2. Droplet concentration

203

TE D

199

In general, the concentration of droplets in an emulsion influences its texture, stability, appearance, sensory attributes, and nutritional quality (McClements, 2005b; McClements & Rao,

205

2011). Droplet concentration is usually characterized in terms of the dispersed phase volume

206

fraction (φ), which is the volume of emulsion droplets (VD) divided by the total volume of

207

emulsion (VE): φ = VD/VE. Practically, it is often more convenient to express the droplet

208

concentration in terms of the dispersed phase mass fraction (φm), which is the mass of emulsion

209

droplets (mD) divided by the total mass of emulsion (mE): φm = mD/mE. When the densities of the

210

two phases are equal, the mass fraction is equivalent to the volume fraction. It is particularly

211

important to convert the droplet concentration to the appropriate units when comparing

212

experimental work with theoretical predictions.

AC C

EP

204

213

In beverage emulsions, controlling the droplet concentration is important for a number of

214

reasons. Beverage emulsions are often prepared in a concentrated form (> 10% oil) because this

ACCEPTED MANUSCRIPT 9

facilitates handling and transport, but they are highly diluted when they are introduced into the

216

final product (< 0.1% oil). The amount to which an emulsion concentrate is diluted influences

217

the appearance of a final product, since emulsion turbidity or cloudiness increases with oil

218

droplet concentration. Dilution also influences the total amount of flavor molecules present in

219

final products, as well as their partitioning between the oil and water phases (Choi, et al., 2009).

220

In the concentrate, droplet concentration has a major impact on the rheological properties of the

221

system. From a practical point of view, it may be important to have a high oil loading in the

222

concentrate emulsion so as to reduce transport and storage costs, but not have the oil content so

223

high that the product is unstable or cannot easily be dispersed into the final product.

224

3.3. Droplet size distribution

SC

The size distribution of the droplets in a beverage emulsion has a strong impact on its

M AN U

225

RI PT

215

physical stability (e.g., to gravitational separation, flocculation, coalescence and Ostwald

227

ripening) and its optical properties (e.g., lightness and color) (McClements, 2005b). Beverage

228

manufacturers must therefore specify the optimum droplet size distribution required for their

229

particular product based on the properties required, e.g., optical clarity and shelf-life. They must

230

then develop a formulation and manufacturing process that can reliably produce a beverage with

231

this droplet size distribution. Immediately after the product has been manufactured it is usually

232

important to measure the droplet size distribution to ensure that it has met the specified quality

233

criteria, e.g., using light scattering instruments. It may also be important to measure changes in

234

the droplet size distribution of the product during storage or after an accelerated storage test to

235

predict its long-term stability (McClements, 2007).

EP

236

TE D

226

The particle size distribution (PSD) of an emulsion specifies the concentration of droplets within different size classes, and can be conveniently measured using various commercially

238

available instruments (McClements, 2005b). When presenting or interpreting PSD data on a

239

beverage emulsion it is important to pay particular attention to the manner in which the particle

240

concentration and particle size are presented. The concentration of particles within a particular

241

size class is usually presented as either the volume or number percent, whereas the size of the

242

particles in a particular size class is usually presented as either the mid-point particle radius or

243

diameter. The same PSD may look very different if it is plotted as volume versus particle size or

244

as number versus particle size. Typically, particle volume versus particle diameter is the most

AC C

237

ACCEPTED MANUSCRIPT 10

widespread and informative way of presenting particle size data. Commercial beverage

246

emulsions are always polydisperse systems that can be characterized as being "monomodal",

247

"bimodal" or "multimodal" depending on whether there are one, two, or more peaks in the

248

particle size distribution. Typically, beverage manufacturers would like to produce a final

249

product that has a narrow monomodal distribution, as this usually provides the best long-term

250

stability.

251

RI PT

245

In many practical situations it is important to have knowledge of the full PSD of a beverage emulsion since this contains information about the size characteristics of all of the particles

253

present, as well as providing insights into the possible origin and nature of any instability

254

mechanisms. For example, it may be possible to detect a small population of large particles that

255

may cause problems with creaming during long-term storage (i.e., ringing). In addition, by

256

measuring changes in the PSD overtime it is sometimes possible to distinguish between different

257

instability mechanisms (e.g., coalescence versus Ostwald ripening). Nevertheless, in some

258

situations it is more convenient to represent the full particle size distribution by a measure of its

259

central tendency and spread. The mean, median, or modal particle sizes are often used as

260

measures of the central tendency, whereas the relative standard deviation is often used as a

261

measure of spread (Walstra, 2003). The mean particle size is the most widely used method of

262

representing the central tendency of emulsion particle size distributions in the beverage industry.

263

It is important to realize that a number of different mean particles sizes can be derived from

TE D

M AN U

SC

252

a full PSD and each mean size can have a different magnitude and physical meaning

265

(McClements, 2007). The three most commonly used mean particle sizes are the number-

266

weighted mean diameter (dN or d10 = Σnidi/ Σni), the surface-weighted mean diameter (dS or d32 =

267

Σnidi3/ Σnidi2) and the volume-weighted mean diameter (dV or d43 = Σnidi4/ Σnidi3). Generally, the

268

volume-weighted mean diameter is more sensitive to the presence of large particles than the

269

number-weighted mean diameter, and so it often provides the most rigorous test of the physical

270

stability of a beverage emulsion, i.e., if d43 is small then the emulsion is more likely to remain

271

stable. Appreciable differences between the values of d10, d32 and d43 generally indicate that the

272

particle size distribution is broad or multimodal. One must therefore be very careful when

273

interpreting or reporting particle size data to identify which mean particle size value (and

274

corresponding relative standard deviation) is being used. It should also be noted that mean

AC C

EP

264

ACCEPTED MANUSCRIPT 11

275

values should be treated with caution when used to represent highly polydisperse emulsions (e.g.,

276

aggregated systems), and it is always useful to examine the full particle size distribution.

277

Commercial beverage manufacturers usually develop a set of standardized particle size criteria that they use to determine whether a particular batch of product has the desired

279

physicochemical characteristics, e.g., long-term stability and optical properties. For example, a

280

manufacturer might specify that that mean droplet diameter (d43) of a particular class of products

281

should be < 500 nm, and that > 90% of the droplets should be smaller than 800 nm. The precise

282

criteria used will depend on the product being manufactured (especially whether it should be

283

clear or opaque).

284

3.4. Droplet charge

SC

The droplets in most beverage emulsions have an electrical charge because of adsorption of

M AN U

285

RI PT

278

ionic species to their surfaces, e.g., proteins, ionic polysaccharides, ionic surfactants,

287

phospholipids, fatty acids, and some small ions (McClements, 2005b). The electrical

288

characteristics of a droplet surface depend on the type, concentration and organization of the

289

ionized species present, as well as the ionic composition and physical properties of the

290

surrounding aqueous phase. The electrical charge on the oil droplets in a beverage may be

291

important for a number of reasons: it determines the stability of the droplets to aggregation due

292

to its influence of the magnitude, range and sign of electrostatic interactions; it determines the

293

interactions of droplets with other charged species in an emulsion e.g., ions (such as calcium or

294

iron), or polyelectrolytes (such as proteins or polysaccharides); it influences how the droplets

295

interact with electrically charged surfaces, such as storage vessels, bottles, cups, and the mouth;

296

it influences the behavior of the droplets in an electrical field, which is important for measuring

297

their charge using electrophoresis.

EP

AC C

298

TE D

286

The electrical characteristics of a droplet in an emulsion are usually characterized in terms

299

of its surface charge density (σ), electrical potential (Ψ0), and/or ζ−potential (ζ) (Hunter, 1986).

300

The surface charge density is the amount of electrical charge per unit surface area, which

301

depends on the net number of ionized groups per unit interfacial area. The electrical potential is

302

the amount of energy required to increase the surface charge density from zero to σ. The

303

electrical potential depends on the surface charge density, but also on the ionic composition of

304

the surrounding medium due to electrostatic screening effects. At a fixed surface charge density,

ACCEPTED MANUSCRIPT 12

the electrical potential decreases with increasing ionic strength due to these effects. The zeta-

306

potential (ζ) is the electrical potential at the "shear plane", which is defined as the distance away

307

from the droplet surface below which the counter-ions remain strongly attached to the droplet

308

when it moves in an electrical field. Practically, the ζ-potential is a better representation of the

309

electrical characteristics of an oil droplet because it inherently accounts for the adsorption of any

310

counter ions or ionic species to the droplet surface. In addition, the ζ-potential is more

311

convenient to measure than the surface charge density or electrical potential (Hunter, 1986).

312

Typically, the electrical characteristics of the droplets in an emulsion are determined by

313

measuring the ζ-potential versus pH under appropriate measurement conditions (such as ionic

314

composition).

SC

Droplet aggregation is inhibited in many beverage emulsions by using ionic emulsifiers that

M AN U

315

RI PT

305

adsorb to the droplet surfaces and prevent them from coming close together because of

317

electrostatic repulsion (Dickinson, 1992b; Friberg, et al., 2004; McClements, 2005b).

318

Electrostatic repulsion plays a major role in determining the aggregation stability of fat droplets

319

coated by charged emulsifiers that only form thin layers that generate short range steric

320

repulsion, such as globular proteins and ionic surfactants. On the other hand, electrostatic

321

repulsion is less important in systems where the fat droplets are coated by emulsifiers that form

322

thick interfacial layers that generate long range steric repulsion, such as polysaccharides (gum

323

arabic and modified starch). For electrostatically-stabilized emulsions, the magnitude of the ζ-

324

potential should be greater than about 20 mV to produce systems that are stable during long-term

325

storage. For sterically-stabilized emulsions, the droplet charge may not be important in terms of

326

their physical stability, but it may still be important in systems where chemical reactions occur

327

within the oil droplets that are induced by water-soluble ionic species, such as oxidation of ω-3

328

fatty acids by transition metals.

329

3.5. Interfacial properties

EP

AC C

330

TE D

316

The boundary between the oil and water phases in an emulsion consists of a narrow region

331

(≈ 1 to 50 nm thick) that surrounds each oil droplet, and contains a mixture of oil, water, and

332

emulsifier molecules, as well as possibly other molecular species, such as mineral ions,

333

polyelectrolytes, and polar lipids. The interfacial region makes up a significant fraction of the

334

volume of a droplet when the droplet diameter is less than about 1 µm (McClements & Rao,

ACCEPTED MANUSCRIPT 13

2011), and is therefore particularly important in beverage emulsions since they usually contain

336

droplets considerably smaller than this size . The interfacial region can influence many

337

important physicochemical and sensory properties of beverages emulsions, including their

338

stability, rheology, mouthfeel, and flavor. For this reason, it is often important to have

339

knowledge about the interfacial properties of the droplets in a beverage emulsion, and to

340

establish the major factors that influence them. Some of the most important properties of the

341

interfacial region are: composition; structural organization; thickness; rheology; interfacial

342

tension; and charge. These properties are determined by the type, concentration and interactions

343

of any surface-active species present, as well as by the events that occur before, during, and after

344

emulsion formation, e.g., complexation, competitive adsorption, layer-by-layer formation

345

(Dickinson, 2003). As mentioned earlier, the electrical charge on the droplet interface influences

346

its interaction with other charged molecules, as well as its stability to aggregation. The thickness

347

and rheology of the interfacial region influences the stability of emulsions to gravitational

348

separation, coalescence and flocculation, and determines the rate at which molecules leave or

349

enter the droplets (Dickinson, 2003; McClements, 2005b). For example, the ability of interfacial

350

coatings to prevent droplet flocculation is strongly influenced by their thickness.

M AN U

SC

RI PT

335

Beverage manufacturers should therefore be aware of the nature of the interfacial region

352

surrounding the oil droplets in their products, and the fact that they may be able to manipulate its

353

properties to improve product performance.

354

3.6. Colloidal interactions

EP

355

TE D

351

The attractive and repulsive colloidal interactions that operate between the oil droplets in beverage emulsions determine their stability to flocculation and coalescence, which in turn

357

influences their creaming stability and rheology (Friberg, et al., 2004; McClements, 2005b). The

358

colloidal interactions between two oil droplets can be described in terms of an interaction

359

potential (w(h)), which is the energy required to bring two droplets from an infinite distance

360

apart to a surface-to-surface separation of h (Figure 4). The overall interaction potential is made

361

up from contributions from various types of interactions, with the most important being van der

362

Waals, steric, electrostatic, depletion, and hydrophobic interactions (Israelachvili, 2011;

363

McClements, 2005b). These individual interactions vary in their sign (attractive or repulsive),

364

magnitude (weak to strong) and range (short to long). Each of the individual interactions usually

AC C

356

ACCEPTED MANUSCRIPT 14

has a simple dependence on surface-to-surface separation, but the sum of the interactions can

366

exhibit a more complex dependence. For example, the interaction potential between two oil

367

droplets coated by a layer of charged polymer molecules would have a number of maximum and

368

minimum values at certain separations, such as short- and long-range energy barriers, and

369

primary and secondary minima (Figure 4). Generally, droplets tend to aggregate when attractive

370

interactions dominate, but remain as individual entities when repulsive interactions dominate

371

(McClements, 2005b).

372

RI PT

365

It is particularly important for scientists working in the beverage industry to identify and understand the major colloidal interactions operating between the droplets in their particular

374

product. This knowledge can then be used to establish the optimum approach for maintaining

375

product stability during production, transport and storage. For example, if a beverage emulsion

376

is stabilized by a protein-based emulsifier, then electrostatic repulsive interactions will play an

377

important role in preventing droplet aggregation. In this situation, the system will be sensitive to

378

environmental changes that reduce the magnitude and range of the electrostatic repulsion acting

379

between droplets, such as altering the pH or adding salts (particularly multivalent counter-ions).

380

On the other hand, if the beverage emulsion is stabilized by a polysaccharide-based emulsifier,

381

then steric repulsive interactions will be most important for preventing droplet aggregation. In

382

this case, the product will be much less sensitive to droplet aggregation when the pH or ionic

383

strength is changed. In this latter case, emulsion stability depends on the thickness and

384

hydrophilicity of the interfacial layer, which will depend on the molecular characteristics of the

385

polysaccharide molecules. A summary of the major colloidal interactions in beverage emulsions

386

is given in Table 2.

387

4. Physicochemical Properties

M AN U

TE D

EP

AC C

388

SC

373

The physicochemical properties of beverage emulsions play an important role during the

389

manufacturing process, as well as in determining the perceived quality attributes of the final

390

product. The most important physicochemical attributes of these systems are briefly discussed in

391

this section. Some of these properties are more important in the concentrated form of the

392

beverage, whereas others are more important in the diluted form.

ACCEPTED MANUSCRIPT 15

393 394

4.1. Optical properties The first cue that a consumer uses to judge the quality or desirability of a finished beverage product is its visual appearance (provided it is packaged or poured into a transparent container,

396

such as a bottle or cup). Each type of beverage product is expected to have a particular

397

appearance depending on its nature, e.g., a dark brown cola, a cloudy orange juice, or a clear

398

green lime juice. From a scientific viewpoint, emulsion appearance is categorized in terms of

399

their opacity and color, which can be quantitatively described using tristimulus color coordinates,

400

such as the L*a*b* system (McClements, 2005b). In this color system, L* represents the

401

lightness, and a* and b* are color coordinates: where +a* is the red direction, -a* is the green

402

direction; +b* is the yellow direction, -b* is the blue direction; low L* is dark and high L* is

403

light. The opacity of an emulsion can therefore by characterized by the lightness (L*), while the

404

color intensity can be characterized by the chroma: C = (a*2 + b*2)1/2. The color intensity is

405

usually inversely related to the lightness, so that the chroma decreases (fades) when the lightness

406

increases. The optical properties of emulsions are mainly determined by the relative refractive

407

index, concentration, and size distribution of the droplets they contain (Chanamai &

408

McClements, 2002b; Danviriyakul, McClements, Decker, Nawar, & Chinachoti, 2002;

409

McClements, 2005b). The lightness of an emulsion tends to increase with increasing refractive

410

index contrast and increasing droplet concentration, and has a maximum value at a particular

411

droplet size. This has important implications for the development of beverage products that

412

should be either clear or opaque. In general, the lightness of emulsions increases steeply as the

413

oil droplet concentration increases from about 0 to 5 wt%, but then increases more gradually at

414

higher droplet concentrations (Figure 5).

SC

M AN U

TE D

EP

As mentioned earlier, some beverages are expected to have optical clarity, whereas others

AC C

415

RI PT

395

416

are expected to be cloudy. Optimizing the initial particle size distribution of a beverage

417

emulsion, as well as inhibiting any changes in the particle size during storage, is therefore a

418

particularly important part of designing a commercial product with the desired optical properties.

419

For clear products, the majority of droplets should be less than about 50 nm in diameter so that

420

the light scattering is very weak (Wooster, Golding, & Sanguansri, 2008a). The scattering

421

efficiency of the individual oil droplets determines the maximum amount of oil phase that can be

422

incorporated into a clear beverage before it becomes noticeably cloudy. As a rule of thumb, a

423

turbidity of 0.05 cm-1 (at 600 nm) can be considered to be a rough cut-off point between

ACCEPTED MANUSCRIPT 16

transparent and cloudy products. For cloudy products, the majority of droplets should be

425

between about 200 and 400 nm in diameter so that the light scattering is very strong

426

(McClements, 2002). In this case, the scattering efficiency of the individual oil droplets will

427

determine the minimum amount of a clouding emulsion required to reach a particular turbidity in

428

the final product.

429

4.2. Rheology

430

RI PT

424

The rheological properties of beverage emulsions are also an important factor determining their manufacture and utilization. Most beverage emulsions are initially manufactured in a

432

concentrated form, which is diluted appreciably during the production of the final beverage

433

product. The droplet concentration in the beverage concentrate typically ranges from 3 to 30%,

434

while that in the final product is typically < 0.1%. Industrially, the rheology of the beverage

435

concentrate is important since it influences the ease of mixing, flow through a pipe, and

436

packaging. A manufacturer typically wants to have as high an oil loading as possible, without

437

the product becoming too viscous or gel-like to handle easily. This requires careful control of

438

the total droplet concentration in the system. The droplet concentration in the final beverage

439

concentration is usually so low that the rheology is dominated by the properties of the aqueous

440

continuous phase (see discussion below).

TE D

M AN U

SC

431

The rheology of dilute colloidal dispersions is normally characterized by the shear viscosity

442

(Genovese, Lozano, & Rao, 2007; McClements, 2005b). When the droplet concentration is less

443

than about 5% (φ < 0.05), the shear viscosity can be described by Einstein’s equation:

444

446 447

η = η 0 (1 + 2.5φ )

(1)

AC C

445

EP

441

Here, η is the viscosity of the overall system, η0 is the viscosity of the continuous phase, and

448

φ is the disperse phase volume fraction. This equation under-predicts the viscosity of colloidal

449

dispersions with higher droplet concentrations due to the effects of droplet-droplet interactions.

450

Einstein’s equation was derived assuming that the colloidal particles were rigid isolated spheres

451

surrounded by a Newtonian fluid. Nevertheless, it still provides a good approximation of the

452

rheological properties of dilute beverage emulsions because the flow of the liquid within the oil

453

phase is inhibited by the emulsifier coating. The above equation shows that the viscosity of a

ACCEPTED MANUSCRIPT 17

dilute emulsion increases linearly with droplet concentration, but that the most important factor

455

affecting the overall rheology is the viscosity of the continuous phase. Thus the most effective

456

means of controlling the viscosity of a dilute beverage emulsion is to change the viscosity of the

457

continuous phase, e.g., by adding sugars or polymer thickening agents.

458

RI PT

454

The viscosity of concentrated emulsions can be described by a semi-empirical equation that

459

takes into account droplet-droplet interactions (Berli, Deiber, & Quemada, 2005; McClements,

460

2005b; Quemada & Berli, 2002): −2

461

 φ η = η 0 1 −  φc

462

Here, φ is the disperse phase volume fraction, and φc (≈ 0.63) is a critical disperse phase

(2)

SC

  

volume fraction above which the droplets are so closely packed together that they cannot easily

464

flow past each other. This equation shows that the viscosity of an emulsion increases with

465

increasing droplet concentration, gradually initially and then steeply as the droplets become more

466

closely packed (Figure 5). Around and above the droplet concentration where close packing

467

occurs, the emulsion becomes highly viscosity and may exhibit solid-like characteristics, such as

468

visco-elasticity and plasticity (Berli, et al., 2005; McClements, 2005b; Quemada & Berli, 2002).

469

In flocculated systems the critical concentration where the system becomes highly viscous or

470

solid-like may be much lower than in a non-flocculated system. It is therefore important for

471

beverage manufactures to consider the influence of droplet concentration and interactions on the

472

rheological properties of emulsion concentrates (Genovese, et al., 2007; McClements, 2005b;

473

Walstra, 2003).

474

4.3. Molecular distribution and release characteristics

TE D

EP

AC C

475

M AN U

463

A beverage emulsion may contain a number of constituents that partition into different

476

phases within the product, e.g., oil, aqueous, interfacial, or gas phases (McClements, 2005b).

477

The physical location of some of these constituents may have a major impact on the quality

478

attributes of the final product. For example, many flavor oils contain constituents that are

479

chemically unstable and therefore prone to degradation during storage (e.g., citral). The

480

chemical stability of these constituents is often dependent on their molecular environment. For

481

example, the rate of acid-catalyzed degradation of citral has been shown to occur considerably

482

faster when it is located within an aqueous phase than when it is present within an oil phase

ACCEPTED MANUSCRIPT 18

(Choi, et al., 2009; Choi, et al., 2010b). This would suggest that it is better to keep citral under

484

neutral conditions or within the emulsion concentrate as long as possible before the final dilution

485

into the acid phase is carried out. The perceived flavor profile of beverage emulsions depends on

486

the distribution of volatile molecules between the liquid and gas phases. Increasing the oil

487

content of an emulsion decreases the concentration of hydrophobic (KOW > 1) volatiles in the

488

headspace and therefore reduces the perceived flavor profile (Figure 6). This phenomenon is

489

important to take into account when reformulating a beverage product so that it contains a

490

different fat concentration, e.g., fortification with a bioactive lipid such as ω-3 oils.

RI PT

483

The location of a constituent within a beverage emulsion is governed by its equilibrium

492

partition coefficients (e.g., oil-water, oil-air, oil-interface) and its mass transport kinetics through

493

the system (McClements, 2005b). When a beverage emulsion is placed in the mouth there is a

494

redistribution of flavor molecules, with some of the aroma compounds leaving the product and

495

entering the nasal cavity. The rate at which flavor molecules leave the droplets in beverage

496

emulsions is usually extremely quick (< 0.1 s for KOW < 1000), and therefore droplet dimensions

497

tend to have little impact on the flavor release profile (McClements, 2005b). Nevertheless, it

498

may be possible to encapsulate oil droplets within hydrogel matrices to slow down the release of

499

flavor molecules within the mouth.

500

5. Beverage Emulsion Shelf-Life

TE D

M AN U

SC

491

One of the most important factors determining the commercial viability of beverage

502

emulsions is their ability to resist changes in their physical and chemical properties after their

503

production. Beverage emulsions experience a range of environmental stresses during their

504

manufacture, transport, storage, and utilization that may reduce their shelf lives: mechanical

505

forces (e.g., stirring, flow through a pipe, centrifugation, vibrations, and pouring); temperature

506

variations (e.g., freezing, chilling, warming, pasteurization, and sterilization); exposure to light

507

(e.g., natural or artificial visible or ultraviolet waves); exposure to oxygen; variations in solution

508

properties (e.g., pH and mineral composition of water). Exposure to these environmental

509

stresses may promote emulsion instability through a variety of physicochemical mechanisms:

510

loss of ingredient functionality (e.g., changes in solubility, surface activity, or stabilization

511

capacity); acceleration of chemical degradation reactions (e.g., oxidation, polymerization, or

AC C

EP

501

ACCEPTED MANUSCRIPT 19

hydrolysis); acceleration of physical instability mechanisms, (e.g., flocculation, coalescence or

513

Ostwald ripening). In this section, a brief overview of some of the major instability mechanisms

514

in beverage emulsions is given, and some suggestions for preventing them from occurring are

515

provided.

516

5.1. Physical Stability

517

RI PT

512

Emulsions are thermodynamically unfavorable systems that tend to break down over time due to a variety of physicochemical mechanisms (Figure 2), including gravitational separation,

519

flocculation, coalescence and Ostwald ripening (Dickinson, 1992a; Friberg, et al., 2004;

520

McClements, 2005b). All of these instability mechanisms lead to a change in the structural

521

organization of the various components within the system, rather than in the type of molecules

522

present. Nevertheless, changes in the chemical structure of active components can lead to

523

changes in physical stability, and vice versa.

524

5.1.1. Gravitational Separation

M AN U

525

SC

518

Gravitational separation is one of the most common forms of physical instability in commercial beverage emulsions, and it may take the form of either creaming or sedimentation

527

depending on the relative densities of the oil droplets and the surrounding aqueous phase.

528

Creaming is the upward movement of droplets when they have a lower density than the aqueous

529

phase, whereas sedimentation is the downwards movement of droplets when they have a higher

530

density than the aqueous phase. The oil phases used in beverage emulsions consist primarily of

531

triacylglycerol and/or flavor oils, which have lower densities than water and so creaming is more

532

prevalent (Table 3). However, if a beverage emulsion contained an excess of weighting agent

533

within the oil phase then it may be prone to sedimentation. A beverage emulsion is also prone to

534

sedimentation if it contains very small oil droplets covered by relatively thick and dense

535

interfacial layers (see below) (McClements, 2011).

AC C

EP

TE D

526

536

One of the most common problems reported in beverage emulsions is “ringing”, which is the

537

accumulation of a visible ring of oil droplets on the top of a product. This effect is due to droplet

538

creaming, which may have occurred because a population of droplets in the initial emulsion was

539

too large, or because some droplet growth occurred during storage, e.g., due to flocculation,

540

coalescence, or Ostwald ripening (see later sections). To a first approximation, the velocity that

541

an oil droplet moves upwards in a dilute emulsion due to gravity is given by Stokes’ law:

ACCEPTED MANUSCRIPT 20 2 2 grparticle ( ρ particle − ρ 0 )

542

v=−

543

where, v is the creaming velocity, rparticle is the particle radius, ρparticle is the particle

544

density, ρ0 is the aqueous phase density, η0 is the aqueous phase viscosity, and g is the

545

acceleration due to gravity. This equation shows that the rate of droplet creaming should

546

decrease as the droplet size decreases, the density contrast decreases, or the aqueous phase

547

viscosity increases. Gravitational forces cause droplets to move either upwards or downwards

548

depending on their density relative to the surrounding aqueous phase. Hence, if only

549

gravitational forces operated, then the droplets would accumulate at either the top or the bottom

550

of an emulsion. In practice, droplets may also move because of Brownian motion associated

551

with the thermal energy of the system. Brownian motion favors the random distribution of the

552

droplets throughout the entire volume of the emulsion, rather than their accumulation at either

553

the top or bottom.

554

containing relatively large droplets (r > 100 nm), whereas Brownian motion forces tend to

555

dominate droplet movement in emulsions containing smaller droplets (McClements, 2011).

556

Consequently, emulsions become more stable to creaming or sedimentation as the particle size

557

decreases because the creaming velocity decreases (v ∝ r2) and because Brownian motion effects

558

increase.

RI PT

SC

M AN U

Gravitational forces tend to dominate droplet movement in emulsions

TE D

559

(3)

9η 0

The above calculations assume that the particles in beverage emulsions are homogeneous spheres consisting entirely of oil phase. In practice, the particles in beverage emulsions actually

561

have a core-shell structure, consisting of an oil core and an interfacial shell. In this case, the

562

overall particle radius is given by rparticle = rcore + δ, and the overall particle density (ρparticle)

563

depends on the densities of the core (ρC) and shell (ρS) materials and the volume fraction of the

564

shell (Φ S):

AC C

EP

560

565

ρ particle = Φ S ρ S + (1 − Φ S ) ρ C

566

The shell layer usually has a higher density than the oil or aqueous phases, so that an

(4)

567

increase in the volume fraction of the shell layer will tend to increase the overall particle density.

568

This has important implications for preventing gravitational separation in beverage emulsions

569

with small droplets sizes since it reduces the density contrast between the particles and aqueous

570

phase. In addition, very small particles may actually sediment rather than cream if they contain

ACCEPTED MANUSCRIPT 21

571

sufficiently thick and dense emulsifier layers. Thus, it should be possible to produce density

572

matched particles in beverage emulsions by controlling the oil core size and the thickness of the

573

adsorbed emulsifier layer. The above discussion has highlighted a number of approaches that can be used to inhibit or

RI PT

574

prevent gravitational separation in beverage emulsions. First, gravitational separation can be

576

prevented by matching the density of the dispersed (oil) and continuous (aqueous) phases. The

577

density of the aqueous phase typically varies from about 1000 to 1050 kg m-3, depending on the

578

amount of sugars and other solutes present (Table 3). The density of most oil phases is less than

579

this value, and therefore oil droplets will tend to move upwards. As already mentioned, the

580

density of the core-shell particles within a beverage emulsion can be matched to the surrounding

581

aqueous phase by adding a weighting agent to the oil phase, or by controlling the thickness and

582

density of the emulsifier layer. Second, gravitational separation can be inhibited by reducing the

583

size of the droplets in the emulsion, since the creaming velocity is proportional to the droplet size

584

squared (Stoke’s Law). If the droplets are sufficiently small, then Brownian motion effects will

585

dominate and the system will remain stable to creaming or sedimentation. Third, gravitational

586

separation can be inhibited by increasing the viscosity of the aqueous phase, e.g., by adding

587

thickening or gelling agents. This approach may not always be viable since it will also influence

588

the texture and mouthfeel of the final product.

M AN U

TE D

589

SC

575

Another approach some beverage manufacturers have used to mask the undesirable effects of creaming (“ringing”) on the appearance of a product is to design the packaging so as to

591

obscure the effect, e.g., with appropriate placement of the labels or cap.

592

5.1.2. Droplet Aggregation

The aggregation state of the droplets in a beverage emulsion is important because it

AC C

593

EP

590

594

influences the stability of the product to gravitational separation. Changes in particle size during

595

storage may also influence other important quality attributes of beverage products, such as their

596

appearance (cloudiness or homogeneity). The tendency for droplet aggregation to occur in a

597

beverage emulsion depends on the balance of attractive and repulsive forces operating between

598

the droplets (see earlier). The nature of the colloidal interactions operating in a particular

599

beverage emulsion depends on the physicochemical properties of the oil, water and interfacial

600

phases (e.g., dielectric constant and refractive index), oil core characteristics (such as radius),

ACCEPTED MANUSCRIPT 22

interfacial shell characteristics (such as thickness, charge, packing, rheology and

602

hydrophobicity), and the properties of the intervening fluid (such as pH, ionic strength, osmotic

603

pressure, and temperature). To a first approximation the overall colloidal interactions between a

604

pair of droplets in a beverage emulsion can be described by the sum of the van der Waals (wVDV),

605

electrostatic (wE), and steric (wS) interactions (McClements, 2005b):

606 607

w(h) = wVDV(h) + wE(h) + wS(h)

(5)

608

The van der Waals interactions are attractive, whereas the steric and electrostatic

SC

609

RI PT

601

interactions are usually repulsive (Table 2). The van der Waals attraction operates between all

611

kinds of droplets and would always cause aggregation if there were no opposing repulsive forces.

612

The magnitude and range of the steric repulsion depend on the thickness and chemistry of the

613

interfacial layer, whereas the magnitude and range of the electrostatic repulsion depend on the

614

droplet charge (ζ-potential) and the ionic composition of the aqueous phase. To design a product

615

that is stable to droplet aggregation one must assure that the repulsive interactions dominate the

616

attractive interactions. This is usually achieved by using an emulsifier that generates repulsive

617

interactions between the droplets. The emulsifiers used in the beverage industry typically

618

stabilize the droplets against aggregation by generating steric and/or electrostatic repulsive

619

interactions. Emulsifiers that form relatively thick open interfaces (such as polysaccharides and

620

non-ionic surfactants with large hydrophilic head-groups) can generate a steric repulsion that is

621

sufficient strong and long range to overcome the attractive van der Waals interactions, and

622

thereby stabilize the system against aggregation. Emulsifiers that form highly charged interfaces

623

(such as proteins and ionic surfactants) can generate a strong electrostatic repulsion between

624

droplets that prevent aggregation. However, emulsifiers that can only stabilize emulsions due to

625

electrostatic interactions may be prone to instability when the pH or ionic strength is changed.

626

Some emulsifiers use a combination of electrostatic and steric repulsion to stabilize the system,

627

e.g., such as casein and whey proteins.

AC C

EP

TE D

M AN U

610

628

The droplets in emulsions are in continual motion because of the effects of thermal energy,

629

gravity, or applied mechanical forces, and as they move about they frequently collide with their

630

neighbors. After a collision, emulsion droplets may either move apart or remain aggregated,

631

depending on the relative magnitude of the attractive and repulsive interactions between them.

ACCEPTED MANUSCRIPT 23

Droplets aggregate when there is a primary or secondary minimum in the interaction potential

633

that is sufficiently deep and accessible to the droplets (Figure 4). The two major types of

634

aggregation in beverage emulsions are flocculation and coalescence.

635 636

5.1.2.1. Flocculation Droplet flocculation is the process whereby two or more droplets come together to form an

637

aggregate in which the droplets retain their individual integrity (Figure 2). Droplet flocculation

638

is usually detrimental to beverage emulsion quality because it accelerates the rate of gravitational

639

separation thereby reducing their shelf-life. Flocculation can also cause an appreciable increase

640

in the viscosity of beverage emulsion concentrates, and may even lead to the formation of a gel.

641

This may be undesirable since it would influence the transport, handling and dispersibility of the

642

product. Flocculation may occur in beverage emulsions through a variety of different processes

643

that either increase the attractive forces or decrease the repulsive forces operating between the

644

droplets. The mechanism that is important in a particular emulsion depends largely on the nature

645

of the emulsifier used and the solution conditions (e.g., pH, ion type and concentration, and

646

functional ingredients).

SC

M AN U

647

RI PT

632

Reduced electrostatic repulsion: Electrostatically stabilized emulsions may flocculate when the electrostatic repulsion between the droplets is reduced. A number of physicochemical

649

changes may cause this reduction in electrostatic repulsion (Israelachvili, 2011): (i) the pH is

650

altered so that the net charge on the droplets is reduced; (ii) counter-ions bind to the surface of

651

the droplets and reduce their charge (“charge neutralization”); (iii) the ionic strength of the

652

aqueous phase is increased to screen the electrostatic interactions (“electrostatic screening”).

653

Protein-coated oil droplets are particularly sensitive to flocculation due to reduction in the

654

electrostatic repulsion between them when the pH or ionic composition is altered (Demetriades,

655

Coupland, & McClements, 1997a; McClements, 2004).

EP

AC C

656

TE D

648

Increased depletion attraction: The presence of non-adsorbing colloidal entities in the

657

continuous phase of an emulsion, such as biopolymers or surfactant micelles, generates an

658

increase in the attractive force between the droplets due to an osmotic effect associated with the

659

exclusion of the colloidal entities from a narrow region surrounding each droplet (Israelachvili,

660

2011). This attractive force increases as the concentration of colloidal entities increases, until

661

eventually it becomes large enough to overcome the repulsive interactions between the droplets and

ACCEPTED MANUSCRIPT 24

causes them to flocculate. This type of droplet aggregation is usually referred to as depletion

663

flocculation. The presence of relatively high concentrations of non-adsorbed biopolymer

664

emulsifiers (gum arabic and modified starch) have been shown to induce depletion flocculation in

665

model beverage emulsions (Chanamai & McClements, 2001). Depletion flocculation may also be

666

promoted by other kinds of biopolymers that might be used in beverages, such as maltodextrin,

667

pectin, xanthan gum, and carrageenan (Cao, Dickinson, & Wedlock, 1990; Cho & McClements,

668

2009; Gu, Decker, & McClements, 2004; Gunning, Hibberd, Howe, & Robins, 1988).

669

RI PT

662

Increased hydrophobic interactions: This type of interaction is important in emulsions that contain droplets that have some non-polar regions exposed to the aqueous phase. A good

671

example of this type of interaction is the effect of thermal processing on the flocculation stability

672

of oil-in-water emulsions stabilized by globular proteins (Demetriades, Coupland, &

673

McClements, 1997b). At room temperature, whey protein stabilized emulsions (pH 7) are stable

674

to flocculation because of the large electrostatic repulsion between the droplets, but when they

675

are heated above 70 oC they become unstable. The globular proteins adsorbed to the surface of

676

the droplets unfold above this temperature and expose non-polar amino acids that were originally

677

located in their interior. Exposure of these non-polar amino acids increases the hydrophobic

678

character of the droplet surface and therefore leads to flocculation because of the increased

679

hydrophobic attraction between the droplets.

TE D

M AN U

SC

670

Formation of biopolymer bridges: Many types of biopolymer promote flocculation by

681

forming bridges between two or more droplets. Biopolymers may adsorb either directly to the

682

bare oil surfaces of the droplets or to the adsorbed emulsifier molecules that form the interfacial

683

layer. To be able to bind to the droplets there must be a sufficiently strong attractive interaction

684

between segments of the biopolymer and the droplet surface. The most common types of

685

interaction that operate in food emulsions are hydrophobic and electrostatic (Dickinson, 2003).

686

For example, a positively charged biopolymer (such as chitosan) might adsorb to the surface of

687

two negatively charged emulsion droplets causing them to flocculate (Ogawa, Decker, &

688

McClements, 2003) or a negatively charged biopolymer (such as pectin, carrageenan or xanthan)

689

might adsorb to the surface of two positively charged droplets causing them to flocculate

690

(Dickinson, 2003; Guzey & McClements, 2006).

691 692

AC C

EP

680

The development of a suitable strategy to prevent droplet flocculation in a particular beverage emulsion therefore depends on identification of the physicochemical origin of

ACCEPTED MANUSCRIPT 25

flocculation in that system. In general, flocculation can be prevented by ensuring that the

694

repulsive forces dominate the attractive forces, and that there are no additives that can promote

695

bridging.

696 697

5.1.2.2. Coalescence Coalescence is the process whereby two or more liquid droplets merge together to form a

698

single larger droplet (Figure 2). Coalescence causes emulsion droplets to cream or sediment

699

more rapidly because of the increase in their particle size. In beverage emulsions, coalescence

700

eventually leads to the formation of a layer of oil on top of the material, which is referred to as

701

oiling off. This process is one of the main reasons for the shiny oily layers often seen on top of

702

unstable beverage emulsions.

SC

The susceptibility of a beverage emulsion to droplet coalescence is highly dependent on the

M AN U

703

RI PT

693

nature of the emulsifier used to stabilize the system, since this instability mechanism involves

705

two or more droplets fusing together. In general, the susceptibility of oil droplets to coalescence

706

is determined by the nature of the forces that act between the droplets (i.e. gravitational,

707

colloidal, hydrodynamic and mechanical forces) and the resistance of the interfacial layer to

708

rupture. The stability of emulsions to coalescence can be improved by preventing the droplets

709

from coming into close proximity for extended periods, e.g., by preventing droplet flocculation,

710

preventing the formation of a creamed layer, or having too high droplet concentrations

711

(McClements, 2005b). Alternatively, one can control the properties of the interfacial layer

712

surrounding the oil droplets to make it more resistant to rupture, e.g., by selecting an appropriate

713

emulsifier or other additives that alter surface properties.

714

5.1.3. Ostwald Ripening

EP

AC C

715

TE D

704

This susceptibility of a beverage emulsion to Ostwald ripening (OR) is mainly determined

716

by the solubility of the oil phase in the aqueous phase: the higher the solubility, the more

717

unstable the emulsion. Oil phases with very low water-solubilities (such as the vegetable oils

718

used in clouding emulsions) do not exhibit OR, but oil phases with relatively high water-

719

solubilities (such as flavor or essential oils) may be highly unstable. Mechanistically, OR is the

720

process whereby the size of the oil droplets in an oil-in-water emulsion increases over time due

721

to diffusion of oil molecules from small to large droplets through the intervening aqueous phase

722

(Kabalnov, 2001; Kabalnov & Shchukin, 1992). The driving force for this effect is the fact that

ACCEPTED MANUSCRIPT 26

the water-solubility of an oil contained within a spherical droplet increases as the radius of the

724

droplet decreases, which means that there is a higher concentration of solubilized oil molecules

725

in the aqueous phase surrounding a small droplet than surrounding a larger one (Kabalnov &

726

Shchukin, 1992; McClements, 2005b). The presence of this concentration gradient means that

727

solubilized oil molecules tend to move from the immediate vicinity of smaller droplets to that of

728

larger droplets. This leads to an increase in mean droplet size over time, which can be described

729

by the following equation once steady state conditions have been achieved (Kabalnov &

730

Shchukin, 1992):

RI PT

723

732

d (t ) 3 − d ( 0 ) 3 = ω t =

32 9

αS ∞ Dt

(6)

M AN U

733 734

SC

731

Here, d(t) is the number-weighted mean droplet diameter at time t, d0 is the initial number-

735

weighted mean droplet diameter, ω is the Ostwald ripening rate, α =2γVm/RT, S∞ is the water-

736

solubility of the oil phase in the aqueous phase, D is the translational diffusion coefficient of the

737

oil molecules through the aqueous phase, Vm is the molar volume of the oil, γ is the oil-water

738

interfacial tension, R is the gas constant, and T is the absolute temperature. The most important factor determining the stability of a beverage emulsion to OR is the

740

water-solubility of the oil phase (S∞) (Weiss, Herrmann, & McClements, 1999). For this reason

741

OR is not usually a problem for emulsions prepared using oils with a very low water-solubility,

742

such as long chain triglycerides (e.g., corn, soy, sunflower, or fish oils). On the other hand, OR

743

may occur rapidly for emulsions prepared using oils with an appreciable water-solubility, such as

744

flavor oils and essential oils (Li, Le Maux, Xiao, & McClements, 2009; McClements, et al.,

745

2012; Wooster, Golding, & Sanguansri, 2008b). OR can be retarded in these systems by adding

746

a substance known as a ripening inhibitor. A ripening inhibitor is a non-polar molecule that is

747

soluble in the oil phase but insoluble in the water phase, e.g., a long chain triacylglycerol (such

748

as corn oil). This type of molecule can inhibit OR by generating an entropy of mixing effect that

749

counter-balances the curvature effects.

AC C

EP

TE D

739

750

Consider an oil-in-water emulsion that contains droplets comprised of two different lipid

751

components: a water-insoluble component and a water-soluble component. The water-soluble

752

molecules will diffuse from the small to the large droplets due to OR. Consequently, there will

753

be a greater concentration of water-insoluble molecules in the smaller droplets than in the larger

ACCEPTED MANUSCRIPT 27

droplets after OR occurs. Differences in the composition of emulsion droplets are

755

thermodynamically unfavorable because of the entropy associated with mixing: it is more

756

favorable to have the two lipids distributed evenly throughout all of the droplets, rather than to

757

be located in particular droplets. Consequently, there is a thermodynamic driving force that

758

operates in the opposite direction to the OR effect. The change in droplet size distribution with

759

time then depends on the concentration and solubility of the two components within the oil

760

droplets. This approach has previously been used to improve the stability of food-grade

761

nanoemulsions, such as those containing short chain triglycerides, essential oils, and flavor oils

762

(Li, et al., 2009; McClements, et al., 2012; Wooster, et al., 2008b). An example of this effect is

763

shown in Figure 7 which shows that droplet growth in orange oil-in-water emulsions during

764

storage can be inhibited by adding a sufficiently high concentration of corn oil (the ripening

765

inhibitor) (McClements, et al., 2012),. Orange oil (4-fold) has a relatively high solubility in

766

water, and therefore is highly prone to OR, which leads to an appreciable increase in mean

767

droplet size during storage. On the other hand, corn oil has a very low solubility in water, and

768

therefore it can retard OR if it is incorporated into the oil phase prior to homogenization. These

769

results show that incorporating ≥ 10% corn oil into the oil phase was sufficient to inhibit OR in

770

these systems (Figure 7). OR may also be retarded by adding certain kinds of weighting agents

771

(such as ester gums) since these substances also have a very low water solubility and therefore

772

act as ripening inhibitors (Lim, et al., 2011).

773

5.2. Chemical Stability

SC

M AN U

TE D

EP

774

RI PT

754

A number of lipophilic compounds that may be present in beverage emulsions can undergo chemical degradation during storage, which leads to a loss of color, flavor and/or nutrients. A

776

few representative examples of chemical degradation of lipophilic components in oil-in-water

777

emulsions are given below.

778

AC C

775

Citrus Degradation. Several mechanisms lead to the chemical decomposition of citrus flavor

779

components (such as citral, d-limonene, and citronellal), including oxidation, hydrolytic

780

reactions, the formation of terpene alcohols, and the cyclization of terpene aldehydes (Clark,

781

Powell, & Radford, 1977; Kimura, Iwata, & Nishimura, 1982; Kimura, Nishimura, Iwata, &

782

Mizutani, 1983a, 1983b). Acid-catalyzed decomposition and oxidation reactions change the

783

desirable flavor profile of citrus oils by reducing the concentration of desirable flavor

ACCEPTED MANUSCRIPT 28

784

components and increasing the concentration of undesirable flavor components (Tan, 2004;

785

Ueno, Masuda, & Ho, 2004). The beverage industry would therefore like to identify effective

786

strategies for preventing these undesirable chemical degradation reactions. There has been a great deal of research on establishing the major factors that influence the

RI PT

787

chemical degradation of citral because this is one of the most important flavor compounds found

789

in commercial beverages. The degradation rate of citral in aqueous solutions has been shown to

790

increase with decreasing pH (Choi, et al., 2009) (Figure 8). Most commercial beverages have

791

acidic aqueous phases and are therefore highly susceptible to flavor loss during storage due to

792

this acid-catalyzed mechanism. The chemical stability of citral has been shown to be much

793

higher when it is located within an oil phase than in an aqueous phase (Choi, et al., 2009).

794

Consequently, the chemical degradation of citral in beverage emulsions can be improved by

795

ensuring that the citral molecules are located primarily in an oil phase rather than in the aqueous

796

phase. Indeed, studies have shown that citral stability can be improved by increasing the oil

797

droplet concentration (Choi, et al., 2009) or by adding surfactant micelles to the aqueous phase

798

(Choi, et al., 2010b), although these strategies may not be practical for most commercial

799

products. It was proposed that citral stability may be improved by encapsulating it within solid

800

lipid particles rather than within liquid oil droplets, however the opposite was found to be true

801

experimentally, which was attributed to the expulsion of the citral molecules into the aqueous

802

phase after droplet crystallization (Mei, et al., 2010). Addition of various kinds of natural

803

antioxidants to flavor oil emulsions has also been shown to improve the stability of citral to

804

chemical degradation (Yang, Tian, Ho, & Huang, 2011). The oil droplets in beverage emulsions

805

are surrounded by a coating of emulsifier molecules, and so it may be possible to improve the

806

stability of the citral molecules within them by engineering the properties of the interfacial layer

807

(Decker & McClements, 2001; Given Jr., 2009). Indeed, studies have shown that citral

808

degradation was faster in flavor oil droplets coated by an anionic surfactant than those coated by

809

a non-ionic or cationic surfactant, which was attributed to differences in the accumulation of

810

catalytic protons near the droplet surfaces (Choi, Decker, Henson, Popplewell, & McClements,

811

2010a). A high local concentration of protons is believed to accelerate the citral degradation

812

mechanism at the droplet surfaces. Coating flavor oil droplets with a cationic biopolymer layers

813

has also been shown to improve the stability of citral to chemical degradation (Djordjevic,

814

Cercaci, Alamed, McClements, & Decker, 2007, 2008; Yang, Tian, Ho, & Huang, 2012)

AC C

EP

TE D

M AN U

SC

788

ACCEPTED MANUSCRIPT 29

815

Polyunsaturated Lipid Degradation. There has been great interest in the beverage industry in fortifying products with ω-3 lipids (such as flax, fish and algal oils) since these lipids have

817

been claimed to have health benefits and are currently under-consumed by the general

818

population. Nevertheless, there are many technical difficulties associated will incorporating

819

these lipids into beverage products due to their high susceptibility to oxidation. Lipid oxidation

820

affects the quality of emulsion-based products, influencing their flavor, odor, and nutritive value

821

(Frankel, Satué-Gracia, Meyer, & German, 2002). The oxidation of polyunsaturated lipids is a

822

highly complex series of chemical reactions that is initiated when a lipid interacts with an

823

oxygen reactive species, and proceeds through molecular cleavage and oxygen addition reactions

824

to the formation of a wide variety of volatile compounds (McClements & Decker, 2000; Waraho,

825

McClements, & Decker, 2011). The rate at which oxidation takes place is dependent on several

826

factors: the molecular structure of the lipids; storage conditions; the presence of pro-oxidants and

827

antioxidants; and the structural organization of the system. Based on this knowledge a variety of

828

strategies have been developed to inhibit or prevent lipid oxidation in emulsified products:

829

addition of oil-soluble and water-soluble antioxidants; chelation of pro-oxidant transition metals;

830

engineering the interface to prevent pro-oxidants from coming into close proximity to lipid

831

substrates; controlling environmental conditions, such as exposure to heat, oxygen, or light.

832

Carotenoid degradation. Carotenoids are natural compounds found in many fruits and

833

vegetables that are may be used in foods an colorants or nutraceuticals because of their potential

834

health benefits (Mayne, 1996; Ryan, O'Connell, O'Sullivan, Aherne, & O'Brien, 2008). One of

835

the major factors currently limiting the incorporation of carotenoids into many food and

836

beverage products is their high susceptibility to chemical degradation. In particular, carotenoids

837

have a conjugated polyunsaturated hydrocarbon chain that makes them highly prone to

838

autoxidation (Boon, McClements, Weiss, & Decker, 2009). A number of factors have previously

839

been shown to promote the oxidation of carotenoids, including highly acidic environments

840

(Konovalov & Kispert, 1999), light (Mortensen & Skibsted, 1996), heat (Mader, 1964), singlet

841

oxygen (Krinsky, 1998), transition metals (Gao & Kispert, 2003; Williams, et al., 2001), and free

842

radicals (Liebler & McClure, 1996; Woodall, Lee, Weesie, Jackson, & Britton, 1997). Once

843

carotenoid degradation has been initiated a number of secondary reaction products may form,

844

including epoxides, endoperoxides, apocarotenals and apocarotenones (Gao & Kispert, 2003;

AC C

EP

TE D

M AN U

SC

RI PT

816

ACCEPTED MANUSCRIPT 30

845

Woodall, et al., 1997; Yamauchi, Miyake, Inoue, & Kato, 1993). The chemical degradation of

846

carotenoids leads to color fading, and may reduce their beneficial health properties.

847

Recent studies have examined the influence of interfacial properties (i.e., emulsifier type), storage conditions (i.e., pH, ionic strength, and temperature) and antioxidant addition (i.e.,

849

vitamin E, Coenzyme Q10, EDTA and ascorbic acid) on the chemical degradation of β-carotene

850

encapsulated within oil-in-water nanoemulsions (Qian, Decker, Xiao, & McClements, 2012).

851

The rate of β-carotene degradation was found to increase with decreasing pH and increasing

852

temperature, was faster for a non-ionic surfactant (Tween 20) than for a protein (β-

853

lactoglobulin), and decreased with increasing antioxidant addition to either the oil or aqueous

854

phase.

855

5.3. Defining the End of Shelf Life

SC

M AN U

856

RI PT

848

The end of the shelf life of a product can be defined as the time when it becomes unacceptable to consumers, which depends on the rate of the various physical and chemical

858

instability mechanisms occurring. A product may become unacceptable when a ring of oil

859

droplets is visible at the top of the bottle, when the flavor components decompose/oxidize and

860

create an unacceptable flavor profile, when the color changes beyond an acceptable level, or

861

when the product is microbiologically unsafe to consume. A beverage manufacture should

862

establish quantitative criteria that can be used to establish the end of the shelf life of their

863

particular product. They should then develop a systematic testing scheme that can be used to

864

predict the shelf life of products.

865

6. Beverage Emulsion Manufacture

AC C

EP

TE D

857

866

Beverage emulsions are usually prepared using a two-step process: a beverage emulsion

867

concentrate (3 – 30 wt% oil) is prepared, which is then diluted extensively to create the finished

868

product (< 0.1 wt% oil) (Tan, 2004). In this section, we briefly describe the major characteristics

869

of the concentrate and finished products, and then review some of the major manufacturing

870

approaches for forming stable beverage emulsions.

871

Beverage emulsion concentrates: When beverages are prepared using high energy

872

homogenization methods (see below), all the aqueous phase components are usually mixed

873

together and all of the oil phase components are mixed together prior to homogenization. The

ACCEPTED MANUSCRIPT 31

aqueous phase often has to be heated and mechanically agitated to facilitate dissolution and

875

dispersion of water-soluble components (such as emulsifiers, thickening agents, buffers, minerals

876

and other functional ingredients). Similarly, the oil phase may also have to be heated and

877

mechanically agitated to facilitate the melting and dispersion of any antioxidants, weighting

878

agents, ripening inhibitors, or colors. Once the oil and aqueous phases have been prepared they

879

are blended together using a high-shear mixer to form a coarse emulsion (d ≈ 1 to 10 µm), which

880

is then homogenized using a mechanical device to form a fine emulsion (d ≈ 0.1 to 1 µm). When

881

beverages are prepared using low energy homogenization methods a different approach may be

882

taken (see below). In this case, water-soluble surfactants and some other water-soluble

883

components may initially be incorporated into the oil phase, which is then mixed with the

884

aqueous phase. This process can lead to the spontaneous formation of a microemulsion,

885

nanoemulsion, or emulsion depending on system composition and preparation procedure. After

886

preparation the beverage emulsion concentrate is often pasteurized to reduce the microbial load,

887

and then stored or transported to the place where it will be used.

M AN U

SC

RI PT

874

Finished Product: The finished product is created by diluting the beverage emulsion

889

concentrate with another aqueous phase, which may contain various other ingredients, such as

890

colors, flavors, preservatives, pH regulators etc. Typically, the concentrate is diluted 500-1000

891

times to produce a final product that often has an oil concentration < 20 mg per liter for a ready-

892

to-drink product (Given, 2009). The final product may be homogenized again to ensure that any

893

non-polar colors, flavors, and preservatives are incorporated into the oil droplets. Appropriate

894

selection of ingredients and processing conditions may lead to beverage products with shelf lives

895

longer than 12 months. However, the perceived quality of a product may deteriorate after

896

extended storage due to detrimental changes in its physical or chemical properties. Beverage

897

emulsions are susceptible to various physical instability mechanisms that can lead to undesirable

898

changes in appearance, such as ringing and oiling off (see earlier section). Beverage emulsions

899

are also liable to undesirable quality changes due to chemical degradation, e.g., changes in flavor

900

profile due to degradation of flavor molecules (such as citral) or color fading due to degradation

901

of colorants (such as carotenoids). These challenges can often be overcome using appropriate

902

packaging materials and/or proper product formation. The design and formulation of successful

903

soft drink products requires careful selection of functional ingredients (oil type, emulsifiers,

904

weighting agents, ripening inhibitors, antioxidants, etc.), preparation conditions (e.g.,

AC C

EP

TE D

888

ACCEPTED MANUSCRIPT 32

905

homogenization method and operating conditions), and storage conditions (e.g., exposure to

906

elevated temperatures, light, and oxygen). A number of different methods can be used to form beverage emulsion concentrates. In

908

general, these approaches can be categorized as either high-energy or low-energy approaches

909

depending on the underlying physical principle of droplet formation (Acosta, 2009; Anton &

910

Vandamme, 2009; Leong, Wooster, Kentish, & Ashokkumar, 2009; Pouton & Porter, 2006;

911

Tadros, Izquierdo, Esquena, & Solans, 2004). High-energy approaches utilize mechanical

912

devices (“homogenizers”) that generate intense forces capable of disrupting and intermingling

913

the oil and aqueous phases leading to the formation of very fine oil droplets (Figure 9). The

914

most commonly used homogenizers utilized in the beverage industry for forming emulsions are

915

high pressure valve homogenizers, but microfluidizers and ultrasonic methods may also be used

916

(Gutierrez, et al., 2008; Leong, et al., 2009; Velikov & Pelan, 2008; Wooster, et al., 2008b).

917

High-energy approaches are probably the most common method used for preparing beverage

918

emulsions at present because they are capable of large-scale production, and they can be used to

919

prepare emulsions from a variety of different starting materials. Low energy approaches rely on

920

the spontaneous formation of fine oil droplets within mixed surfactant-oil-water systems when

921

the solution or environmental conditions are altered (Anton, Benoit, & Saulnier, 2008;

922

Bouchemal, Briancon, Perrier, & Fessi, 2004; Chu, Ichikawa, Kanafusa, & Nakajima, 2007;

923

Freitas, Merkle, & Gander, 2005; Tadros, et al., 2004; Yin, Chu, Kobayashi, & Nakajima, 2009).

924

A number of different low energy approaches have been developed, and some of these are

925

suitable for utilization within the beverage industry, e.g., phase inversion and spontaneous

926

emulsification methods (Figure 10). The minimum particle size that can be produced using

927

either approach depends on many different factors, which are highlighted in the sections below.

928

6.1. High-Energy Approaches

SC

M AN U

TE D

EP

AC C

929

RI PT

907

The size of the droplets generated by high energy approaches is determined by a balance

930

between two opposing processes occurring within the homogenizer: droplet disruption and

931

droplet coalescence (Jafari, Assadpoor, He, & Bhandari, 2008). Only those mechanical devices

932

that are capable of generating extremely intense disruptive forces are capable of producing the

933

tiny droplets required for most beverage emulsion applications (Figure 9), i.e., high-pressure

934

valve homogenizers, microfluidizers, and ultrasonic devices (Leong, et al., 2009; Tadros, et al.,

ACCEPTED MANUSCRIPT 33

2004). The smallest droplet size that can be produced by a particular high-energy device

936

depends on homogenizer design (e.g., flow and force profiles), homogenizer operating

937

conditions (e.g., energy intensity, duration), environmental conditions (e.g. temperature), sample

938

composition (e.g., oil type, emulsifier type, concentrations), and the physicochemical properties

939

of the component phases (e.g., interfacial tension, viscosity) (Kentish, et al., 2006; Wooster, et

940

al., 2008b).

941

RI PT

935

High energy homogenizers are widely used to produce beverage emulsions because they can be utilized with a wide variety of different types of oils and emulsifiers. Once the

943

homogenization conditions have been optimized, beverage emulsions can be produced using

944

triacylglycerol oils or flavor oils as the oil phase, and proteins, polysaccharides, phospholipids,

945

or surfactants as emulsifiers. Thus, high-energy methods are suitable for producing both cloud

946

emulsions and flavor emulsions. Even so, the size of the droplets produced depends strongly on

947

the characteristics of the oil and emulsifier used (see below). For example, it is usually easier to

948

produce very small droplets when the oil phase has a low viscosity and/or interfacial tension

949

(e.g., flavor oils) than when it has a high viscosity and/or interfacial tension (e.g., triacylglycerol

950

oils).

951

6.1.1. High Pressure Valve Homogenizers

M AN U

TE D

952

SC

942

High pressure valve homogenizers are currently the most common high-energy method of producing beverage emulsions. Initially, a coarse emulsion is produced using a high shear mixer

954

and then this is fed directly into the inlet of the high pressure valve homogenizer. The

955

homogenizer has a pump that pulls the coarse emulsion into a chamber on its backstroke and

956

then forces it through a narrow valve at the end of the chamber on its forward stroke (Figure 9).

957

As the coarse emulsion passes through the valve it experiences a combination of intense

958

disruptive forces that cause the larger droplets to be broken down into smaller ones. Different

959

nozzle designs are available to increase the efficiency of droplet disruption. The droplet size

960

produced using a high pressure valve homogenizer usually decreases as the number of passes

961

and/or the homogenization pressure increases. It also depends on the viscosity ratio of the two

962

phases (usually oil and water) being homogenized. Small droplets can only usually be produced

963

when the disperse-to-continuous phase viscosity ratio falls within a certain range (0.05 < ηD/ηC <

964

5) (Tadros, et al., 2004; Walstra, 1993, 2003). Finally, it is important to have sufficient

AC C

EP

953

ACCEPTED MANUSCRIPT 34

emulsifier present to cover all the new droplet surfaces formed during homogenization, and to

966

use an emulsifier that can rapidly adsorb to the droplet surfaces to prevent re-coalescence (Jafari,

967

et al., 2008). Usually, there is a linear relationship between the logarithm of the homogenization

968

pressure (P) and the logarithm of the droplet diameter (d): log d ∝ log P, with the constant of

969

proportionality depending on homogenizer type (McClements, 2005b). To reduce the droplet

970

size to the level required in beverage emulsions it is sometimes necessary to operate at extremely

971

high pressures and to use multiple passes through the homogenizer. Even then, it is only

972

possible under certain circumstances to obtain droplets less than 100 nm in radius (e.g., high

973

emulsifier levels, low interfacial tensions, and appropriate viscosity ratios).

974

6.1.2. Microfluidizers

SC

The formation of beverage emulsions using a microfluidizer also involves forcing a coarse

M AN U

975

RI PT

965

emulsion through a narrow orifice under high pressure to facilitate droplet disruption. However,

977

the design of the channels through which the emulsion is made to flow within a microfluidizer is

978

different from that of a high pressure valve homogenizer (Figure 9). The microfluidizer divides

979

an emulsion into two streams that are then made to impinge on each other in an interaction

980

chamber. Intense disruptive forces are generated within the interaction chamber when the two

981

fast moving streams of emulsion collide, leading to highly efficient droplet disruption.

TE D

976

A number of studies have examined the potential application of microfluidizers for the

983

production of model beverage emulsions (Dalgleish, West, & Hallett, 1997; Henry, Fryer, Frith,

984

& Norton, 2010; Klein, Aserin, Svitov, & Garti, 2010). These studies have shown that small

985

droplets can be produced provided that conditions are optimized to facilitate droplet disruption

986

and inhibit droplet coalescence. The droplet size tends to decrease with increasing

987

homogenization pressure, number of passes, emulsifier concentration, and decreasing disperse-

988

to-continuous phase viscosity ratio (Wooster, et al., 2008b). Again, there is usually a linear

989

relationship between the logarithm of homogenization pressure and the logarithm of the droplet

990

diameter: log d ∝ log P. Recent studies have identified some of the major factors influencing the

991

formation of fine emulsions using microfluidizers (Qian & McClements, 2011). This study

992

showed that the logarithm of the mean droplet diameter decreased linearly with the logarithm of

993

the homogenization pressure for both an ionic surfactant (SDS) and a globular protein (β-

994

lactoglobulin) (Figure 11). However, the slope of the log(d) versus log(P) relationship was

AC C

EP

982

ACCEPTED MANUSCRIPT 35

appreciably steeper for the surfactant (-0.57) than for the protein (-0.29), which was attributed to

996

the fact that the protein may adsorb more slowly to the droplet surfaces, and that it may form a

997

viscoelastic coating that inhibits further droplet breakup. The dependence of the mean droplet

998

diameter on viscosity ratio (ηD/ηC) was also examined by preparing emulsions using different oil

999

phase and aqueous phase compositions. For the ionic surfactant there was a distinct decrease in

1000

mean droplet diameter with decreasing viscosity ratio, which suggested that droplet disruption

1001

within the homogenizer became easier as the viscosity of the two phases became more similar.

1002

On the other hand, little change was found in mean droplet size with viscosity ratio when a

1003

globular protein was used as an emulsifier, which again may be due to the relatively slow

1004

adsorption of the protein and its ability to form a coating that inhibits further droplet disruption.

1005

6.1.3. Ultrasonic Homogenizers

SC

M AN U

1006

RI PT

995

Beverage emulsions can also be formed continuously using ultrasonic homogenizers. This type of homogenizer utilizes high intensity ultrasonic waves to generate intense disruptive forces

1008

(mainly generated by cavitation) that break the oil and water phases into very small droplets

1009

(Kentish, et al., 2006; Leong, et al., 2009; Lin & Chen, 2008). Batch and continuous ultrasonic

1010

homogenizers are available for producing emulsions (Leong, et al., 2009). However, continuous

1011

ultrasonic homogenizers are probably the most commonly used methods for the large scale

1012

production of fine emulsions (Figure 9). The size of the droplets produced using these devices

1013

tends to decrease as the intensity of the ultrasonic waves is increased or the residence time in the

1014

disruption zone is increased (Abismail, Canselier, Wilhelm, Delmas, & Gourdon, 1999; Maa &

1015

Hsu, 1999). The homogenization efficiency also depends on the type and amount of emulsifier

1016

present, and the viscosity of the oil and aqueous phases (Jafari, He, & Bhandari, 2006; Kentish,

1017

et al., 2006; Leong, et al., 2009; Maa & Hsu, 1999). Ultrasonic homogenizers are particularly

1018

suitable for low-viscosity fluids, but are less suitable for more viscous systems.

1019

6.2. Low-Energy Approaches

EP

AC C

1020

TE D

1007

The formation of emulsions using low-energy approaches relies on the spontaneous

1021

formation of oil droplets in surfactant-oil-water mixtures when either their composition or their

1022

environment is altered in a specific way (Anton, et al., 2008; Anton & Vandamme, 2009;

1023

Bouchemal, et al., 2004; Yin, et al., 2009). A number of methods for preparing fine emulsions

1024

are based on the low-energy approach, including spontaneous emulsification and phase inversion

ACCEPTED MANUSCRIPT 36

methods (Figure 10) (Anton, et al., 2008; Anton & Vandamme, 2009; Fernandez, Andre, Rieger,

1026

& Kuhnle, 2004; Maestro, Sole, Gonzalez, Solans, & Gutierrez, 2008). Some of these low

1027

energy methods are already used in the beverage industry for forming oil-in-water emulsions,

1028

whereas others may also be for certain applications.

1029

RI PT

1025

Low energy approaches are often more effective at producing small droplet sizes than high energy approaches, but they are often more limited in the types of oils and emulsifiers that can be

1031

used. For example, it is currently not possible to use proteins or polysaccharides as emulsifiers

1032

in most of the low energy approaches used to form very fine emulsions. Instead, relatively high

1033

concentrations of synthetic surfactants are usually required to form stable emulsions by these

1034

approaches, which limit their use for some beverage applications.

1035

6.2.1. Spontaneous Emulsification Methods

M AN U

1036

SC

1030

In this method an emulsion is spontaneously formed when an organic phase and an aqueous phase are mixed together (Anton & Vandamme, 2009; Miller, 1988; Pouton & Porter, 2006).

1038

The organic phase usually contains oil, hydrophilic surfactant, and possibly water-miscible

1039

solvent, whereas the aqueous usually contains water and possibly cosolvent and cosurfactant.

1040

The two phases can be brought together in various ways: the organic phase may be titrated into

1041

the aqueous phase (Anton & Vandamme, 2009), or the aqueous phase may be titrated into the

1042

organic phase (Sonneville-Aubrun, et al., 2009). When the two phases come into contact the

1043

hydrophilic surfactant and/or solvent move from the oil phase to the aqueous phase, which

1044

generates tiny oil droplets at the phase boundary (Horn & Rieger, 2001). The size of the droplets

1045

produced can be controlled by varying the compositions of the two initial phases, as well as the

1046

mixing conditions.

EP

The spontaneous emulsification method was recently compared with a high energy method

AC C

1047

TE D

1037

1048

(microfluidization) of producing fine emulsions (Yang, Marshall-Breton, Leser, Sher, &

1049

McClements, 2012). The surfactant-oil-water system used consisted of 15.4 wt% non-ionic

1050

surfactant, 23.1 wt% MCT, and 61.5% water, with the surfactant containing a 50:50 mixture of a

1051

hydrophilic (Tween 80) and lipophilic (Tween 85) surfactant. In the high energy method, the

1052

surfactant, oil and water were blended together, and then passed through a microfluidizer. In the

1053

low energy method, the oil and surfactant were mixed together, and then simply titrated into the

1054

water using slow stirring. The microfluidization method produced droplets with a diameter of

ACCEPTED MANUSCRIPT 37

about 110 nm, whereas the spontaneous emulsification method could produce droplets with

1056

diameters around 140 nm. This simple experiment demonstrated that nanoemulsions could be

1057

produced using the spontaneous emulsification method, provided that the system composition

1058

was optimized, i.e., surfactant, oil, and water contents. However, a much higher surfactant-to-oil

1059

ratio (SOR) is required to produce small droplets using the low energy method than with the high

1060

energy method. The influence of SOR on the mean droplet diameter of vitamin E nanoemulsions

1061

produced using the spontaneous emulsification approach is shown in Figure 12. These

1062

measurements show that there is a minimum in the mean droplet size at a particular SOR, and

1063

therefore the composition of the system has to be carefully controlled to obtain a particular

1064

droplet size.

The ability of fine emulsions to be formed using the spontaneous emulsification method

M AN U

1065

SC

RI PT

1055

depends on the system composition (surfactant, oil, and water levels), surfactant type, oil type,

1067

and preparation method (order of addition, mixing speeds). This approach is suitable for

1068

producing flavor or cloud emulsions, provided that an appropriate surfactant is used (such as a

1069

non-ionic surfactant with a particular HLB number) and the surfactant-to-oil ratio is optimized.

1070

The main advantage of this method is that it is inexpensive, simple, and requires no major

1071

equipment, whereas the disadvantage is that relatively high surfactant levels are required (SOR ≈

1072

1)

1073

6.2.2. Phase Inversion Methods

A number of methods have been developed to formulate fine emulsions that depend on

EP

1074

TE D

1066

inducing a phase inversion from a W/O to O/W form, e.g., phase inversion temperature (PIT)

1076

and emulsion inversion point (EIP) methods (Fernandez, et al., 2004; Thakur, Villette, Aubry, &

1077

Delaplace, 2008). In the PIT method the change from one type of an emulsion to another involves a

1078

transitional phase inversion, whereas in the EIP method it involves a catastrophic phase inversion

1079

(Figure 13). A transitional phase inversion occurs when the surfactant properties are altered by

1080

adjusting a formulation variable, such as temperature, pH, or ionic strength. A catastrophic phase

1081

inversion occurs when the ratio of the oil-to-water phases is altered while the surfactant properties

1082

remain constant.

AC C

1075

ACCEPTED MANUSCRIPT 38

6.2.2.1. Phase Inversion Temperature Method The PIT method is based on changes in the optimum curvature and relative solubility of

1085

non-ionic surfactants with changing temperature (Anton, Gayet, Benoit, & Saulnier, 2007; Anton

1086

& Vandamme, 2009; Gutierrez, et al., 2008). Emulsions containing very fine droplets can be

1087

spontaneously formed using the PIT method by varying the temperature-time profile of

1088

appropriate mixtures of surfactant, oil, and water. This type of phase inversion usually involves

1089

the controlled transformation of an emulsion from one type to another (e.g., W/O to O/W or vice

1090

versa) through an intermediate state. This intermediate state is typically a liquid crystalline or

1091

bicontinuous microemulsion phase. The driving force for this type of phase inversion is the

1092

change in the physicochemical properties of the surfactant molecules caused by temperature

1093

changes, e.g., molecular geometry and solubility (Figure 10). The molecular geometry of a

1094

surfactant can be characterized by a packing parameter, p (Israelachvili, 2011):

1096

SC

p=

M AN U

1095

RI PT

1083 1084

aT aH

(7)

where, aT and aH are the cross-sectional areas of the lipophilic tail-group and hydrophilic headgroup respectively. Surfactants spontaneously associate in water to form monolayers that have a

1098

curvature that allows the most efficient packing of the surfactant molecules. The optimum

1099

curvature of a monolayer depends on the packing parameter of the surfactant: for p < 1, the

1100

optimum curvature is convex (favoring O/W systems); for p = 1, monolayers with zero curvature

1101

are preferred (favoring lamellar structures); and, for p > 1 the optimum curvature is concave

1102

(favoring W/O systems). For non-ionic surfactants, the head group is relatively large compared to

1103

the tail group (p < 1) at temperatures below the PIT and so O/W emulsions are favored. Upon

1104

heating, the head group becomes progressively dehydrated and so the packing parameter increases.

1105

At the phase inversion temperature (PIT), the head group and tail group have similar sizes (p = 1)

1106

and so liquid crystals or microemulsions are formed. Above the PIT, the head group is relatively

1107

small compared to the tail group (p > 1) and so W/O emulsions are favored. The relative solubility

1108

of non-surfactants in the oil and water phases also changes with temperature because of head group

1109

dehydration, which has also been used to interpret the PIT phenomenon (Anton, et al., 2007; Anton

1110

& Vandamme, 2009). At low temperatures, the head group is highly hydrated and so the surfactant

1111

tends to be more soluble in water. As the temperature is raised and the head group becomes

1112

progressively dehydrated and the solubility of the surfactant in water decreases, while its solubility

AC C

EP

TE D

1097

ACCEPTED MANUSCRIPT 39

in oil increases. At a particular temperature (≈ PIT), the solubility of the surfactant in the oil and

1114

water phases is approximately equal. At higher temperatures, the surfactant becomes more soluble

1115

in the oil phase than in the water phase. Thus at high temperatures the surfactant is mainly located

1116

in the oil phase, and when the system is cooled it has a tendency to move into the aqueous phase.

1117

Thus, oil droplets are formed by a mechanism similar to spontaneous emulsification.

1118

RI PT

1113

The PIT method is relatively straightforward to implement (Figure 10). A mixture of

surfactant, oil, and water (SOW) is initially heated up to a temperature around or slightly above the

1120

PIT, which leads to the formation of a microemulsion or liquid crystalline phase. The SOW system

1121

is then quench cooled to a temperature well below the PIT with continuous stirring, which leads to

1122

the spontaneous formation of an oil-in-water emulsion or nanoemulsion (Anton & Vandamme,

1123

2009). An example of the phase behavior of a SOW mixture consisting of a non-ionic surfactant

1124

(13% Tween 80), a flavor oil (10 wt% lemon oil), and water (77%) upon heating and cooling is

1125

shown in Figure 14 (Rao & McClements, 2010). Initially, the surfactant, oil and water were

1126

blended together to form a coarse O/W emulsion that was optically opaque. Upon heating the

1127

system becomes transparent when the PIT is reached, and then becomes opaque when heated above

1128

the PIT due to formation of a W/O emulsion. Upon cooling, the system goes from turbid to

1129

transparent indicating that a nanoemulsion was formed (d = 45 nm).

M AN U

TE D

1130

SC

1119

One of the limitations of the phase inversion temperature method is that the emulsions produced are often highly prone to droplet coalescence when they are stored at temperatures

1132

approaching the PIT. This could be a problem in food and beverage applications that require some

1133

form of thermal treatment (such as pasteurization, sterilization or cooking) or that are stored at

1134

elevated temperatures (e.g., in warm or hot climates). Recently, we developed an approach to

1135

overcome this problem by forming emulsions using a non-ionic surfactant with a relatively low PIT,

1136

and then diluting the resulting emulsions in a solution containing another surfactant with a high PIT

1137

(Rao & McClements, 2010). This second surfactant partially displaces some of the original

1138

surfactant from the droplet surfaces, thereby altering the optimum curvature and PIT of the

1139

surfactant monolayer, as well as increasing the repulsive interactions between the droplets. This

1140

approach could be used in the food and beverage industry to form very fine droplets using the PIT

1141

method, and then stabilizing them by diluting them in a different surfactant solution.

1142 1143

AC C

EP

1131

The PIT method can only be used for certain combinations of oils and surfactants. The PIT must be within a practical range that can be implemented within the food industry (e.g. somewhere

ACCEPTED MANUSCRIPT 40

between 40 and 90 ºC) – if it is too high then it will not be possible to create emulsions, and if it is

1145

too low the emulsions formed will be highly unstable to droplet growth. The PIT tends to increase

1146

with increasing molecular weight and hydrophobicity of oil molecules, which means that for many

1147

triacylglycerol oils the PIT is too high to practically reach. On the other hand, flavor oils seem to be

1148

amenable to emulsification using the PIT method. Nevertheless, it is still important to select an

1149

appropriate type and amount of surfactant to ensure the PIT is in the correct range.

1150 1151

6.2.2.2. Emulsion Inversion Point (EIP) Method The emulsion inversion point (EIP) method can also be implemented very easily (Figure 10).

1152

It simply involves titrating increasing amounts of an aqueous phase into an organic phase to induce

1153

a catastrophic phase inversion from a W/O to an O/W system. Initially, an organic phase is

1154

prepared that contains oil and a hydrophilic surfactant. The aqueous phase is then slowly titrated

1155

into this organic phase with constant stirring. As the amount of water added increases the system

1156

converts from a W/O emulsion, to an O/W/O multiple emulsion, to an O/W emulsion (Jahanzad,

1157

Crombie, Innes, & Sajjadi, 2009; Sajjadi, 2006). The formation of these multiple emulsions

1158

requires that the hydrophilic surfactant is initially located in the oil phase. At relatively low

1159

surfactant concentrations the formation of multiple emulsions is suppressed and only relatively

1160

large oil droplets are produced in the final emulsion, which are similar in size to those that would

1161

be produced if the surfactant had been dissolved in the water phase prior to homogenization

1162

(Jahanzad, et al., 2009). At relatively high surfactant concentrations, multiple emulsions are

1163

formed during the titration process, and the final oil droplet size within the O/W emulsions is

1164

determined by the size of the inner oil droplets in the intermediate O/W/O emulsions. The value

1165

of the critical concentration where phase inversion occurs, as well as the size of the oil droplets

1166

produced, depends on process variables, such as the stirring speed, the rate of water addition, and

1167

the emulsifier concentration (Thakur, et al., 2008). The emulsifiers used in catastrophic phase

1168

inversion are usually limited to small molecule surfactants that can stabilize both W/O emulsions (at

1169

least over the short term) and O/W emulsions (over the long term).

SC

M AN U

TE D

EP

AC C

1170

RI PT

1144

We have recently used the EIP method to produce emulsions with relatively small droplet

1171

diameters (< 100 nm) (Ostertag, Weiss, & McClements, 2012). We titrated water into an organic

1172

phase containing a mixture of oil and water-miscible surfactant. The droplet size produced by this

1173

method depended on: (i) oil type: medium chain triglycerides (MCT) < flavor oils (orange and

ACCEPTED MANUSCRIPT 41

limonene) < long chain triglycerides (olive, grape, sesame, peanut and canola oils); (ii) surfactant

1175

type: Tween 80 < Tween 20 < Tween 85; (iii) surfactant concentration: smaller droplets were

1176

produced at higher surfactant-to-oil ratios (SOR); (iv) surfactant location: surfactant initially in oil <

1177

surfactant initially in water. As an example, the data for the influence of oil type on droplet size

1178

produced using the EIP method is shown in Figure 15. The EIP method was also compared to a

1179

high energy method (microfluidization). Small droplets (d < 160 nm) could be produced by both

1180

methods, but that much less surfactant was needed for the high energy method (SOR ≥ 0.1) than the

1181

low energy method (SOR ≥ 0.7). The EIP method may therefore be suitable for producing fine

1182

emulsions suitable for use in the beverage industry without the need for expensive homogenization

1183

equipment, provided that having relatively high amounts of surfactant present is not a problem. In

1184

the finished product, the levels of surfactant may be quite small (even when SOR ≈ 1), since the

1185

beverage emulsion concentrates are diluted so much.

1186

7. Beverage Emulsion Ingredients

1187

7.1. Oil Phase Components

1188

7.1.1. Flavor Oils

SC

M AN U

TE D

1189

RI PT

1174

Flavor oils, such as orange, lemon and peppermint oils, are widely used in the food and beverage industries as flavoring agents, since they contain volatile constituents with

1191

characteristic aroma profiles. The composition of natural flavor oils depends on their biological

1192

origin, the extraction procedure used to isolate them, and any subsequent processing steps. For

1193

example, lemon oil is usually obtained from lemon peel by cold pressing, and it may then be

1194

further refined using steam distillation and various other methods (Misharina, Terenina,

1195

Krikunova, & Medvedeva, 2010). Commercially, flavor oils are available in various forms

1196

(“folds”) that differ in chemical composition due to differences in isolation and processing

1197

procedures used. Flavor oils extracted by cold pressing are usually referred to as single-fold (1×)

1198

oils, while those that have undergone further processing (such as distillation) are referred to as

1199

higher fold oils e.g., three-fold (3×), five-fold (5×), or ten-fold (10×) oils (Gamarra, Sakanaka,

1200

Tambourgi, & Cabral, 2006). Normally, higher-fold oils have different and more intense flavor

1201

profiles than lower-fold oils. Differences in the chemical composition of flavor oils also leads to

AC C

EP

1190

ACCEPTED MANUSCRIPT 42

differences in their physicochemical properties (e.g., water solubility, density, viscosity,

1203

refractive index, and optical properties), which influences their ability to form and stabilize

1204

beverage emulsions (Rao & McClements, 2012a; Rao & McClements, 2012c). In recent studies,

1205

our laboratory has examined the influence of lemon oil fold (1×, 3×, 5× and 10×) on the

1206

formation and properties of oil-in-water microemulsions, nanoemulsions and emulsions (Rao &

1207

McClements, 2012b; Rao & McClements, 2012c). The composition, molecular characteristics,

1208

and physicochemical properties of the four lemon oils were tabulated. The main constituents in

1209

1× lemon oil were monoterpenes (> 90 %), whereas the major constituents in 10× lemon oil were

1210

monoterpenes (≈ 35%), sesquiterpenes (≈ 14%) and oxygenates (≈ 33%). The concentration of

1211

both polar and non-polar components increased with increasing oil fold, while intermediate polar

1212

components decreased. The density, interfacial tension, viscosity, and refractive index of the

1213

lemon oils increased as the oil fold increased (i.e., 1× < 3× < 5× < 10×). The stability of oil-in-

1214

water emulsions produced by high pressure homogenization was strongly influenced by lemon

1215

oil type. Emulsions produced using lower fold oils (1×, 3×, 5×) were highly unstable to droplet

1216

growth during storage, with the growth rate increasing with increasing storage temperature and

1217

decreasing oil fold. Droplet growth was mainly attributed to Ostwald ripening i.e., the diffusion

1218

of water-soluble lemon oil components from small to large droplets. Emulsions prepared using

1219

the highest fold oil (10×) were stable to droplet growth, which was attributed to the presence of

1220

an appreciable fraction of low water-solubility constituents that acted as “ripening inhibitors”

1221

(see later). On the other hand, the lower fold oils were more suitable for forming

1222

microemulsions than the higher fold oils, presumably because they contained constituents that

1223

more easily fit into micelle structures.

SC

M AN U

TE D

EP

A number of recent studies have examined the major factors influencing the formation and

AC C

1224

RI PT

1202

1225

stability of beverage emulsions containing flavor oils, including the preparation method used

1226

(Cheng & McClements, 2011; Rao & McClements, 2011, 2012c), emulsifier type (Mirhosseini,

1227

Tan, Aghlara, et al., 2008; Mirhosseini, Tan, Hamid, & Yusof, 2008; Qian & McClements,

1228

2011), droplet charge (Choi, et al., 2010a), addition of antioxidants (Yang, et al., 2011), presence

1229

of surfactant micelles (Choi, et al., 2010b), and the physical state of the oil phase (Mei, et al.,

1230

2010). Many flavor oils used in beverage emulsions are unstable to chemical degradation during

1231

storage. For example, citral decomposes rapidly during storage at acidic pH by a series of

1232

cyclization and oxidation reactions (Choi, et al., 2009). This effect is highlighted in Figure 16,

ACCEPTED MANUSCRIPT 43

which shows the amount of citral remaining in aqueous solution during storage at different pH

1234

values: the degradation rate increases rapidly as the pH is decreased. Citral degradation reduces

1235

the level of desirable aroma compounds present as well as increasing the levels of undesirable

1236

aroma compounds, thus limiting the shelf life of many beverages. The rate of citral degradation

1237

in emulsions is affected by environmental and compositional factors, such as temperature,

1238

oxygen availability, and antioxidant addition (Choi, et al., 2009; Liang, Wang, Simon, & Ho,

1239

2004a, 2004b; Ueno, Kiyohapa, Ho, & Masuda, 2006; Ueno, et al., 2003). It also depends on the

1240

location of the citral molecules within an emulsion, e.g., oil, water, or interfacial regions (Choi,

1241

et al., 2009; Choi, et al., 2010b). If a labile component is located within an environment where it

1242

is isolated from other components that promote its chemical degradation, then it is possible to

1243

reduce the degradation rate. Since citral degradation occurs predominantly in acidic aqueous

1244

solutions, it is possible to reduce its degradation rate by altering its partitioning between the oil

1245

and aqueous phases. For example, when citral is located predominantly within a non-polar

1246

environment, such as the interior of a lipid droplet or surfactant micelle, it is partially protected

1247

from degradation. This effect is highlighted in Figure 17, which shows that the rate of citral

1248

degradation in oil-in-water emulsions decreases as the total fat droplet concentration increases.

1249

This effect can be attributed to the fact that a lower fraction of the citral is present within the

1250

acidic aqueous phase as the fat content increases due to equilibrium partitioning effects (Figure

1251

18). This phenomenon may be important for beverage emulsion concentrates that have relatively

1252

high fat contents, since in this case an appreciable fraction of the citral will be present within the

1253

oil phase. Choosing an oil type and concentration that ensures that most of the citral is present

1254

within the oil droplets, rather than the aqueous phase, will then help to inhibit citral degradation

1255

during storage. On the other hand, once an emulsion concentrate is diluted to form the finished

1256

product nearly all of the flavor molecules partition into the aqueous phase and therefore any

1257

protective effect is lost.

1258

7.1.2. Cloud Oils

AC C

EP

TE D

M AN U

SC

RI PT

1233

1259

The oil droplets in cloud emulsions are composed mainly of non-flavor oils, such as

1260

triacylglycerol or terpene oils. Triacylglycerol oils are typically derived from natural sources

1261

such as vegetable, corn, canola and sunflower oils, whereas terpene oils are usually isolated by

1262

distillation of natural flavor oils. These oils have very low water-solubilities and are therefore

ACCEPTED MANUSCRIPT 44

stable to Ostwald ripening (unlike flavor oils). The main purpose of cloud oils is to produce

1264

droplets that scatter light strongly so as to give a product a desirable turbid or cloudy appearance,

1265

and so it is important that the droplets are within the size range where efficient light scattering

1266

occurs (typically around 200 to 500 nm). For commercial applications, it is also important that

1267

the droplets remain physically and chemically stable during storage, transport, and handling.

1268

Cloud oils do not usually contribute directly to the flavor profile of beverage emulsions, but they

1269

may indirectly influence flavor due to their effects on the partitioning of flavor molecules

1270

between the oil, water, and headspace regions (McClements, 2005b). They may also adversely

1271

affect the flavor profile of a beverage if they are susceptible to chemical degradation, e.g., the

1272

oxidation of unsaturated lipids products rancid off-flavors.

SC

The physicochemical properties of cloud oils may also play an important role in the

M AN U

1273

RI PT

1263

formation and stability of cloud emulsions. The viscosity of cloud oils will influence their ability

1275

to form small droplets during the homogenization process, with higher viscosities normally

1276

leading to larger droplets. The density of cloud oils determines the creaming velocity of the

1277

droplets within an emulsion, and may therefore influence their long-term stability. The melting

1278

characteristics of cloud oils may also be important for certain applications since partially

1279

crystalline oil droplets are susceptible to partial coalescence leading to droplet aggregation and

1280

phase separation (McClements, 2005b). This may be particularly important for beverages that

1281

are stored in a refrigerator for prolonged periods. In this case, it may be necessary to use an oil

1282

phase that remains liquid at refrigerator temperatures (e.g., winterized or fractionated oil).

1283

7.1.3. Nutraceutical Lipids

EP

1284

TE D

1274

There has been growing interest in the incorporation of various kinds of nutraceutical lipids into commercial food and beverage products, such as polyunsaturated fats, conjugated linoleic

1286

acid, carotenoids, phytosterols, and fat-soluble vitamins. These nutraceutical lipids are either

1287

mixed with the other oil phase components prior to homogenization or they are encapsulated

1288

separately and then introduced intro the product. The incorporation of nutraceutical lipids into

1289

commercial beverage emulsions is often challenging because they are prone to various physical

1290

or chemical degradation mechanisms. Each type of nutraceutical lipid has its own particular

1291

challenges depends on its physicochemical properties, such as solubility, oil-water partition

1292

coefficient, melting point, and chemical stability (McClements, Decker, Park, & Weiss, 2009).

AC C

1285

ACCEPTED MANUSCRIPT 45

Beverage products must therefore be carefully formulated taking into account the potential

1294

problems associated with the specific types of nutraceutical lipids present.

1295 1296

7.1.3.1. Polyunsaturated Lipids There has been growing interest in the incorporation of polyunsaturated fatty acids (PUFAs),

1297

and especially omega-3 (ω-3) fatty acids, into food and beverage products because of their

1298

potential health benefits (Ruxton, Reed, Simpson, & Millington, 2004, 2007; Siddiqui, et al.,

1299

2004). Omega-3 fatty acids are unsaturated fatty acids that have a double bond that is three

1300

carbon atoms from the methyl end of the molecule. The most common ω-3 fatty acids in food

1301

oils are α-linolenic acid (ALA, 18:3), eicosapentaenoic acid (EPA, 20:5) and docosahexaenoic

1302

acid (DHA, 22:6), with EPA and DHA being the most bioactive. PUFAs are normally

1303

incorporated into foods as part of triacylglycerol molecules, but they may also be incorporated as

1304

monoacylglycerols, diacylglycerols, or phospholipids. Consumption of adequate amounts of ω-3

1305

fatty acids has been linked to decreased risks of various diseases, including cardiovascular

1306

disease, immune response disorders, mental disorders, and poor infant development (Orchard,

1307

Pan, Cheek, Ing, & Jackson, 2012; Tur, Bibiloni, Sureda, & Pons, 2012). This has prompted

1308

food manufacturers to attempt to fortify their products with ω-3 fatty acids at levels sufficiently

1309

high to have a beneficial biological effect (McClements & Decker, 2000; Waraho, et al., 2011).

1310

However, PUFAs are extremely susceptible to oxidative deterioration, which causes problems

1311

for the long-term storage of these products (Arab-Tehrany, et al., 2012).

SC

M AN U

TE D

The oxidation of ω-3 fatty acids in emulsions involves a complex series of chemical

EP

1312

RI PT

1293

reactions that results in the production of rancid off-flavors. The control of lipid oxidation often

1314

requires a combination of different approaches to achieve adequate long-term stability, including

1315

control of initial ingredient quality, removal/deactivation of pro-oxidants (such as oxygen or

1316

transition metals), addition of anti-oxidants, and interfacial engineering (McClements & Decker,

1317

2000; Waraho, et al., 2011). A more detailed discussion of the chemical stability of PUFA was

1318

given in an earlier section.

1319 1320

7.1.3.2. Fat Soluble Antioxidants Lipophilic antioxidants can be incorporated into the oil phase of oil-in-water emulsions to

1321

inhibit the oxidation of encapsulated chemically labile substances (such as ω-3 oils or

1322

carotenoids), thereby extending product shelf life (Boon, McClements, Weiss, & Decker, 2010;

AC C

1313

ACCEPTED MANUSCRIPT 46

McClements & Decker, 2000; Waraho, et al., 2011). Alternatively, they can be encapsulated as

1324

functional food components themselves so as to make an “antioxidant” claim on a product label.

1325

Some commonly used lipophilic antioxidants utilized within the food industry include alpha-

1326

tocopherols, ascorbyl palmitate, BHT, BHA, and rosemary extracts (McClements & Decker,

1327

2008). These five molecules act as free radical scavengers and they each have the ability to

1328

absorb two free radicals before becoming inactive. The addition of these antioxidants typically

1329

increases the lag phase of the oxidation reaction.

1330 1331

7.1.3.3. Fat Soluble Vitamins There is considerable interest in incorporating various types of fat-soluble vitamins (A, D, E,

1332

and K) and nutraceuticals (e.g., carotenoids, phytosterols, flavonoids, and curcumin) into food

1333

and beverage products to improve their nutritional value (Sagalowicz & Leser, 2010; Velikov &

1334

Pelan, 2008). In this section, we use recent studies on the incorporation of vitamin E into

1335

emulsions to demonstrate some of the potential benefits and challenges associated with its use in

1336

beverages. The term “vitamin E” actually refers to a group of fat-soluble vitamins that are

1337

widely used as functional ingredients in food, pharmaceutical, and cosmetic preparations, with α-

1338

tocopherol being the most biologically active form (Chiu & Yang, 1992). The major biological

1339

function of Vitamin E appears to be as an oil-soluble antioxidant, although other health benefits

1340

have been claimed, including reducing cardiovascular disease, diabetes, and cancer (Sylvester, et

1341

al., 2011; Traber, Frei, & Beckman, 2008; Weng-Yew & Brown, 2011). For these reasons, there

1342

has been interest in fortifying many foods and beverages with Vitamin E (Sagalowicz & Leser,

1343

2010). Vitamin E is unstable to oxidation and may therefore be lost during processing, storage,

1344

and utilization of commercial products (Gawrysiak-Witulska, Siger, & Nogala-Kalucka, 2009;

1345

Yoon & Choe, 2009). For this reason, vitamin E acetate (rather than vitamin E) is used in many

1346

food and beverage applications since it has a higher oxidative stability. After consumption

1347

vitamin E acetate is broken down to vitamin E in the gastrointestinal tract by the action of

1348

pancreatic esterases (Brisson, et al., 2008). The recommended daily intake (RDI) of Vitamin E

1349

is 15 mg/day (Gonnet, Lethuaut, & Boury, 2010; Institute.of.Medicine, 2000).

1350

AC C

EP

TE D

M AN U

SC

RI PT

1323

Vitamin E is a highly lipophilic molecule that cannot be directly dispersed into aqueous

1351

solutions (Sagalowicz & Leser, 2010). Instead, it must be incorporated into an appropriate

1352

colloidal delivery system prior to dispersion (Gonnet, et al., 2010). A number of previous

ACCEPTED MANUSCRIPT 47

studies have shown that vitamin E can be successfully incorporated into emulsion-based delivery

1354

systems, such as microemulsions (Chiu & Yang, 1992; Feng, Wang, Zhang, Wang, & Liu,

1355

2009b), nanoemulsions (Hatanaka, et al., 2010; Li, et al., 2011; Shukat, Bourgaux, & Relkin,

1356

2012) and emulsions (Chen & Wagner, 2004; Gonnet, et al., 2010). Encapsulation of vitamin E

1357

has been reported to improve its physicochemical stability during storage and its biological

1358

activity after consumption (Cortesi, Esposito, Luca, & Nastruzzi, 2002; Zuccari, Carosio, Fini,

1359

Montaldo, & Orienti, 2005). Indeed, studies have suggested that the oral bioavailability of

1360

Vitamin E is increased when it is delivered in colloidal form rather than in bulk form (Feng,

1361

Wang, Zhang, Wang, & Liu, 2009a), which would make beverages particularly suitable delivery

1362

systems for this type of nutraceutical lipid.

1363 1364

7.1.3.4. Carotenoids Carotenoids are a diverse group of lipophilic compounds that contribute to the yellow,

1365

orange and red colors of many foods, and that have also been claimed to have some health

1366

benefits. They are polyenes consisting of 3 to 13 conjugated double bonds and in some cases 6

1367

carbon ring structures at one or both ends of the molecule. Carotenoids containing oxygen are

1368

known as xanthophylls (e.g. lutein and zeaxanthin) while those without oxygen are known as

1369

carotenes (e.g., lycopene and β-carotene). The carotenoids have been proposed to exhibit several

1370

potential health benefits: lutein and zeaxanthin may decrease age related macular degeneration

1371

and cataracts (Stringham & Hammond, 2005); lycopene may decrease the risk of prostate cancer

1372

(Basu & Imrhan, 2007). Carotenoids are relatively stable to chemical degradation when they are

1373

present in their natural environment. However, they become highly unstable to degradation once

1374

they are isolated, becoming susceptible to light, oxygen, pH, and temperature (Xianquan, Shi,

1375

Kakuda, & Yueming, 2005). Consequently, dispersion of carotenoids into ingredient systems

1376

can result in their rapid degradation (Heinonen, Haila, Lampi, & Piironen, 1997; Ribeiro, Ax, &

1377

Schubert, 2003). Carotenoids can be degraded by reactions that cause the loss of double bonds

1378

or scission of the molecule. In addition, the double bonds in carotenoids can undergo

1379

isomerization to the cis configuration (Xianquan, et al., 2005). Isomerization reactions might

1380

actually be beneficial since cis isomers of carotenoids such as lycopene are thought to be more

1381

bioavailable and bioactive (Schieber & Carle, 2005). An additional challenge to using

1382

carotenoids as ingredients in functional foods is their high melting point, making them crystalline

AC C

EP

TE D

M AN U

SC

RI PT

1353

ACCEPTED MANUSCRIPT 48

at food storage and body temperatures. A more detailed discussion of the factors governing the

1384

chemical degradation of carotenoids was given earlier.

1385 1386

7.1.3.5. Phytosterols Phytosterols and phytostanols are a group of bioactive lipids derived from plants that have

1387

been shown to be have biological activity (Chaiyasit, Elias, McClements, & Decker, 2007). The

1388

fortification of foods and beverages with these components has become popular due to their

1389

ability to decrease total and low density lipoprotein cholesterol in humans, which is mainly

1390

attributed to their ability to inhibit the absorption of dietary cholesterol (Ostlund, 2004; Wong,

1391

2001). Intake of 1.6 g phytosterols/day results in an approximately 10% reduction in LDL

1392

cholesterol (Hallikainen, Sarkkinen, & Uusitupa, 2000). The intestinal absorption of

1393

phytosterols is very low and so dietary phytosterols do not have adverse effects on health.

M AN U

SC

RI PT

1383

Incorporation of phytosterols and phytostanols into aqueous-based beverages is particularly

1395

difficult due to their low-water solubility, low-oil solubility, high melting point, and tendency to

1396

form crystals. Some of these problems have been overcome by esterification of phytosterols

1397

with polyunsaturated fatty acids. After consumption, the phytosterol esters are hydrolyzed by

1398

enzymes in the gastrointestinal tract to produce free fatty acids and phytosterols. The initial

1399

applications of phytosterols were in high fat foods (e.g. margarine and spreads) where

1400

solublization and dispersion within the oil phase were relatively simple.

1401

introduced into aqueous-based foods, they need to be either suspended or emulsified.

1402

Phytosterols are also susceptible to chemical degradation (oxidation) during storage

1403

(Bortolomeazzi, Cordaro, Pizzale, & Conte, 2003; Cercaci, Rodriguez-Estrada, Lercker, &

1404

Decker, 2007; Dutta, 1997; Lambelet, et al., 2003; Soupas, Juntunen, Lampi, & Piironen, 2004).

1405

At present, it is unclear whether oxidized forms of phytosterols lose their bioactivity or have

1406

some toxicity as has been observed for oxidized cholesterol. Phytosterols and phytostanols must

1407

therefore be encapsulated in delivery systems that will maintain their physical and chemical

1408

stability in beverage emulsions.

1409

7.1.4. Fat Soluble Colorants

1410

For phytosterols to be

AC C

EP

TE D

1394

A number of natural fat-soluble compounds have intense colors and can therefore be used as

1411

pigments in foods, including paprika, lycopene, and β-carotene (Gibbs, Kermasha, Alli, &

1412

Mulligan, 1999; Kandansamy & Somasundaram, 2012; Wrolstad & Culver, 2012). Some of

ACCEPTED MANUSCRIPT 49

these components also have biological activity, and may therefore be used as nutraceutical lipids

1414

also. Many colorants from natural sources are highly unstable and may chemically degrade and

1415

fade rapidly during storage, e.g., the carotenoids discussed earlier (Qian, et al., 2012). It is

1416

therefore important to establish the mechanism of chemical degradation for each type of fat

1417

soluble colorant used in a beverage product and to determine the major factors affecting its

1418

degradation (such as pH, light, oxygen, pro-oxidants) so that an effective delivery system can be

1419

designed to ensure appropriate product shelf-life.

1420

7.2. Emulsifiers

1421

7.2.1. Mechanism of Action

SC

Emulsifiers are surface-active molecules that are used in emulsions to facilitate droplet

M AN U

1422

RI PT

1413

breakup within the homogenizer (thereby forming smaller droplets) and to prevent droplet

1424

growth after formation (thereby increasing long-term stability) (McClements, 2005b; Stauffer,

1425

1999). During homogenization, emulsifiers absorb to the oil-water interface, which leads to a

1426

reduction in the interfacial tension thereby facilitating further droplet disruption. Once they have

1427

adsorbed to the droplet surface the emulsifiers should prevent the droplets from coalescing with

1428

each other, which means they must for a protective coating around the oil droplets. An effective

1429

emulsifier system should meet a number of criteria to prevent droplet coalescence within the

1430

homogenizer (Jafari, et al., 2008): (1) the emulsifier molecules must be present at a sufficiently

1431

high concentration to cover all of the oil-water surface formed: (2) the emulsifier molecules must

1432

form a coating around the droplet surfaces faster than the droplets collide with each other; and,

1433

(3) the adsorbed emulsifier molecules must form a coating that prevents the droplets from

1434

coming into close proximity and merging together (McClements, 2005b). There are a number of

1435

food-grade emulsifiers that fulfill these requirements, but they differ considerably in their

1436

effectiveness at forming and stabilizing beverage emulsions.

1437

7.2.2. Factors Influencing Selection

1438

AC C

EP

TE D

1423

Emulsion Formation: Two of the most important factors to consider when selecting an

1439

appropriate emulsifier for forming a beverage emulsion is its ability to produce small droplets

1440

during homogenization, and the minimum amount of emulsifier required to form a stable system

1441

(McClements, 2005b). The performance of an emulsifier can be characterized by plotting the

ACCEPTED MANUSCRIPT 50

mean droplet diameter versus emulsifier concentration under standardized homogenization

1443

conditions, such as homogenizer type, energy input, and duration (McClements, 2007). A

1444

typical dependence of mean droplet diameter on emulsifier concentration for oil-in-water

1445

emulsions produced using a high pressure homogenizer is shown in Figure 19. These curves can

1446

be divided into two regions: an emulsifier-limited region and a homogenizer-limited region.

1447

In the emulsifier-limited region the droplet size decreases with increasing emulsifier

1448

concentration and the size of the droplets is primarily limited by the total amount of emulsifier

1449

present (rather than by homogenization conditions). The minimum size of the stable droplets

1450

that can be produced in this region is given by (McClements, 2007):

1451

SC

6 ⋅ Γ⋅ φ 6 ⋅ Γ⋅ φ = cS c'S (1 − φ )

(8)

M AN U

d min =

RI PT

1442

Here, Γ is the surface load of the emulsifier at saturation (kg m–2), φ is the disperse phase volume fraction (dimensionless), cS is the concentration of emulsifier in the emulsion (kg m–3),

1453

and c′S is the concentration of emulsifier in the aqueous phase (kg m–3). This equation shows

1454

that the droplet size should by linearly related to the inverse of the surfactant concentration in

1455

this region. In this region, the homogenizer initially produces small droplets but there is

1456

insufficient emulsifier present to completely cover them all, and so they tend to coalesce with

1457

each other and form larger droplets under they are covered and protected from coalescence.

1458

TE D

1452

In the homogenizer-limited region there is more than enough emulsifier present to cover all of the droplets formed by the homogenizer, but the disruptive energy generated by the

1460

homogenizer is insufficient to form any smaller droplets. In this region, the droplet size is

1461

therefore limited by the maximum disruptive energy that the homogenizer can generate, rather

1462

than by the emulsifier concentration. In this region, the droplet size could be reduced further by

1463

increasing the homogenization pressure to that smaller droplets are produced.

AC C

EP

1459

1464

Emulsifiers differ in the abilities to form small droplets and in the amounts required to form

1465

stable emulsions due to differences in their molecular and physicochemical characteristics. The

1466

most important physical characteristics are: (i) the rate at which they absorb to the oil-water

1467

interface during homogenization; (ii) the amount that they reduce the interfacial tension after

1468

absorption; and, (iii) the amount required to cover a unit area of interface (surface load).

1469

Typically, the size of the droplets produced under standardized conditions tends to decrease with

1470

increasing absorption kinetics, decreasing interfacial tension, and decreasing surface load.

ACCEPTED MANUSCRIPT 51

Emulsion Stability: The nature of the emulsifier used to form a beverage emulsion has a

1472

pronounced influence on its physical and chemical stability, and so beverage manufacturers must

1473

carefully select an emulsifier with appropriate stability characteristics. When selecting an

1474

emulsifier it is important to establish the set of conditions that it is expected to work under (e.g.

1475

pH range, temperature range, ionic composition, solvent composition, and mechanical abuse),

1476

which is likely to be highly product specific. Emulsifiers may vary considerably in their ability

1477

to form stable emulsions under different environmental conditions. For example, many globular

1478

proteins form stable emulsions at low ionic strengths and pH values far from their isoelectric

1479

point, but tend to flocculate at high ionic strengths or pH values close to their isoelectric point

1480

(McClements, 2004). In addition, they tend to flocculate when heated above their thermal

1481

denaturation temperature due to an increase in droplet surface hydrophobicity. Emulsions

1482

stabilized by some types of non-ionic surfactants are also unstable at elevated temperatures (Rao

1483

& McClements, 2010), in this case because the rate of droplet coalescence increases when the

1484

surfactant approaches its phase inversion temperature (PIT). Emulsions stabilized by

1485

amphiphilic polysaccharides, such as gum arabic and modified starch, tend to be have good

1486

stability to droplet aggregation across a wide range of pH values, ionic strengths, and

1487

temperatures (Chanamai & McClements, 2002b; Charoen, et al., 2011; Qian, Decker, Xiao, &

1488

McClements, 2011). However, these emulsifiers are often not very efficient at producing small

1489

droplets, or they have to be used at high levels. The stability characteristics of a particular

1490

emulsifier are largely determined by its influence on the colloidal interactions operating between

1491

oil droplets (McClements, 2005b). Globular proteins and ionic surfactants tend to form thin

1492

interfacial layers that mainly stabilize droplets against aggregation through electrostatic

1493

repulsion. On the other hand, polysaccharides and non-ionic surfactants tend to form thick

1494

hydrophilic interfacial layers that mainly stabilize droplets through steric repulsion. Emulsions

1495

stabilized by electrostatic interactions are particularly susceptible to environmental changes that

1496

alter the magnitude or range of the electrical charge, e.g., pH and ionic strength. An example of

1497

differences in the electrical characteristics and aggregation stability of emulsions stabilized by a

1498

protein (whey protein isolate) and two polysaccharides (gum Arabic and modified starch) are

1499

shown in Figure 20. The protein-coated droplets tend to aggregate near their isoelectric point

1500

due to their low net charge, whereas polysaccharide-coated droplets are stable across the whole

1501

pH range due to their thick adsorbed layers.

AC C

EP

TE D

M AN U

SC

RI PT

1471

ACCEPTED MANUSCRIPT 52

Legal and Labeling Requirements: Traditionally, many of the emulsifiers used in food and

1503

beverage applications were synthetic small-molecule surfactants due to their low cost and high

1504

efficiency (Kralova & Sjoblom, 2009). However, there has been increasing interest within the

1505

food and beverage industry in replacing synthetic ingredients with natural alternatives so as to

1506

create product labels that are more consumer-friendly. Natural ingredients often have a higher

1507

price than synthetic ingredients, and poorer functional performances. The change from artificial

1508

to natural food components can also increase manufacturing costs and reduce product shelf-life.

1509

In the following sections, we provide a brief overview of some of the synthetic and natural

1510

emulsifiers that are currently available for use in beverage emulsions.

1511

7.2.3. Ingredient Examples

SC

In principle, many different kinds of emulsifiers can be used to stabilize the oil-in-water

M AN U

1512

RI PT

1502

emulsions used in the beverage industry. In practice, the vast majority of commercial beverage

1514

emulsions are made with a limited number of emulsifiers, mainly gum arabic and modified

1515

starch. In this section, we provide an overview of both currently used and potential emulsifiers

1516

that can be utilized in beverage products, and classify them according to the nature of the

1517

molecules involved (i.e., polysaccharides, proteins, or surfactants).

1518 1519

7.2.3.1. Polysaccharide-based emulsifiers Gum Arabic: Currently, gum arabic (also known as gum acacia) is probably the most widely

1520

used emulsifier in the beverage industry to formulate cloud and flavor emulsions (Given, 2009;

1521

Tan, 2004). Gum arabic is a natural exudate harvested from acacia trees in sub-Saharan Africa,

1522

especially Sudan (Buffo, Reineccius, & Oehlert, 2002). Exudates from Acacia seyal and senegal

1523

have both been approved for use in food, but the beverage industry generally acknowledges A.

1524

senegal as the “gold standard” (Reiner, Reineccius, & Peppard, 2010). Acacia senegal consists

1525

of at least three high molecular weight biopolymer fractions. It has been proposed that the

1526

surface-active fraction consists of branched arabinogalactan blocks attached to a polypeptide

1527

backbone (Dickinson, 2003; Jayme, Dunstan, & Gee, 1999; Phillips & Williams, 2000). The

1528

hydrophobic polypeptide chain is believed to anchor the molecules to the droplet surface, while

1529

the hydrophilic carbohydrate blocks extend into the surrounding aqueous phase. The interfacial

1530

layer formed by gum arabic is believed to provide stability against droplet aggregation mainly

1531

through steric repulsion, but with some contribution from electrostatic repulsion also. The

AC C

EP

TE D

1513

ACCEPTED MANUSCRIPT 53

influence of a variety of processing conditions on gum arabic functionality has been examined

1533

(Buffo & Reineccius, 2001; Buffo, Reineccius, & Oehlert, 2001; Buffo, et al., 2002; Chanamai

1534

& McClements, 2002b). For example, it has been shown that gum arabic-coated oil droplets

1535

remain stable to flocculation when exposed to a wide range of environmental conditions, e.g., pH

1536

(3 to 9), ionic strength (0 to 500 mM NaCl; 0 to 25 mM CaCl2) and thermal treatment (30 to 90

1537

ºC) (Chanamai & McClements, 2002b; Charoen, et al., 2011). Nevertheless, gum arabic has a

1538

relatively low affinity for oil-water interfaces compared to most other surface-active

1539

biopolymers, which means that it has to be used at relatively high concentrations to form stable

1540

emulsions. For example, as much as 20 % gum arabic may be required to produce a stable 12

1541

wt% oil-in-water emulsion (Tse & Reineccius, 1995a). For this reason, its application as an

1542

emulsifier is restricted to products that have relatively low droplet concentrations, e.g., beverage

1543

emulsions. In addition, frequent problems have been reported in obtaining reliable sources of

1544

consistently high quality gum arabic, which has led many food scientists to investigate

1545

alternative sources of biopolymer emulsifiers for use in beverages, such as modified starch and

1546

various proteins (Chanamai & McClements, 2002a; Charoen, et al., 2011; Garti, 1999; Kim,

1547

Morr, & Schenz, 1996; Qian, et al., 2011; Tan, 1998; Trubiano, 1995a). Gum arabic has a high

1548

water-solubility and a relatively low solution viscosity compared to other gums, which facilitates

1549

its application as an emulsifier.

SC

M AN U

TE D

1550

RI PT

1532

Modified Starch: Natural (unmodified) starches have poor surface activity due to their hydrophilic glucose backbones. Nevertheless, starches can be chemically modified to make

1552

them effective emulsifiers by attaching non-polar chains along their backbones (Trubiano,

1553

1995b). These types of modified starch are widely used as emulsifiers in the beverage industry.

1554

One of the most commonly used modified starches is an octenyl succinate derivative of waxy-

1555

maize (Stauffer, 1999; Trubiano, 1995b). It consists primarily of amylopectin that has been

1556

chemically modified to contain a side-group that is anionic and non-polar. These side-groups

1557

anchor the molecule to the oil droplet surface, while the hydrophilic starch chains protrude into

1558

the aqueous phase and protect droplets against aggregation through steric repulsion. Because the

1559

dominant stabilizing mechanism is steric repulsion, emulsions stabilized by modified starch are

1560

resistant to changes in pH (3 to 9), ionic strength (0 to 100 mM NaCl; 0 to 25 mM CaCl2) and

1561

temperature (30 to 90 ºC) (Chanamai & McClements, 2002b; Charoen, et al., 2011; Qian, et al.,

1562

2011). Like gum arabic, modified starch has a relatively low interfacial activity (compared to

AC C

EP

1551

ACCEPTED MANUSCRIPT 54

proteins or surfactants), and so a large excess must be added to ensure that all the droplet

1564

surfaces are adequately coated. For example, it is had been recommended that a 1:1 mass ratio

1565

of modified starch to oil was required to produce a stable oil-in-water emulsion (Tse &

1566

Reineccius, 1995a). Nevertheless, there have recently been technical advances in the

1567

manufacture of modified starches that have led to emulsifiers that can be used at much lower

1568

levels (1:5 mass ratio) to form stable emulsions with small droplet sizes (Charoen, et al., 2011;

1569

Qian, et al., 2011). Studies have shown that if the concentration of non-adsorbed modified starch

1570

in the aqueous phase surrounding the oil droplets is too high, it may promote depletion

1571

flocculation (Chanamai & McClements, 2001). Modified starches usually come in powdered or

1572

granular forms that are easily dispersible in cold water.

SC

Other polysaccharides: A number of other surface-active polysaccharides have been studied

M AN U

1573

RI PT

1563

1574

for their potential use as emulsifiers in oil-in-water emulsions. Although these polysaccharides

1575

are not widely used in the beverage industry at present they may find some applications in the

1576

future. There has been debate about the molecular origin of the surface activity of these

1577

polysaccharides, and about whether their ability to stabilize emulsions is primarily due to their

1578

surface activity or their ability to thicken the aqueous phase (Dickinson, 2003). Cellulose is not a good emulsifier in its natural state because it forms strong intermolecular

TE D

1579

hydrogen bonds, which make it insoluble in water. Nevertheless, cellulose can be physically,

1581

chemically, or enzymatically modified to produce food-grade ingredients that have interfacial

1582

activity and can be used as emulsifiers (Garti, et al., 1993), such as methyl cellulose (MC),

1583

hydroxypropyl cellulose (HPC), and methyl hydroxypropyl cellulose (MHPC). These

1584

ingredients are all non-ionic polymers that are soluble in cold water, but tend to become

1585

insoluble when the solution is heated above a critical temperature (around 50 – 90 ºC). They

1586

have good stability to pH (2 – 11), salt and freeze-thaw cycling, which may be beneficial in a

1587

number of beverage emulsion applications.

AC C

EP

1580

1588

Certain types of pectin have been shown to be surface active molecules that can able to

1589

stabilize oil-in-water emulsions, e.g., beet pectin (Jung & Wicker, 2012; Nakauma, et al., 2008).

1590

The origin of this effect is believed to be due to the pectin having some protein molecules

1591

associated with it. The amphiphilic protein portion anchors the pectin molecules to the oil phase,

1592

while the hydrophilic polysaccharide chains protrude into the aqueous phase thereby providing

1593

some steric repulsion. In addition, pectin molecules usually have a relatively high negative

ACCEPTED MANUSCRIPT 55

1594

charge because of the presence of carboxylic groups (pKa = 3.5), and therefore they can also

1595

generate an electrostatic repulsion (Nakauma, et al., 2008).

1596

Soy soluble polysaccharide (SSPS) is a soluble, acidic polysaccharide extracted from okara that has a molecular structure that closely resembles pectin (Nakamura, Takahashi, Yoshida,

1598

Maeda, & Corredig, 2004; Nakamura, Yoshida, Maeda, Furuta, & Corredig, 2004). Similar to

1599

gum Arabic and beet pectin, SSPS is believed to have a protein backbone that anchors the

1600

molecule to the oil phase and hydrophilic anionic polysaccharide chains that protrude into the

1601

aqueous phase and provide stabilization through steric and electrostatic repulsion (Nakauma, et

1602

al., 2008).

SC

1603

RI PT

1597

Corn fiber gum (CFG) is a by-product of the milling of corn (Yadav, Johnston, Hotchkiss, & Hicks, 2007). Preliminary studies have shown that CFG can facilitate the formation and stability

1605

of oil-in-water emulsions (Yadav, et al., 2007). The surface activity of this polysaccharide may

1606

also be due to the presence of small amounts of protein or lipids that are associated with it.

1607

M AN U

1604

Research is continuing in this area, and it is likely that a variety of other polysaccharidebased emulsifiers will be identified that may be suitable for stabilization of beverage emulsions.

1609 1610

7.2.3.2. Protein-based emulsifiers Proteins are widely used in the food industry to stabilize various commercial food and

1611

beverage products, but they are not commonly used to stabilize soft drinks. Surface-active

1612

proteins can be isolated from various sources, including milk, meat, fish, and plants, but the most

1613

widely used in the food industry are those derived from bovine milk: casein and whey proteins.

1614

The interfacial coatings formed by proteins are usually relatively thin and electrically charged,

1615

and hence the major mechanism preventing droplet flocculation in protein-stabilized emulsions

1616

is electrostatic repulsion (rather than steric repulsion). Consequently, protein-stabilized

1617

emulsions are particularly sensitive to pH and ionic strength effects (Figures 20 and 21), and

1618

will tend to flocculate at pH values close to their isoelectric points or when the ionic strength

1619

exceeds a certain level (Demetriades, et al., 1997a, 1997b; Dickinson, 2010; McClements, 2004).

1620

Proteins can typically be used at much lower levels that amphiphilic polysaccharides to

1621

stabilize oil-in-water emulsions. For example, as much as 20 % gum arabic may be required to

1622

produce a stable 12.5 wt% oil-in-water emulsion, whereas less than 1% whey protein can be used

1623

(Chanamai & McClements, 2002a; Tse & Reineccius, 1995b). On the other hand, globular

AC C

EP

TE D

1608

ACCEPTED MANUSCRIPT 56

protein stabilized emulsions are much more susceptible to pH, ionic strength, and thermal

1625

processing than polysaccharide stabilized emulsions as shown in Figure 20 (Chanamai &

1626

McClements, 2002a). The stability of protein-stabilized emulsions to pH, salts, and thermal

1627

processing can often be improved by mixing them with ionic polysaccharides either before or

1628

after homogenization, such as gum arabic, pectin, alginate, and carrageenan (Guzey &

1629

McClements, 2006, 2007; Harnsilawat, Pongsawatmanit, & McClements, 2006; Klein, et al.,

1630

2010). These ionic polysaccharides form an interfacial complex with the adsorbed protein

1631

molecules, which increases the steric and electrostatic repulsion between the droplets

1632

(Dickinson, 2011).

SC

1633

RI PT

1624

A number of other methods have also been developed to improve the emulsifying properties of protein ingredients, including limited hydrolysis to form peptides, modification of protein

1635

structure by chemical, physical, enzymatic or genetic means, and blending of the proteins with

1636

other ingredients, although not all of these methods are legally permitted at present.

1637

M AN U

1634

Globular Proteins: Native globular proteins tend to have fairly compact spheroid structures with the non-polar amino acids contained within their hydrophobic interiors (away from water)

1639

and the polar amino acids located at the exterior (in contact with water). There are many sources

1640

of globular proteins in foods that are surface-active and capable of forming and stabilizing

1641

emulsions, such as those from whey, soy, egg, pea, and various other plants (McClements,

1642

2005b). Globular proteins tend to form relatively thin interfacial coatings around fat droplets,

1643

and therefore primarily stabilize emulsions against flocculation through electrostatic repulsion

1644

(McClements, 2004). They are therefore particularly sensitive to droplet flocculation at pH

1645

values near the isoelectric point (where the droplet charge is low) or at high salt concentrations

1646

(where the droplet charge is screened). Emulsions stabilized by globular proteins are also

1647

particularly sensitive to heat treatments, because these proteins unfold when the temperature

1648

exceeds the thermal denaturation temperature thereby exposing reactive non-polar and sulfhydryl

1649

groups (Kim, Decker, & McClements, 2002). These reactive groups increase the attractive

1650

interactions between droplets, which may lead to droplet flocculation through increased

1651

hydrophobic attraction and disulfide bond formation.

AC C

EP

TE D

1638

1652

The most common source of globular proteins in the food industry is the whey protein

1653

derived from bovine milk. A number of different whey protein ingredients are available for use

1654

in the food industry as emulsifiers, including whey protein concentrate (WPC), whey protein

ACCEPTED MANUSCRIPT 57

isolate (WPI), and highly purified protein fractions such as β-lactoglobulin and α-lactalbumin.

1656

These ingredients normally come as white powders that are dissolved in water prior to

1657

utilization. WPC and WPI are commonly used as emulsifiers in commercial food and beverage

1658

products, whereas β-lactoglobulin and α-lactalbumin are more frequently used in fundamental

1659

studies of protein functionality in research laboratories. Whey protein actually contains a

1660

mixture of globular proteins: ≈ 55% β-lactoglobulin; ≈ 24% α-lactalbumin; ≈ 15%

1661

immunoglobulins; and ≈ 5% serum albumin (McClements, 2005b). The isoelectric point of the

1662

major whey proteins is around pH 5, and therefore droplets stabilized by whey proteins tend to

1663

flocculate around this pH, and especially between about pH 4 to 6 (Figure 20). The thermal

1664

denaturation temperature of whey proteins is around 70 to 80 ºC, and so they are prone to droplet

1665

aggregation when heated above this temperature, especially in the presence of salts that screen

1666

electrostatic interactions (Kim, et al., 2002).

SC

M AN U

1667

RI PT

1655

Flexible Proteins: Another group of protein-based emulsifiers commonly used in foods can be classified as flexible proteins (Dickinson, 1992a). These molecules tend to have a more

1669

flexible random coil type structure (although they often retain some secondary structure). The

1670

main sources of flexible proteins are gelatin from fish and animals, and caseins from bovine

1671

milk. Gelatin is a relatively high molecular weight protein derived from fish, porcine, or bovine

1672

collagen (McClements, 2005b). It is typically prepared by disrupting the native structure of

1673

collagen by boiling it in the presence of either acid (Type A gelatin) or base (Type B gelatin).

1674

The isoelectric point of Type A gelatin (∼7-9) tends be higher than that of Type B gelatin (∼5).

1675

Gelatin exists as a random coil molecule at relatively high temperatures, but undergoes a coil-to-

1676

helix transition upon cooling below a critical temperature, which is about 10 to 25 ºC for animal

1677

gelatin and about 0 to 5 ºC for fish gelatin. Gelatin has been shown to be surface-active and

1678

capable of acting as an emulsifier in oil-in-water emulsions. Nevertheless, it often produces

1679

relatively large droplet sizes when used in isolation, and therefore it may be utilized with other

1680

ingredients to improve its effectiveness at forming and stabilizing emulsions.

AC C

EP

TE D

1668

1681

There are various casein-based emulsifiers that can be purchased commercially with

1682

different compositions and functional properties, such as sodium caseinate, calcium caseinate,

1683

and purified protein fractions, such as a β-casein. There are four main protein fractions in

1684

casein: αS1 (∼ 44%), αS2 (∼ 11%), β (∼ 32%) and κ (∼ 11%). Casein-coated droplets have been

1685

shown to be unstable to droplet flocculation at pH values (3.5 – 5.3) close to the protein’s

ACCEPTED MANUSCRIPT 58

isoelectric point. Caseinate-stabilized emulsions tend to be more stable to heating than whey

1687

protein-stabilized emulsions, presumably because the relatively flexible casein molecules do not

1688

undergo appreciable heat-induced conformational changes like globular proteins do.

1689 1690

7.2.3.3. Small-molecule Surfactants In general, small-molecule surfactants consist of a polar head-group and a non-polar tail

1691

group (Kralova & Sjoblom, 2009; McClements, 2005b; Stauffer, 1999). The head group may be

1692

non-ionic, anionic or cationic, while the tail group may vary in the number, length, and degree of

1693

unsaturation of the chains. Many surfactants can be used to fabricate beverage emulsions using

1694

either low or high energy approaches, which is different from amphiphilic biopolymers that can

1695

only be used to fabricate them using high energy methods. In addition, many surfactants are able

1696

to form smaller oil droplets than amphiphilic biopolymers using high energy methods because

1697

they can adsorb more quickly to the oil-water interfaces during homogenization and lower the

1698

interfacial tension more. Surfactants are also able to stabilize emulsions at lower concentrations

1699

than amphiphilic biopolymers. Despite these advantages, the food industry has been trying to

1700

replace synthetic small molecule surfactants with emulsifiers that are natural and more label

1701

friendly due to consumer demands. Recently, some natural small molecule surfactants have

1702

become commercially available, such as quillaja saponin. A summary of some small-molecule

1703

surfactants that could be utilized in beverage emulsions is given in Table 4.

SC

M AN U

TE D

1704

RI PT

1686

Tweens: Tweens (also sold under the name polysorbates) are synthetic non-ionic surfactants that consist of a non-polar fatty acid group esterified to a polar polyoxyethylene sorbitan group.

1706

The nature of the fatty acid and polyoxyethylene chains determines the nomenclature and

1707

functional properties of the surfactant. Tween 20 (monolaurate), Tween 40 (monopalmitate),

1708

Tween 60 (monostearate), and Tween 80 (monooleate) are hydrophilic molecules with relatively

1709

high HLB values (14.9 and 16). They are therefore predominantly water-soluble, form micelles

1710

in aqueous solutions, and stabilize oil-in-water emulsions. Since Tweens are nonionic they tend

1711

to produce droplets that are stable to aggregation over a wide range of pH and ionic strength

1712

values. However, Tween-coated droplets may become unstable to coalescence at elevated

1713

temperatures close to the phase inversion temperature.

1714 1715

AC C

EP

1705

Sugar esters: Sugar esters are synthetic nonionic surfactants that consist of a polar sugar esterified to a fatty acid chain. There has been growing interest in the utilization of sugar esters

ACCEPTED MANUSCRIPT 59

as emulsifiers within the food and beverage industry (Szuts, Budai-Szucs, Eros, Otomo, &

1717

Szabó-Révész, 2010), which can be attributed to their desirable sensory attributes, low toxicity,

1718

and high biodegradability (Sadtler, Guely, Marchal, & Choplin, 2004). In addition, sugar esters

1719

are produced from natural products, such as sucrose and vegetable oil, and are therefore often

1720

perceived as being more label-friendly than petrochemical-based surfactants (Huck-Iriart,

1721

Candal, & Herrera, 2009). Sucrose fatty acid esters are non-ionic surfactants that have a sucrose

1722

moiety as a hydrophilic head group and one or more fatty acids as lipophilic tail groups. Sucrose

1723

monoesters have a single fatty acid attached to the sucrose molecule, which means they are

1724

relatively hydrophilic, water-soluble, have high HLB numbers, and can stabilize O/W emulsions

1725

(Garti, Clement, Leser, Aserin, & Fanun, 1999; Glatter, et al., 2001). A number of researchers

1726

have previously examined the structure, properties, and functionality of sucrose monoesters

1727

(Fanun, 2010; Garti, Aserin, & Fanun, 2000; Garti, Clement, Fanun, & Leser, 2000; Garti, et al.,

1728

1999). These studies have shown that sucrose monoesters can form a range of different colloidal

1729

structures depending on system composition and temperature, e.g., microemulsions,

1730

nanoemulsions or emulsions. Emulsions formed using sucrose monopalmitate (SMP) as the only

1731

emulsifier are highly unstable to aggregation when stored at acidic pH values (pH 3.5), which

1732

has been attributed to a reduction of droplet charge (Rao & McClements, 2011). However,

1733

emulsions that are stable across a wide range of pH can be produced by mixing SMP with a co-

1734

surfactant (such as lecithin) that increases the electrostatic repulsion between droplets at low pH

1735

(Choi, et al., 2011).

TE D

M AN U

SC

RI PT

1716

Quillaja Saponin: Quillaja saponin is a natural surfactant extracted from the bark of the

1737

Quillaja saponaria Molina tree. It has been found to contain surface active components that are

1738

capable of forming surfactant micelles and stabilizing oil-in-water emulsions (Mitra & Dungan,

1739

1997; Waller & Yamasaki, 1996a, 1996b; Yang, Leser, Sher, & McClements, 2013; Yang &

1740

McClements, 2013). The major components within this extract have been reported to be saponins

1741

(van Setten, ten Hove, Wiertz, Kamerling, & van de Werken, 1998; van Setten, van de Werken,

1742

Zomer, & Kersten, 1995) , which are high molecular weight glycosides consisting of a sugar

1743

moiety attached to a triterpene or a steroid aglycone (Hostettmann & Marston, 1995). The

1744

saponins are surface active substances because they contain both hydrophilic regions (such as

1745

rhamnose, xylose, arabinose, galactose, fucose, and glucuronic acid) and hydrophobic regions

1746

(such as quillaic acid and gypsogenic acid) on the same molecule (Mitra & Dungan, 1997; Sidhu

AC C

EP

1736

ACCEPTED MANUSCRIPT 60

1747

& Oakenfull, 1986). Recent experiments have shown that it can be used to form emulsions

1748

containing small droplets that are stable to changes in pH, ionic strength, and temperature (Yang,

1749

et al., 2013; Yang & McClements, 2013).

1750

A food ingredient based on the quillaja saponin extract has recently been produced by the National Starch Company (Bridgewater, NJ) under the trade name Q-Naturale®. This ingredient

1752

was released in 2008 and has been marketed as a replacement for gum acacia in beverage

1753

emulsions. According to the suppliers, Q-Naturale® can be used at a low level, has high oil

1754

loading, and is stable in some forms of alcoholic beverages, unlike most forms of gum acacia or

1755

modified food starch. It is also organic certified and is stable in products stored at ambient or

1756

cold temperatures for up to a year. The ingredient also comes in a liquid form, so that dissolution

1757

and hydration steps are unnecessary.

SC

M AN U

1758

RI PT

1751

Aside from its use as a beverage emulsion emulsifier, quillaja extract has also been used as a foam-stabilizing agent in beverages. The extract is added to the aqueous phase of a beverage to

1760

stabilize foam that forms at the neck of the beverage. This foam is considered a positive attribute

1761

in root beer and some citrus juices in various parts of the world.

1762 1763

7.2.3.4. Mixed Emulsifier Systems There has been increasing interest in using mixed emulsifier systems with the aim of

1764

producing more label-friendly food ingredients. Proteins tend to be better at producing small

1765

emulsion droplets when used at low concentrations than polysaccharides, whereas

1766

polysaccharides tend to be better at producing emulsions that are more stable to environmental

1767

conditions than proteins, e.g., pH, ionic strength, temperature, freeze-thaw cycling (McClements

1768

2004). The beneficial attributes of these two kinds of biopolymer can be combined to produce

1769

small emulsion droplets with good environmental stability. A number of researchers have shown

1770

that protein-polysaccharide complexes may have better emulsifying properties than either of the

1771

biopolymers used in isolation (Dickinson, 2003, 2011; Guzey & McClements, 2006). These

1772

complexes may be held together either by physical or covalent interactions, and may be formed

1773

either before or after homogenization (McClements 2004). Ingredients based on protein-

1774

polysaccharide interactions will have to be legally acceptable, economically viable and show

1775

benefits over existing ingredients before they find widespread utilization in the food industry. It

AC C

EP

TE D

1759

ACCEPTED MANUSCRIPT 61

1776

should be noted that gum arabic is actually a naturally occurring protein-polysaccharide complex

1777

that is already widely used in the food industry as an emulsifier. As mentioned earlier, sucrose monoesters produce emulsions that are not very stable under

1779

acidic conditions when they are used in isolation, but their stability can be greatly improved by

1780

combining them with co-surfactants, such as lysolecithin.

1781

7.3. Weighting Agents

1782

7.3.1. Mechanism of Action

Weighting agents are additives incorporated into the oil phase of certain types of beverage

SC

1783

RI PT

1778

emulsions to inhibit gravitational separation of the oil droplets (McClements, 2005b). As

1785

discussed earlier, the creaming rate of a droplet within an emulsion is given by Stokes’ law

1786

(McClements, 2005a). This equation predicts that the rate of creaming is proportional to the

1787

droplet radius squared (r2), the density contrast between the two phases (∆ρ), and the reciprocal

1788

of the aqueous phase viscosity (η0) (Equation 3). If ∆ρ is positive the particles move downwards

1789

(sediment), but if ∆ρ is negative they move upwards (creaming). The stability of a beverage

1790

emulsion to gravitational separation can therefore be improved by ensuring that the density of the

1791

oil droplets is similar to that of the surrounding aqueous phase. The densities of some of the

1792

major components found in beverage emulsions are listed in Table 3. The densities of flavor

1793

oils and vegetable oils are considerably lower than those of water and aqueous sugar solutions.

1794

Consequently, droplets containing these oils tend to move upward during storage leading to

1795

“ringing”, i.e., the formation of a visible ring of fat droplets at the surface of the product. The

1796

creaming rate can be reduced by decreasing the density contrast between the oil droplets and the

1797

surrounding aqueous phase (Equation 3). The most common means of increasing the density of

1798

the oil phase so that it matches that of the aqueous phase is to add weighting agents, which are

1799

hydrophobic components that have a density that is considerably greater than water (Table 3).

1800

7.3.2. Factors Influencing Selection

AC C

EP

TE D

M AN U

1784

1801

A number of different weighting agents are available for utilization within commercial

1802

beverage products, such as brominated vegetable oil (BVO), SAIB (Sucrose acetate isobutyrate),

1803

ester gum, or dammar gum (Table 3). There are a number of different factors that will

ACCEPTED MANUSCRIPT 62

1804

determine which of these different components is suitable for application within a specific

1805

beverage product. Physicochemical factors: The bulk physicochemical properties of commercial weighting

1807

agents vary considerably, which alters their ease of use and functional properties. The amount of

1808

a weighting agent that is required to match the density of the oil and aqueous phases depends on

1809

its density: the higher the density, the smaller the amount needed. Weighting agents also vary in

1810

their rheological and solubility properties, which will influence the preparation of the oil phase

1811

and the formation of stable emulsions by homogenization. Many beverage emulsions are

1812

sweetened by adding sugars to the aqueous phase. Sugars increase the aqueous phase density

1813

and would therefore be expected to increase the density contrast between the oil and aqueous

1814

phases, which should accelerate creaming. However, sugars also increase the viscosity of the

1815

continuous phase. These two effects largely cancel each other out, so that the creaming rate in

1816

sugar solutions is not much greater than in pure water (Chanamai & McClements, 2000).

SC

M AN U

1817

RI PT

1806

As mentioned previously, the overall density of an emulsion droplet actually depends on the density and volume of the oil phase and the interfacial layer. The interfacial layer is usually

1819

made up of emulsifiers (such as surfactants, phospholipids, proteins, or polysaccharides) that

1820

have a higher density than both the oil and the water phases. When the thickness of the

1821

interfacial layer is appreciable compared to the dimensions of the oil core, then the overall

1822

density of the emulsion droplet (oil core + interfacial layer) may become similar to that of the

1823

aqueous phase, and so the adsorbed emulsifier acts as a weighting agent itself (McClements,

1824

2011).

EP

1825

TE D

1818

Legal and Labeling Factors: There are legal limits on the amount of each type of weighting agent that can be used in beverages, which must be adhered to when formulating a project.

1827

These legal limits vary from country to country and therefore a manufacturer must be aware of

1828

the appropriate legislation in each country the product will be marketed in. There are also

1829

differences in the “label-friendliness” of different weighting agents. Dammar and ester gum can

1830

be considered to be natural ingredients, whereas BVO and SAIB are synthetic ingredients.

AC C

1826

1831

Brominated vegetable oil (BVO) and ester gum are the two most commonly used weighting

1832

agents in beverage emulsions. However, many countries have restricted the amount of weighting

1833

agents that can be found in the final product. For example, the United States restricts BVO and

1834

ester gum’s use to 15 and 100 ppm per serving, respectively. This low concentration of allowable

ACCEPTED MANUSCRIPT 63

weighting agents restricts the type of products that BVO and ester gum can be used in. Beverage

1836

emulsions are one of the few products where these weighting agents can be used to reduce the

1837

creaming rate, because the oil phase content is so low (<0.1%). It is also possible to blend two or

1838

more types of weighting agents together to achieve density matching but remain below their

1839

individual restriction levels.

1840

7.3.3. Ingredient Examples

1841 1842

7.3.3.1. Brominated Vegetable Oil Brominated vegetable oil (BVO) is made when bromine is added to the double bonds of the

1843

triacylglycerol molecules in corn, soybean, cotton seed, or olive oil. It typically has a density

1844

between 1240 and 1330 kg m-3, and a permitted usage level of 15 ppm per serving in the United

1845

States. Nevertheless, there has recently been consumer concern about the presence of BVO in

1846

soft drinks, which has led some major beverage manufacturers to replace it with more natural

1847

alternatives. In 2013, a leading beverage manufacturer removed BVO from a major product line

1848

after receiving consumer complaints and petitions against its use. Even though BVO is permitted

1849

for use in the United States, it has been banned from use in beverages in the European Union and

1850

Asia.

1851 1852

7.3.3.2. Ester Gum Ester gum is a hydrophobic polymer made when glycerol is esterified to gum rosin

1853

harvested from trees (such as pine trees). It is normally supplied as a crystalline solid that can be

1854

incorporated into the oil phase. Ester gum may not be considered to be a natural food ingredient

1855

due to the esterification step used in its preparation,. Nevertheless, it is derived from natural

1856

components (gum rosin) and non-animal glycerol, and the name “ester gum” may appear to be

1857

more label friendly to consumers than BVO or SAIB. Ester gum has a density of around 1080

1858

kg m-3 and a permitted usage level of 100 ppm per serving in the United States (FDA, USA). It

1859

performs similarly to BVO but a greater concentration of ester gum has to be added to the oil

1860

phase in order to raise the density (Chanamai & McClements, 2000). Aside from its traditional

1861

use as a weighting agent, recent studies have shown that ester gum also retards droplet growth in

1862

model flavor emulsions through Ostwald ripening, i.e., it acts like a ripening inhibitor (Lim, et

1863

al., 2011). This behavior would probably be demonstrated by all weighting agents because they

1864

are highly hydrophobic materials with low water-solubilities.

AC C

EP

TE D

M AN U

SC

RI PT

1835

ACCEPTED MANUSCRIPT 64

7.3.3.3. Sucrose Acetate Isobutyrate Sucrose acetate isobutyrate (SAIB) is a synthetic weighting agent produced by the

1867

esterification of sucrose with acetic and isobutyric anhydrides. SAIB is usually supplied in the

1868

form of a high viscosity transparent liquid that can be mixed with the oil phase prior to

1869

homogenization. It has a molecular weight of ≈ 845 g mol-1, a refractive index of ≈ 1.454, and a

1870

density of ≈ 1146 kg m-3. SAIB also has good stability to chemical degradation due to lipid

1871

oxidation. The level of SAIB in beverages is currently restricted to 300 ppm per serving in the

1872

United States (FDA, USA).

1873 1874

7.3.3.4. Dammar Gum Dammar gum is a natural weighting agent that is isolated from an exudate of

1875

Caesalpinaceae and Dipterocarpaceae shrubs. Of the four weighting agents mentioned in this

1876

review, dammar gum has the lowest density at 1060 kg m-3 (Table 3). Dammar gum is approved

1877

for use in several countries, but it is not currently approved for use in the USA and it so does not

1878

have GRAS status.

1879

7.4. Ripening Inhibitors

SC

M AN U

As discussed previously, a major form of physical instability in many beverage emulsions is

TE D

1880

RI PT

1865 1866

Ostwald ripening, which is the process whereby the mean droplet size grows during storage due

1882

to diffusion of oil molecules from small droplets to large droplets (Kabalnov, 2001; Kabalnov &

1883

Shchukin, 1992). Once steady state conditions have been achieved, the increase in mean droplet

1884

diameter over time can be described by Equation 6 given earlier (Kabalnov & Shchukin, 1992).

1885

This equation predicts that the Ostwald ripening rate increases as the solubility of the oil phase in

1886

the aqueous phase increases. Many flavor oils have a relatively high water-solubility and are

1887

therefore particularly prone to droplet growth through this mechanism (Rao & McClements,

1888

2012b). On the other hand, triacylglycerol oils (such as corn, sunflower, or canola oils) contain

1889

large non-polar triacylglycerol molecules with relatively low water-solubilities, and therefore

1890

cloud emulsions prepared from these types of oil are not prone to Ostwald ripening (Coupland,

1891

Weiss, Lovy, & McClements, 1996).

1892

AC C

EP

1881

An effective means of inhibiting or preventing Ostwald ripening in an emulsion is to add a

1893

hydrophobic component with a very low water-solubility to the oil phase prior to

1894

homogenization (Kabalnov, Pertzov, & Shchukin, 1987). This hydrophobic compound is

ACCEPTED MANUSCRIPT 65

referred to as a “ripening inhibitor” due to its ability to slow droplet growth through this

1896

mechanism (Lim, et al., 2011; McClements, et al., 2012). The ability of low water-solubility

1897

compounds to retard Ostwald ripening in emulsions can be attributed to an entropy of mixing

1898

effect (Kabalnov, et al., 1987), which is discussed in more detail in Section 5.1.3.

1899

RI PT

1895

Recent studies have shown that Ostwald ripening may be inhibited in flavor oil emulsions by adding long chain triglycerides (such as corn oil) (McClements, et al., 2012) or by adding

1901

weighting agents (such as ester gum) (Lim, et al., 2011). It has also been proposed that some of

1902

the water-insoluble components within flavor oils themselves may act as ripening inhibitors,

1903

which may account for the better stability of high-fold lemon oils (e.g., 10×) to Ostwald ripening

1904

than low-fold lemon oils (e.g., 1× and 3×) (Rao & McClements, 2012c).

1905

7.5. Thickening Agents

1906

7.5.1. Mechanism of Action

M AN U

1907

SC

1900

Thickening agents are ingredients that are added to beverage emulsions to increase the viscosity of the aqueous phase, and therefore retard droplet creaming and alter product

1909

mouthfeel. Thickening agents are usually hydrophilic biopolymers that dissolve in the aqueous

1910

phase and occupy a volume that is much larger than the volume of the polymer chain itself

1911

(McClements, 2005b). The effectiveness of a thickening agent at increasing the viscosity of an

1912

aqueous solution is largely determined by its molecular characteristics (e.g., molecular weight,

1913

conformation, and interactions). To a first approximation the viscosity of a dilute aqueous

1914

solution containing a hydrophilic biopolymer can be described by Einstein’s equation:

1917 1918

EP

1916

η = η0 (1 + 2.5 Rvφ)

(9)

AC C

1915

TE D

1908

where φ is the volume fraction of the biopolymer chain, and Rv is the volume ratio (Rv =

1919

Vsphere / Vpolymer), where Vsphere is the effective volume that the biopolymer occupies in solution

1920

(which includes the polymer chain and any trapped solvent), and Vpolymer is the volume occupied

1921

by the polymer chain itself. The higher the value of the volume ratio, the more effective is the

1922

polymer at increasing the viscosity of a solution. Typically, Rv increases as the molar mass,

1923

extension, and self-association of biopolymer molecules increases. Thus, xanthan gum is a

1924

particularly effective thickening agent because it has a high molecular weight and is highly

ACCEPTED MANUSCRIPT 66

extended in aqueous solutions and therefore has a large Rv value. Most thickening agents used in

1926

beverages are long-chain hydrophilic polysaccharides, such as xanthan, locust bean gum, pectin,

1927

or carrageenan. It should be noted that thickening agents can actually promote emulsion

1928

instability under certain conditions, due to their ability to induce bridging or depletion

1929

flocculation. Bridging flocculation can occur when polymer molecules are attracted to the

1930

surfaces of oil droplets, whereas depletion flocculation occurs when they polymer molecules are

1931

dispersed in the continuous phase and their concentration exceeds a particular value.

1932

7.5.2. Factors Influencing Selection

SC

RI PT

1925

There are a number of factors that must be considered when selecting a suitable thickening

1934

agent for use in beverage emulsions. It is important to establish the solution and environmental

1935

conditions under which the thickening agent must function. This usually requires knowledge of

1936

the physicochemical properties of the biopolymer molecules involved, such as helix-coil

1937

transition temperatures (for carrageenan, alginate, pectin), electrical properties (charge groups

1938

and pKa values), hydrophobicity, ion sensitivity, and chemical/enzymatic stability (Whistler &

1939

BeMiller, 1997).

1940

of the biopolymer molecules used (ζ-potential versus pH profile), since electrostatic interactions

1941

influence ingredient interactions, emulsion stability, and mouthfeel. The electrical charge on

1942

biopolymers depends on the nature of any ionic groups along the chain background, as well as

1943

solution conditions (pH and ionic composition). Some biopolymers are neutral (e.g., starch,

1944

cellulose, LBG), some are anionic (e.g., alginate, carrageenan, xanthan), some are cationic (e.g.,

1945

chitosan and polylysine), and some are amphoteric (e.g., gelatin). The magnitude of the

1946

electrical charge on ionic biopolymers depends on the pH relative to the pKa value of the ionic

1947

groups. Anionic polysaccharides tend to be neutral at pH values sufficiently below their pKa

1948

value but negative above, whereas cationic polysaccharides tend to be neutral at pH values

1949

sufficiently above their pKa value but positive below. Proteins are positive below their

1950

isoelectric point, and negative above it (Figure 20). The most common charged groups on

1951

polysaccharide thickening agents are sulfate groups (e.g., carrageenan), carboxyl groups (e.g.,

1952

pectin, alginate, xanthan, carboxymethylcellulose) and amino groups (e.g., chitosan): -SO4H ↔ -

1953

SO4- (pKa ≈ 2); -CO2H ↔ -CO2- (pKa ≈ 3.5); -NH3+ ↔ - NH2 (pKa ≈ 6.5). The most common

1954

charged groups on proteins are carboxyl and amino groups.

M AN U

1933

AC C

EP

TE D

For some applications, it is important to establish the electrical characteristics

ACCEPTED MANUSCRIPT 67

The electrical characteristics of thickening agents in aqueous solutions may also be altered

1956

by interactions with other ionic species in their environment, such as mineral ions, surfactants,

1957

and other biopolymers. Monovalent or multivalent ions such as sodium, potassium, or calcium

1958

may bind to oppositely charged groups on biopolymer chains, altering their overall charge

1959

characteristics. In addition, knowledge of the type of environmental and solution conditions

1960

present within a particular food is often important for selecting the most appropriate thickening

1961

agent. For example, pectin can start to depolymerize when exposed to neutral or alkaline

1962

conditions as a result of a base-catalyzed β-elimination reaction that breaks down its sugar chain.

1963

Over time, this breakdown can result in a decline in viscosity and loss of texture (Sila et al.,

1964

2009).

1965

7.5.3. Ingredient Examples

M AN U

SC

RI PT

1955

There are many different types of thickening agents that can be used in commercial

1967

beverage products. These thickening agents differ in their chemical compositions, molecular

1968

weight distributions, degrees of branching, conformations, and transition temperatures, which

1969

causes variations in their functional performances. A summary of thickening agents that could

1970

be utilized in beverage emulsions is given in Table 5. There are a number of excellent books that

1971

provide detailed information about the molecular, physicochemical, and functional properties of

1972

thickening agents in foods and beverages (Cui, 2005; Imeson, 2010; Phillips & Williams, 2009).

1973

7.6. Sweeteners

1974

7.6.1 Mechanism of Action

EP

1975

TE D

1966

Beverage emulsions are sweetened by adding artificial or natural substances that interact with specific taste receptors on the tongue to provide the perception of sweetness (Fernstrom, et

1977

al., 2012; McCaughey, 2008; Meyers & Brewer, 2008). It has been proposed that mammals

1978

evolved to perceive a pleasant sensation after ingestion of sugar-rich foods because these foods

1979

are a ready source of energy in nature (e.g., fruits and berries). A wide variety of substances

1980

have molecular structures that can induce a sweet sensation, although they vary greatly in the

1981

concentrations required to produce sweetness. Sugar molecules have different sweetness

1982

intensities due to differences in their molecular structures, which alters how they react with the

1983

sweetness receptors in the mouth. Conventionally, the sweetness of substances is compared to

AC C

1976

ACCEPTED MANUSCRIPT 68

that of sucrose (table sugar) which is taken to have a Sweetness of 100. The relative sweetness

1985

of different types of sweetener is established using carefully controlled sensory tests. Fructose is

1986

the sweetest of all common sugars because it has just the right dimensions and bond orientations

1987

to strongly interact with the sweetness receptors in the mouth. However, many natural and

1988

artificial compounds have much strong sweetness intensities than sucrose, and therefore can be

1989

used at much lower levels, which is the basis of many high-intensity sweeteners used in low-

1990

calorie products.

1991

7.6.2 Factors Influencing Selection

SC

1992

RI PT

1984

A number of factors must be considered when selecting a suitable sweetening system for a commercial beverage product. The sweeteners used must provide a flavor profile that consumers

1994

find acceptable and desirable. One of the main limitations of many high-intensity sweeteners is

1995

that they do not produce a flavor profile similar to that normally provided by sugars (Nahon,

1996

Roozen, & DeGraaf, 1996), or that they produce undesirable off-flavors (such as bitterness or

1997

metallic) or mouthfeel (such as astringency) (DuBois & Prakash, 2012). In addition, the

1998

sweeteners must remain physically and chemically stable within the product throughout

1999

processing, transport, and storage. Some sweeteners may breakdown when exposed to elevated

2000

temperatures and are therefore unsuitable for use in beverages that undergo thermal processing.

2001

Concern about overweight and obesity has driven consumer demand for reduced calorie

2002

products, which has led the beverage industry to produce low- and mid-calorie versions of their

2003

conventional high-calorie products. However, many of the high-intensity sweeteners used to

2004

replace sugars have undesirable flavor profiles or are perceived as being “unnatural”, and so the

2005

beverage industry is trying to introduce natural low-intensity sweeteners with desirable flavor

2006

profiles (such as Stevia).

2007

7.6.3. Ingredient Examples

AC C

EP

TE D

M AN U

1993

2008

There are many different types of sweeteners that can be used in commercial beverage

2009

products, which vary according to their chemical structure, origins (natural versus artificial), and

2010

intensity (low- versus high-intensity). The most commonly used low-intensity sweeteners are

2011

sugars (such as sucrose, glucose, and fructose) and sugar alcohols (such as sorbitol, mannitol,

2012

xylitol and erythritol). These substances may be added as individual ingredients, or they may be

2013

part of more complex ingredients, such as high fructose corn syrup, honey, and sugar cane. The

ACCEPTED MANUSCRIPT 69

human body does not have the appropriate enzymes in its digestive tract to break down sugar

2015

alcohols in the same manner that sugars are broken down. Consequently, the body breaks sugar

2016

alcohols down more slowly and less efficiently that sugars, which means they have less calories

2017

per gram (about 50-75%), which may be useful for the development of reduced calorie products.

2018

However, high-intensity sweeteners are more commonly used to produce low-calorie or medium-

2019

calorie beverages, which may be artificial (such as Cyclamate, Aspartame, Acesulfame K,

2020

Saccharin, Sucrolose, and Neotame) or natural (such as Stevia).

2021

8. Commercial Aspects of Beverage Emulsion Formulation

SC

2022

RI PT

2014

Beverage emulsions are developed internally and externally by commercial manufacturers of soft drinks products. External development is usually contracted to flavor houses that often

2024

compete with each another for a flavor contract with a beverage manufacturer. This contract can

2025

be worth thousands to millions of dollars depending on the product and length of the contract.

2026

Factors considered for selecting the most appropriate flavor delivery system include stability,

2027

cost, volume capacity, safety audits, kosher/halal certification, and the ingredient line of the

2028

beverage emulsion. Because of these various factors, the best-tasting, most stable flavor may not

2029

always be chosen as the final flavor used in a commercial application.

TE D

2030

M AN U

2023

Beverage manufacturers normally supply flavor houses with a base of the intended product into which the flavor is intended to be incorporated. A base consists of all ingredient components

2032

except the flavor. By doing this, the manufacturers do not have to disclose confidential

2033

formulations and the flavor houses can develop a flavor specific to the product. Shelf-life studies

2034

(accelerated and non-accelerated) are performed to determine how well the prototype flavors

2035

hold up in the formulation. If chemical or physical instabilities are observed, the flavor can be

2036

reformulated to increase its stability. Depending on the project’s timeline this can take as little as

2037

a few days or up to a year. If the product is being rushed to market and only a few days are

2038

allotted for formulation, the flavor house can rely on past knowledge of how flavors behave in

2039

similar products. This can enable the fast development of a flavor that will have a high-

2040

probability of success in a given formulation. If development is done internally, the same steps

2041

are followed: prototype formulation; shelf-life studies; reformulation (if necessary); and then

2042

commercialization.

AC C

EP

2031

ACCEPTED MANUSCRIPT 70

2043

9. Recent Developments in Beverage Emulsions There have been a number of recent developments and innovations in the beverage industry

2045

over the past few years. In this section, we highlight a number of the most important ones. The

2046

compositions and calorie content of a number of common commercial beverages is summarized

2047

in Table 6, whereas some of the claims that can be made on beverages in given in Table 7.

2048

9.1. “All-natural” Products and Cleaner Ingredient Lines

RI PT

2044

Consumers’ preference for natural products over artificial have caused manufacturers to

2050

claim their products as “Natural” whenever possible. As attractive as a “Natural” claim is, it

2051

comes at a product functionality and financial cost. In most cases, natural ingredients are more

2052

expensive than synthetic and their functional properties are often not as robust. The use of

2053

synthetic ingredients creates a cheaper, more stable product, but also one less attractive to many

2054

consumers. A 2007 study by the University of Southampton (UK) linked synthetic colors and

2055

additives to heightened hyperactivity in children (McCann, et al., 2007). This report led to a

2056

consumer outcry for the removal of synthetic ingredients from food and beverage products aimed

2057

at children, promoting many large companies to begin removing synthetic additives and colors.

2058

One problem with replacing artificial additives with natural ones is that they are often much less

2059

stable within commercial products. For example, natural colors may rapidly degrade in the

2060

presence of acid, light, and heat. Consequently, many commercial products have had to be

2061

reformulated to include effective strategies to extend their shelf-life, e.g., using antioxidants and

2062

other preservatives to slow color fading.

M AN U

TE D

EP

2063

SC

2049

The use of synthetic ingredients not only impacts consumer acceptance, but also limits retail channels a product can compete within. Certain local and national grocery chains prohibit

2065

artificial sweeteners, preservatives, and flavors in the foods they sell. The incorporation of

2066

brominated vegetable oil (BVO), sucrose-acetate isobutyrate (SAIB), Tweens (polysorbate) or

2067

modified-food starch into a product can restrict product sales in these marketplaces. Thus

2068

beverage suppliers have to use alternative weighting agents (such as dammar gum or ester gum)

2069

or emulsifiers (such as gum Arabic) in these products.

2070

AC C

2064

The removal of all synthetic components of a product may be unachievable for functional

2071

and financial reasons. However, manufacturers can remove as many synthetic components as

2072

possible to make “No, No, No” claims. For instance, an unnatural product can still claim “No

ACCEPTED MANUSCRIPT 71

artificial colors, preservatives, flavors, and sweeteners”. A “No, No, No” claim is not as strong

2074

as a “Natural” claim, but it may be seen favorably by consumers. Claims on product packaging

2075

are regulated by the US Food and Drug Administration within the Code of Federal Regulations

2076

(CFR), specifically Title 21.

2077

9.2. Low-Calorie and Mid-Calorie Soft Drinks

2078

RI PT

2073

Consumer perceptions about the potential link between overconsumption of sugary drinks and obesity promoted beverage manufacturers to create a variety of low or zero-calorie products

2080

(Slavin, 2012). These products have been on the market for numerous years, but many

2081

consumers do not like their taste, and so the beverage industry has responded by developing a

2082

number of mid-calorie soft drinks. These drinks contain blends of caloric (e.g., Sucrose or High-

2083

Fructose Corn Syrup) and high-potency sweeteners (e.g., Acesulfame-Potassium, Aspartame,

2084

Rebiana, and Sucralose) to reduce the total calories in sweetened beverage products (typically by

2085

over a half compared to regular products), while still maintaining a desirable flavor profile.

2086

Historically, these kinds of products have not sold well in the marketplace (with a number of

2087

products from different companies failing), but there have been renewed attempts to introduce

2088

them again recently in response to consumer desire for reduced calorie intake.

2089

9.3. Energy Drinks

TE D

M AN U

SC

2079

Energy drinks are marketed as products that can enhance physical or mental activity by

2091

providing a rapid source of energy, nutrients, and stimulants. These drinks may be taken before,

2092

during, or after some proposed activity. They usually contain a blend of ingredients, such as

2093

water, sugars, minerals, proteins, vitamins and/or stimulants (such as caffeine) that are proposed

2094

to replenish the bodies’ reserves and enhance performance. Typically, energy drinks come as a

2095

Ready-to-Drink beverage (serving ~ 8 fluid ounces), Ready-to-Drink shot (serving ~ 2 fluid

2096

ounces), or Liquid Concentrate (dilute beverage serving ~ 8 fluid ounces). All the major

2097

commercial beverage manufacturers have products in this rapidly growing product category.

2098

Emulsion technology may be used in these products to deliver oil-soluble flavors, clouding

2099

agents, preservatives, or nutrients. Further scientific research is required to demonstrate the

2100

claimed biological activity of many of these products. Recently, there has been concern about

2101

the potential risks to certain individuals associated with consuming the large amounts of caffeine

2102

in some of these products.

AC C

EP

2090

ACCEPTED MANUSCRIPT 72

2103

9.4. Beverage Concentrates Beverage concentrates are products that are sold in a concentrated form that consumers then

2105

dilute with water prior to consumption. Concentrates may be in either a fluid or powdered form

2106

and contain ingredients such as sugars, flavors, colors, and nutrients that form a drinkable

2107

beverage upon dilution. Beverage concentrates have been used since the nineteenth century

2108

when soda syrups were diluted by pharmacists prior to consumption by consumers. This process

2109

still occurs within automated soda fountains found in many fast-food restaurants, and within

2110

some home soda beverage makers. More recently, beverages have been developed that contain

2111

high-intensity sweeteners rather than sugars, which enable higher concentration factors to be

2112

attained. One recent successful product based on this principle involves squirting a small

2113

amount of the concentrate from a plastic bottle into a glass of water to form a zero-calorie

2114

beverage. A number of beverage manufacturers have launched products within this category in

2115

recent years.

M AN U

SC

RI PT

2104

Beverage concentrates have some unique developmental issues due to the fact that the

2117

ingredients are present in a highly concentrated form. For example, high acid, sweetener, flavor,

2118

buffer, nutraceuticals and/or color concentrations can lead to accelerated chemical degradation

2119

reactions and shorter shelf lives.

2120

9.5. Fortification with Vitamins, Minerals, and Nutraceuticals

TE D

2116

As discussed earlier, there is a push to fortify many beverage products with ingredients that

2122

are perceived as giving added health benefits, such as vitamins, minerals, antioxidants, peptides,

2123

proteins, dietary fibers, and ω-3 fatty acids. These nutraceutical additives vary greatly in their

2124

molecular, physicochemical, and biological properties, such as physical state, solubility, charge,

2125

polarity, interactions, and flavor profiles. For example, carotenoids and polyunsaturated lipids

2126

may oxidize during storage and produce undesirable flavors or changes in color. Multivalent

2127

mineral ions (such as calcium) may precipitate from solution, promote precipitation of other

2128

components (such as proteins), or cause undesirable mouthfeel. Dietary fibers may increase the

2129

viscosity of a product, promote depletion flocculation, or give a slimy mouthfeel. Consequently,

2130

each beverage product must be carefully formulated to take into account the specific types of

2131

nutraceutical components it contains. This segment of the beverage industry is growing rapidly

AC C

EP

2121

ACCEPTED MANUSCRIPT 73

2132

with consumer demand for products that are perceived to be more beneficial to human health and

2133

wellness.

2134

9.6 Bottled Water There has been an appreciable increase in the global sales of bottled waters in recent years,

RI PT

2135

including enhanced, flavored, and fruit-flavored waters. Enhanced water is fortified with

2137

electrolytes, vitamins and other nutraceuticals, whereas flavored and fruit-flavored waters are

2138

flavored with flavorants and fruit extracts, respectively. Utilization of conventional emulsions in

2139

bottled waters to deliver oil-soluble flavors and nutraceuticals is limited because of the increase

2140

in turbidity caused by light scattering from the oil droplets. However, microemulsions and

2141

nanoemulsions that contain very fine droplets that do not scatter light strongly may be used for

2142

this purpose.

2143

9.7. New Product Innovation

M AN U

2144

SC

2136

New product innovation drives business growth and enables manufacturers to meet current consumer needs. Predicting future trends and having the first product to market can give

2146

manufacturers a huge advantage against competitors. However, being first to market is also a risk

2147

due to the possibility of a new product being rejected by consumers. Reasons for rejection may

2148

include product unfamiliarity or a lack of consumer need for the product.

2149

TE D

2145

One objective of new product innovation is to develop a product that is so novel that it can be patented. The federal government then grants the manufacturer a 20-year monopoly on that

2151

product type. A patented product can be the main driver of strong revenue growth for years, and

2152

a manufacturer can gain additional revenue through lawsuits due to patent infringement. On the

2153

other hand, patent expiration can also drive significant revenue loss once competitors have open

2154

access to the once-patented concept.

AC C

EP

2150

2155

New product innovation does not necessarily mean creating the first prototype of a new

2156

product. Studying trends in other parts of the world and then modifying them to fit the intended

2157

marketplace can also be used to drive innovation. For instance, Acai Berry has been consumed in

2158

Brazil for decades, but it was only in the 2000’s that it became a desired component of food

2159

products in the United States. Suspending grains in beverage products has been carried out in

2160

South America and Europe for many years, but has only recently found its way into North

2161

America.

ACCEPTED MANUSCRIPT 74

Determining a secondary use of manufacturing byproducts can also generate revenue

2163

through the development of functional food ingredients from material once destined for waste.

2164

For example, whey protein was once a byproduct of cheese production, but is not a widely used

2165

functional ingredient in foods and beverages. Fusing elements from two different product

2166

categories is another way to innovate.

RI PT

2162

New product development is an integral part of the food industry and changing out

2168

components of a beverage to meet current demands is a struggle that manufacturers have to

2169

consider every day. Such demands include the removal of ingredients from current products that

2170

consumers perceive as undesirable (such as artificial dyes or High-Fructose Corn Syrup), the

2171

addition of desirable components (such as antioxidants, minerals, or nutraceutical lipids), or the

2172

replacement of expensive ingredients with inexpensive alternatives.

2173

10. Conclusions

M AN U

2174

SC

2167

In principle, beverage emulsions are one of the simplest forms of oil-in-water emulsions used within the food industry, consisting of small oil droplets dispersed within an aqueous

2176

medium. Nevertheless, there are still many challenges associated with developing successful

2177

commercial products. Beverage emulsions are prepared from various ingredients and processing

2178

operations that may vary from run-to-run, which can cause changes in the properties and stability

2179

of finished products. To minimize any undesirable changes in product quality it is important that

2180

beverage manufacturers develop a robust initial formulation and manufacturing operation, and

2181

that they also ensure that ingredient quality and manufacturing conditions are standardized and

2182

monitored. The beverage industry must also reformulate products to meet changing market and

2183

consumer demands: to increase profits by replacing expensive ingredients with less expensive

2184

ones; to remove ingredients that consumers perceive as being undesirable, such as sugar, salt,

2185

and artificial colors, flavors, or preservatives; to fortify beverages with ingredients that

2186

consumers perceive as desirable, such as vitamins, minerals, proteins, nutraceuticals, and fibers.

2187

A thorough understanding of the science and technology underlying the formation, properties,

2188

and stability of beverage emulsions facilitates the development of innovative products, and

2189

allows any production problems to be solved more rapidly. This review article has attempted to

2190

provide a comprehensive overview of the scientific and technological principles underlying

2191

beverage emulsion development.

AC C

EP

TE D

2175

ACCEPTED MANUSCRIPT 75

2192 2193

11. Acknowledgments DJM would like to thank Prof. Eric Dickinson for introducing him to the area of food biopolymers and colloids – it has led to many years of interesting and stimulating work, for

2195

which I am very grateful. We would also like to thank a number of colleagues in industry and

2196

academia who have provided valuable comments and suggestions on the manuscript, including

2197

Ratjika Chanamai, Eric Decker, Yuan Fang, Anil Gaonkar, and Peter Given. We would also like

2198

to acknowledge that this material is partly based upon work supported by grants from the United

2199

States Department of Agriculture (CREES, NRI Grant and NSF-EPA-USDA Grant) and the

2200

Massachusetts Department of Agricultural Resources.

2201

12. References

2202 2203 2204 2205 2206 2207 2208 2209 2210 2211 2212 2213 2214 2215 2216 2217 2218 2219 2220 2221 2222 2223 2224 2225 2226 2227 2228 2229 2230

Abismail, B., Canselier, J. P., Wilhelm, A. M., Delmas, H., & Gourdon, C. (1999). Emulsification by ultrasound: drop size distribution and stability. Ultrasonics Sonochemistry, 6(1-2), 75-83. Acosta, E. (2009). Bioavailability of nanoparticles in nutrient and nutraceutical delivery. Current Opinion in Colloid & Interface Science, 14(1), 3-15. Anton, N., Benoit, J. P., & Saulnier, P. (2008). Design and production of nanoparticles formulated from nano-emulsion templates - A review. Journal of Controlled Release, 128(3), 185-199. Anton, N., Gayet, P., Benoit, J. P., & Saulnier, P. (2007). Nano-emulsions and nanocapsules by the PIT method: An investigation on the role of the temperature cycling on the emulsion phase inversion. International Journal of Pharmaceutics, 344(1-2), 44-52. Anton, N., & Vandamme, T. F. (2009). The universality of low-energy nano-emulsification. International Journal of Pharmaceutics, 377(1-2), 142-147. Arab-Tehrany, E., Jacquot, M., Gaiani, C., Imran, M., Desobry, S., & Linder, M. (2012). Beneficial effects and oxidative stability of omega-3 long-chain polyunsaturated fatty acids. Trends in Food Science & Technology, 25(1), 24-33. Basu, A., & Imrhan, V. (2007). Tomatoes versus lycopene in oxidative stress and carcinogenesis: conclusions from clinical trials. European Journal of Clinical Nutrition, 61(3), 295-303. Benichou, A., Aserin, A., & Garti, N. (2004). Double emulsions stabilized with hybrids of natural polymers for entrapment and slow release of active matters. Advances in Colloid and Interface Science, 108-109, 29-41. Berli, C. L. A., Deiber, J. A., & Quemada, D. (2005). On the viscosity of concentrated suspensions of charged colloids. Latin American Applied Research, 35(1), 15-22. Boon, C. S., McClements, D. J., Weiss, J., & Decker, E. A. (2009). Role of Iron and Hydroperoxides in the Degradation of Lycopene in Oil-in-Water Emulsions. Journal of Agricultural and Food Chemistry, 57(7), 2993-2998. Boon, C. S., McClements, D. J., Weiss, J., & Decker, E. A. (2010). Factors Influencing the Chemical Stability of Carotenoids in Foods. Critical Reviews in Food Science and Nutrition, 50(6), 515-532.

AC C

EP

TE D

M AN U

SC

RI PT

2194

ACCEPTED MANUSCRIPT 76

EP

TE D

M AN U

SC

RI PT

Bortolomeazzi, R., Cordaro, F., Pizzale, L., & Conte, L. S. (2003). Presence of phytosterol oxides in crude vegetable oils and their fate during refining. Journal of Agricultural and Food Chemistry, 51(8), 2394-2401. Bouchemal, K., Briancon, S., Perrier, E., & Fessi, H. (2004). Nano-emulsion formulation using spontaneous emulsification: solvent, oil and surfactant optimisation. International Journal of Pharmaceutics, 280(1-2), 241-251. Brisson, L., Castan, S., Fontbonne, H., Nicoletti, C., Puigserver, A., & Ajandouz, E. H. (2008). Alpha-tocopheryl acetate is absorbed and hydrolyzed by Caco-2 cells - Comparative studies with alpha-tocopherol. Chemistry and Physics of Lipids, 154(1), 33-37. Buffo, R. A., & Reineccius, G. A. (2001). Shelf-life and mechanisms of destabilization in dilute beverage emulsions. Flavour and Fragrance Journal, 16(1), 7-12. Buffo, R. A., Reineccius, G. A., & Oehlert, G. W. (2001). Factors affecting the emulsifying and rheological properties of gum acacia in beverage emulsions. Food Hydrocolloids, 15(1), 53-66. Buffo, R. A., Reineccius, G. A., & Oehlert, G. W. (2002). Influence of time-temperature treatments on the emulsifying properties of gum acacia in beverage emulsions. Journal of Food Engineering, 51(4), 341-345. Cao, Y., Dickinson, E., & Wedlock, D. J. (1990). Creaming and flocculation in emulsions containing polysaccharide. Food Hydrocolloids, 4(3), 185-195. Cercaci, L., Rodriguez-Estrada, M. T., Lercker, G., & Decker, E. A. (2007). Phytosterol oxidation in oil-in-water emulsions and bulk oil. Food Chemistry, 102(1), 161-167. Chaiyasit, W., Elias, R. J., McClements, D. J., & Decker, E. A. (2007). Role of physical structures in bulk oils on lipid oxidation. . Crit. Rev. Food Sci. Nutr., 47, 299-317. Chanamai, R., Horn, G., & McClements, D. J. (2002). Influence of oil polarity on droplet growth in oil-in-water emulsions stabilized by a weakly adsorbing biopolymer or a nonionic surfactant. Journal of Colloid and Interface Science, 247(1), 167-176. Chanamai, R., & McClements, D. J. (2000). Impact of weighting agents and sucrose on gravitational separation of beverage emulsions. Journal of Agricultural and Food Chemistry, 48(11), 5561-5565. Chanamai, R., & McClements, D. J. (2001). Depletion flocculation of beverage emulsions by gum arabic and modified starch. Journal of Food Science, 66(3), 457-463. Chanamai, R., & McClements, D. J. (2002a). Comparison of gum arabic, modified starch, and whey protein isolate as emulsifiers: Influence of pH, CaCl2 and temperature. J. Food Sci., 67(1), 120-125. Chanamai, R., & McClements, D. J. (2002b). Comparison of gum arabic, modified starch, and whey protein isolate as emulsifiers: Influence of pH, CaCl2 and temperature. Journal of Food Science, 67(1), 120-125. Charoen, R., Jangchud, A., Jangchud, K., Harnsilawat, T., Naivikul, O., & McClements, D. J. (2011). Influence of Biopolymer Emulsifier Type on Formation and Stability of Rice Bran Oil-in-Water Emulsions: Whey Protein, Gum Arabic, and Modified Starch. Journal of Food Science, 76(1), E165-E172. Chen, C. C., & Wagner, G. (2004). Vitamin E nanoparticle for beverage applications. Chemical Engineering Research & Design, 82(A11), 1432-1437. Cheng, Q., & McClements, D. J. (2011). Formation of nanoemulsions stabilized by model foodgrade emulsifiers using high-pressure homogenization: Factors affecting particle size. Food Hydrocolloids, In Press.

AC C

2231 2232 2233 2234 2235 2236 2237 2238 2239 2240 2241 2242 2243 2244 2245 2246 2247 2248 2249 2250 2251 2252 2253 2254 2255 2256 2257 2258 2259 2260 2261 2262 2263 2264 2265 2266 2267 2268 2269 2270 2271 2272 2273 2274 2275 2276

ACCEPTED MANUSCRIPT 77

EP

TE D

M AN U

SC

RI PT

Chiu, Y., & Yang, W. (1992). Preparation of vitamin E microemulsion possessing high resistance to oxidation in air. Colloids and Surfaces, 63(3-4), 311-322. Cho, Y. H., & McClements, D. J. (2009). Theoretical Stability Maps for Guiding Preparation of Emulsions Stabilized by Protein-Polysaccharide Interfacial Complexes. Langmuir, 25(12), 6649-6657. Choi, S. J., Decker, E. A., Henson, L., Popplewell, L. M., & McClements, D. J. (2009). Stability of Citral in Oil-in-Water Emulsions Prepared with Medium-Chain Triacylglycerols and Triacetin. Journal of Agricultural and Food Chemistry, 57(23), 11349-11353. Choi, S. J., Decker, E. A., Henson, L., Popplewell, L. M., & McClements, D. J. (2010a). Influence of Droplet Charge on the Chemical Stability of Citral in Oil-in-Water Emulsions. Journal of Food Science, 75(6), C536-C540. Choi, S. J., Decker, E. A., Henson, L., Popplewell, L. M., & McClements, D. J. (2010b). Inhibition of citral degradation in model beverage emulsions using micelles and reverse micelles. Food Chemistry, 122(1), 111-116. Choi, S. J., Decker, E. A., Henson, L., Popplewell, L. M., Xiao, H., & McClements, D. J. (2011). Formulation and properties of model beverage emulsions stabilized by sucrose monopalmitate: Influence of pH and lyso-lecithin addition. Food Research International, 44(9), 3006-3012. Chu, B. S., Ichikawa, S., Kanafusa, S., & Nakajima, M. (2007). Preparation and characterization of beta-carotene nanodispersions prepared by solvent displacement technique. Journal of Agricultural and Food Chemistry, 55(16), 6754-6760. Clark, B. C., Powell, C. C., & Radford, T. (1977). Acid-catalyzed cyclization of citral. Tetrahedron, 33(17), 2187-2191. Cortesi, R., Esposito, E., Luca, G., & Nastruzzi, C. (2002). Production of lipospheres as carriers for bioactive compounds. Biomaterials, 23(11), 2283-2294. Coupland, J. N., Weiss, J., Lovy, A., & McClements, D. J. (1996). Solubilization kinetics of triacyl glycerol and hydrocarbon emulsion droplets in a micellar solution. Journal of Food Science, 61(6), 1114-1117. Cui, S. W. (2005). Food Carbohydrates: Chemistry, Physical Properties and Applications. Boca Raton, FL: Taylor and Francis. Dalgleish, D. G., West, S. J., & Hallett, F. R. (1997). The characterization of small emulsion droplets made from milk proteins and triglyceride oil. Colloids and Surfaces aPhysicochemical and Engineering Aspects, 123, 145-153. Danviriyakul, S., McClements, D. J., Decker, E., Nawar, W. W., & Chinachoti, P. (2002). Physical stability of spray-dried milk fat emulsion as affected by emulsifiers and processing conditions. Journal of Food Science, 67(6), 2183-2189. Decker, E. A., & McClements, D. J. (2001). Transition metal and hydroperoxide interactions: an important determinant in the oxidative stability of lipid dispersions. Inform, 12, 215-255. Demetriades, K., Coupland, J. N., & McClements, D. J. (1997a). Physical properties of whey protein stabilized emulsions as related to pH and NaCl. Journal of Food Science, 62(2), 342-347. Demetriades, K., Coupland, J. N., & McClements, D. J. (1997b). Physicochemical properties of whey protein-stabilized emulsions as affected by heating and ionic strength. Journal of Food Science, 62(3), 462-467. Dickinson, E. (1992a). Introduction to Food Colloids. Cambridge, UK: Royal Society of Chemistry.

AC C

2277 2278 2279 2280 2281 2282 2283 2284 2285 2286 2287 2288 2289 2290 2291 2292 2293 2294 2295 2296 2297 2298 2299 2300 2301 2302 2303 2304 2305 2306 2307 2308 2309 2310 2311 2312 2313 2314 2315 2316 2317 2318 2319 2320 2321 2322

ACCEPTED MANUSCRIPT 78

EP

TE D

M AN U

SC

RI PT

Dickinson, E. (1992b). An Introduction to Food Colloids. Oxford: Oxford Science Publishers. Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the properties of dispersed systems. Food Hydrocolloids, 17(1), 25-39. Dickinson, E. (2010). Flocculation of protein-stabilized oil-in-water emulsions. Colloids and Surfaces B-Biointerfaces, 81(1), 130-140. Dickinson, E. (2011). Mixed biopolymers at interfaces: Competitive adsorption and multilayer structures. Food Hydrocolloids, 25(8), 1966-1983. Dickinson, E., & Stainsby, G. (1982). Colloids in Foods London: Applied Science. Djordjevic, D., Cercaci, L., Alamed, J., McClements, D. J., & Decker, E. A. (2007). Chemical and physical stability of citral and limonene in sodium dodecyl sulfate-chitosan and gum arabic-stabilized oil-in-water emulsions. Journal of Agricultural and Food Chemistry, 55(9), 3585-3591. Djordjevic, D., Cercaci, L., Alamed, J., McClements, D. J., & Decker, E. A. (2008). Stability of citral in protein- and gum arabic-stabilized oil-in-water emulsions. Food Chemistry, 106(2), 698-705. DuBois, G. E., & Prakash, I. (2012). Non-Caloric Sweeteners, Sweetness Modulators, and Sweetener Enhancers. In M. P. Doyle & T. R. Klaenhammer (Eds.), Annual Review of Food Science and Technology, Vol 3 (Vol. 3, pp. 353-380). Palo Alto: Annual Reviews. Dutta, P. C. (1997). Studies on phytosterol oxides .2. Content in some vegetable oils and in french fries prepared in these oils. Journal of the American Oil Chemists Society, 74(6), 659-666. Fanun, M. (2010). Microemulsions with Mixed Nonionic Surfactants and Flavor Oil. Journal of Surfactants and Detergents, 13(3), 321-328. Feng, J.-L., Wang, Z.-W., Zhang, J., Wang, Z.-N., & Liu, F. (2009a). Study on food-grade vitamin E microemulsions based on nonionic emulsifiers. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 339(1-3), 1-6. Feng, J. L., Wang, Z. W., Zhang, J., Wang, Z. N., & Liu, F. (2009b). Study on food-grade vitamin E microemulsions based on nonionic emulsifiers. Colloids and Surfaces aPhysicochemical and Engineering Aspects, 339(1-3), 1-6. Fernandez, P., Andre, V., Rieger, J., & Kuhnle, A. (2004). Nano-emulsion formation by emulsion phase inversion. Colloids and Surfaces a-Physicochemical and Engineering Aspects, 251(1-3), 53-58. Fernstrom, J. D., Munger, S. D., Sclafani, A., de Araujo, I. E., Roberts, A., & Molinary, S. (2012). Mechanisms for Sweetness. Journal of Nutrition, 142(6), 1134S-1141S. Frankel, E. N., Satué-Gracia, T., Meyer, A. S., & German, J. B. (2002). Oxidative Stability of Fish and Algae Oils Containing Long-Chain Polyunsaturated Fatty Acids in Bulk and in Oil-in-Water Emulsions. J Agric Food Chem, 50(7), 2094-2099. Freitas, S., Merkle, H. P., & Gander, B. (2005). Microencapsulation by solvent extraction/evaporation: reviewing the state of the art of microsphere preparation process technology. Journal of Controlled Release, 102(2), 313-332. Friberg, S., Larsson, K., & Sjoblom, J. (2004). Food Emulsions (4 ed.). New York: Marcel Dekker. Gamarra, F. M. C., Sakanaka, L. S., Tambourgi, E. B., & Cabral, F. A. (2006). Influence on the quality of essential lemon (Citrus aurantifolia) oil by distillation process. Brazilian Journal of Chemical Engineering, 23(1), 147-151.

AC C

2323 2324 2325 2326 2327 2328 2329 2330 2331 2332 2333 2334 2335 2336 2337 2338 2339 2340 2341 2342 2343 2344 2345 2346 2347 2348 2349 2350 2351 2352 2353 2354 2355 2356 2357 2358 2359 2360 2361 2362 2363 2364 2365 2366 2367

ACCEPTED MANUSCRIPT 79

EP

TE D

M AN U

SC

RI PT

Gao, Y. L., & Kispert, L. D. (2003). Reaction of carotenoids and ferric chloride: Equilibria, isomerization, and products. Journal of Physical Chemistry B, 107(22), 5333-5338. Garti, N. (1999). Hydrocolloids as emulsifying agents for oil-in-water emulsions. J. Disp. Sci. and Tech., 20(1-2), 327-355. Garti, N., Aserin, A., & Fanun, M. (2000). Non-ionic sucrose esters microemulsions for food applications. Part 1. Water solubilization. Colloids and Surfaces a-Physicochemical and Engineering Aspects, 164(1), 27-38. Garti, N., & Bisperink, C. (1998). Double emulsions: progress and applications. Current Opinion in Colloid & Interface Science, 3(6), 657-667. Garti, N., Clement, V., Fanun, M., & Leser, M. E. (2000). Some characteristics of sugar ester nonionic microemulsions in view of possible food applications. Journal of Agricultural and Food Chemistry, 48(9), 3945-3956. Garti, N., Clement, V., Leser, M., Aserin, A., & Fanun, M. (1999). Sucrose ester microemulsions. Journal of Molecular Liquids, 80(2-3), 253-296. Garti, N., Reichman, D., Hendrickx, H., Dickinson, E., Jackson, L. K., & Bergenstahl, B. (1993). Hydrocolloids as food emulsifiers and stabilizers. Food Structure, 12(4), 411-426. Gawrysiak-Witulska, M., Siger, A., & Nogala-Kalucka, M. (2009). Degradation of tocopherols during near-ambient rapeseed drying. Journal of Food Lipids, 16(4), 524-539. Genovese, D. B., Lozano, J. E., & Rao, M. A. (2007). The rheology of colloidal and noncolloidal food dispersions. Journal of Food Science, 72(2), R11-R20. Gibbs, B. F., Kermasha, S., Alli, I., & Mulligan, C. N. (1999). Encapsulation in the food industry: a review. International Journal of Food Sciences and Nutrition, 50(3), 213-224. Given Jr., P. S. (2009). Encapsulation of flavors in emulsions for beverages. Current Opinion in Colloid and Interface Science, 14, 43-47. Given, P. S. (2009). Encapsulation of Flavors in Emulsions for Beverages. Current Opinion in Colloid & Interface Science, 14(1), 43-47. Glatter, O., Orthaber, D., Stradner, A., Scherf, G., Fanun, M., Garti, N., Clément, V., & Leser, M. (2001). Sugar-ester nonionic microemulsion: Structural characterization. Journal of Colloid and Interface Science, 241(1), 215-225. Gonnet, M., Lethuaut, L., & Boury, F. (2010). New trends in encapsulation of liposoluble vitamins. Journal of Controlled Release, 146(3), 276-290. Gu, Y. S., Decker, E. A., & McClements, D. J. (2004). Influence of pH and iota-carrageenan concentration on physicochemical properties and stability of beta-lactoglobulin-stabilized oil-in-water emulsions. Journal of Agricultural and Food Chemistry, 52(11), 3626-3632. Gunning, P. A., Hibberd, D. J., Howe, A. M., & Robins, M. M. (1988). Gravitational destabilization of emulsions flocculated by non-adsorbed xanthan. Food Hydrocolloids, 2(2), 119-129. Gutierrez, J. M., Gonzalez, C., Maestro, A., Sole, I., Pey, C. M., & Nolla, J. (2008). Nanoemulsions: New applications and optimization of their preparation. Current Opinion in Colloid & Interface Science, 13(4), 245-251. Guzey, D., & McClements, D. J. (2006). Formation, stability and properties of multilayer emulsions for application in the food industry. Advances in Colloid and Interface Science, 128, 227-248. Guzey, D., & McClements, D. J. (2007). Impact of electrostatic interactions on formation and stability of emulsions containing oil droplets coated by beta-lactoglobulin-pectin complexes. Journal of Agricultural and Food Chemistry, 55(2), 475-485.

AC C

2368 2369 2370 2371 2372 2373 2374 2375 2376 2377 2378 2379 2380 2381 2382 2383 2384 2385 2386 2387 2388 2389 2390 2391 2392 2393 2394 2395 2396 2397 2398 2399 2400 2401 2402 2403 2404 2405 2406 2407 2408 2409 2410 2411 2412 2413

ACCEPTED MANUSCRIPT 80

EP

TE D

M AN U

SC

RI PT

Hallikainen, M. A., Sarkkinen, E. S., & Uusitupa, M. I. J. (2000). Plant stanol eaters affect serum cholesterol concentrations of hypercholesterolemic men and women in a dose-dependent manner. Journal of Nutrition, 130(4), 767-776. Harnsilawat, T., Pongsawatmanit, R., & McClements, D. J. (2006). Stabilization of model beverage cloud emulsions using protein-polysaccharide electrostatic complexes formed at the oil-water interface. Journal of Agricultural and Food Chemistry, 54(15), 5540-5547. Hatanaka, J., Chikamori, H., Sato, H., Uchida, S., Debari, K., Onoue, S., & Yamada, S. (2010). Physicochemical and pharmacological characterization of α-tocopherol-loaded nanoemulsion system. International Journal of Pharmaceutics, 396(1-2), 188-193. Heinonen, M., Haila, K., Lampi, A. M., & Piironen, V. (1997). Inhibition of oxidation in 10% oil-in-water emulsions by beta-carotene with alpha- and gamma-tocopherols. Journal of the American Oil Chemists Society, 74(9), 1047-1052. Henry, J. V. L., Fryer, P. J., Frith, W. J., & Norton, I. T. (2010). The influence of phospholipids and food proteins on the size and stability of model sub-micron emulsions. Food Hydrocolloids, 24(1), 66-71. Horn, D., & Rieger, J. (2001). Organic nanoparticles in the aqueous phase - theory, experiment, and use. Angewandte Chemie-International Edition, 40(23), 4331-4361. Hostettmann, K., & Marston, A. (1995). Saponins (First ed.). Cambridge, UK: Cambridge University Press. Huck-Iriart, C., Candal, R., & Herrera, M. (2009). Effects of Addition of a Palmitic Sucrose Ester on Low-Trans-Fat Blends Crystallization in Bulk and in Oil-in-Water Emulsions. Food Biophysics, 4(3), 158-166. Hunter, R. J. (1986). Foundations of Colloid Science (Vol. 1). Oxford: Oxford University Press. Imeson, A. (2010). Food Stabilizers, Thickeners and Gelling Agents. Chichester, UK: John Wiley & Sons. Institute.of.Medicine. (2000). Dietary Reference Intakes for Vitamin C, Vitamin E, Selenium, and Carotenoids. In. Washington, D.C.: The National Academies Press. Israelachvili, J. (2011). Intermolecular and Surface Forces, Third Edition (Third Edition ed.). London, UK: Academic Press. Jafari, S. M., Assadpoor, E., He, Y. H., & Bhandari, B. (2008). Re-coalescence of emulsion droplets during high-energy emulsification. Food Hydrocolloids, 22(7), 1191-1202. Jafari, S. M., He, Y. H., & Bhandari, B. (2006). Nano-emulsion production by sonication and microfluidization - A comparison. International Journal of Food Properties, 9(3), 475485. Jahanzad, F., Crombie, G., Innes, R., & Sajjadi, S. (2009). Catastrophic phase inversion via formation of multiple emulsions: A prerequisite for formation of fine emulsions. Chemical Engineering Research & Design, 87(4A), 492-498. Jayme, M. L., Dunstan, D. E., & Gee, M. L. (1999). Zeta potentials of gum arabic stabilised oil in water emulsions. Food Hydrocolloids, 13(6), 459-465. Jung, J. Y., & Wicker, L. (2012). Laccase mediated conjugation of sugar beet pectin and the effect on emulsion stability. Food Hydrocolloids, 28(1), 168-173. Kabalnov, A. (2001). Ostwald ripening and related phenomena. Journal of Dispersion Science and Technology, 22(1), 1-12. Kabalnov, A., & Wennerstrom, H. (1996). Macroemulsion stability: The oriented wedge theory revisited. Langmuir, 12(2), 276-292.

AC C

2414 2415 2416 2417 2418 2419 2420 2421 2422 2423 2424 2425 2426 2427 2428 2429 2430 2431 2432 2433 2434 2435 2436 2437 2438 2439 2440 2441 2442 2443 2444 2445 2446 2447 2448 2449 2450 2451 2452 2453 2454 2455 2456 2457 2458

ACCEPTED MANUSCRIPT 81

EP

TE D

M AN U

SC

RI PT

Kabalnov, A. S., Pertzov, A. V., & Shchukin, E. D. (1987). Ostwald Ripening in 2-Component Disperse Phase Systems - Application to Emulsion Stability. Colloids and Surfaces, 24(1), 19-32. Kabalnov, A. S., & Shchukin, E. D. (1992). Ostwald Ripening Theory - Applications to Fluorocarbon Emulsion Stability. Advances in Colloid and Interface Science, 38, 69-97. Kandansamy, K., & Somasundaram, P. D. (2012). Microencapsulation of Colors by Spray Drying - A Review. International Journal of Food Engineering, 8(2). Kentish, S., Wooster, T. J., Ashokkumar, A., Balachandran, S., Mawson, R., & Simons, L. (2006). The use of ultrasonics for nanoemulsion preparation. In 3rd Innovative Foods Centre Conference on Food Innovation: Emerging Science, Technologies and Applications (pp. 170-175). Melbourne, AUSTRALIA. Kim, H. J., Decker, E. A., & McClements, D. J. (2002). Role of postadsorption conformation changes of beta-lactoglobulin on its ability to stabilize oil droplets against flocculation during heating at neutral pH. Langmuir, 18(20), 7577-7583. Kim, Y. D., Morr, C. V., & Schenz, T. W. (1996). Microencapsulation properties of gum Arabic and several food proteins: Liquid orange oil emulsion particles. J. Agric. Food Chem., 44(5), 1308-1313. Kimura, K., Iwata, I., & Nishimura, H. (1982). Studies on the deterioration mechanism of lemon flavor .1. Relationship between acid-catalyzed cyclization of citral and deterioration of lemon flavor. Agricultural and Biological Chemistry, 46(5), 1387-1389. Kimura, K., Nishimura, H., Iwata, I., & Mizutani, J. (1983a). Deterioration mechanism of lemon flavor .2. Formation mechanism of off-odor substances arising from citral. Journal of Agricultural and Food Chemistry, 31(4), 801-804. Kimura, K., Nishimura, H., Iwata, I., & Mizutani, J. (1983b). Studies on the deterioration mechanism of lemon flavor .3. Identification of acidic substances from deteriorated citral and effects of anti-oxidant on their formation. Agricultural and Biological Chemistry, 47(7), 1661-1663. Klein, M., Aserin, A., Svitov, I., & Garti, N. (2010). Enhanced stabilization of cloudy emulsions with gum Arabic and whey protein isolate. Colloids and Surfaces B-Biointerfaces, 77(1), 75-81. Konovalov, V. V., & Kispert, L. D. (1999). AM1, INDO/S and optical studies of carbocations of carotenoid molecules. Acid induced isomerization. Journal of the Chemical SocietyPerkin Transactions 2(4), 901-909. Kralova, I., & Sjoblom, J. (2009). Surfactants Used in Food Industry: A Review. Journal of Dispersion Science and Technology, 30(9), 1363-1383. Krinsky, N. I. (1998). The antioxidant and biological properties of the carotenoids. Towards Prolongation of the Healthy Life Span, 854, 443-447. Lambelet, P., Grandgirard, A., Gregoire, S., Juaneda, P., Sebedio, J. L., & Bertoli, C. (2003). Formation of modified fatty acids and oxyphytosterols during refining of low erucic acid rapeseed oil. Journal of Agricultural and Food Chemistry, 51(15), 4284-4290. Leong, T. S. H., Wooster, T. J., Kentish, S. E., & Ashokkumar, M. (2009). Minimising oil droplet size using ultrasonic emulsification. Ultrasonics Sonochemistry, 16(6), 721-727. Li, X., Anton, N., Thi, M. C. T., Zhao, M. J., Messaddeq, N., & Vandamme, T. F. (2011). Microencapsulation of nanoemulsions: novel Trojan particles for bioactive lipid molecule delivery. International Journal of Nanomedicine, 6, 1313-1325.

AC C

2459 2460 2461 2462 2463 2464 2465 2466 2467 2468 2469 2470 2471 2472 2473 2474 2475 2476 2477 2478 2479 2480 2481 2482 2483 2484 2485 2486 2487 2488 2489 2490 2491 2492 2493 2494 2495 2496 2497 2498 2499 2500 2501 2502 2503

ACCEPTED MANUSCRIPT 82

EP

TE D

M AN U

SC

RI PT

Li, Y., Le Maux, S., Xiao, H., & McClements, D. J. (2009). Emulsion-Based Delivery Systems for Tributyrin, a Potential Colon Cancer Preventative Agent. Journal of Agricultural and Food Chemistry, 57(19), 9243-9249. Liang, C. P., Wang, M. F., Simon, J. E., & Ho, C. T. (2004a). Antioxidant activity of plant extracts on the inhibition of citral off-odor formation. Molecular Nutrition & Food Research, 48(4), 308-317. Liang, C. P., Wang, M. F., Simon, J. E., & Ho, C. T. (2004b). Stabilization of citral flavor and fish oil fatty acids by phenolic extracts. Abstracts of Papers of the American Chemical Society, 228, 012-AGFD. Liebler, D. C., & McClure, T. D. (1996). Antioxidant reactions of beta-carotene: Identification of carotenoid-radical adducts. Chemical Research in Toxicology, 9(1), 8-11. Lim, S. S., Baik, M. Y., Decker, E. A., Henson, L., Popplewell, L. M., McClements, D. J., & Choi, S. J. (2011). Stabilization of orange oil-in-water emulsions: A new role for ester gum as an Ostwald ripening inhibitor. Food Chemistry, 128(4), 1023-1028. Lin, C. Y., & Chen, L. W. (2008). Comparison of fuel properties and emission characteristics of two- and three-phase emulsions prepared by ultrasonically vibrating and mechanically homogenizing emulsification methods. Fuel, 87(10-11), 2154-2161. Maa, Y. F., & Hsu, C. C. (1999). Performance of Sonication and microfluidization for liquidliquid emulsification. Pharmaceutical Development and Technology, 4(2), 233-240. Mader, I. (1964). Beta-Carotene - Thermal Degradation. Science, 144(361), 533-&. Maestro, A., Sole, I., Gonzalez, C., Solans, C., & Gutierrez, J. M. (2008). Influence of the phase behavior on the properties of ionic nanoemulsions prepared by the phase inversion composition method. Journal of Colloid and Interface Science, 327(2), 433-439. Mayne, S. T. (1996). Beta-carotene, carotenoids, and disease prevention in humans. The FASEB Journal, 10(7), 690-701. McCann, D., Barrett, A., Cooper, A., Crumpler, D., Dalen, L., Grimshaw, K., Kitchin, E., Lok, K., Porteous, L., Prince, E., Sonuga-Barke, E., Warner, J. O., & Stevenson, J. (2007). Food additives and hyperactive behaviour in 3-year-old and 8/9-year-old children in the community: a randomised, double-blinded, placebo-controlled trial. Lancet, 370(9598), 1560-1567. McCaughey, S. A. (2008). The taste of sugars. Neuroscience and Biobehavioral Reviews, 32(5), 1024-1043. McClements, D. (2005a). Food emulsions: principles, practices, and techniques: CRC. McClements, D. J. (2002). Theoretical prediction of emulsion color. Advances in Colloid and Interface Science, 97(1-3), 63-89. McClements, D. J. (2004). Protein-stabilized emulsions. Current Opinion in Colloid & Interface Science, 9(5), 305-313. McClements, D. J. (2005b). Food Emulsions: Principles, Practice, and Techniques (2nd ed.). Boca Raton: CRC Press. McClements, D. J. (2005c). Food Emulsions: Principles, Practices, and Techniques. In (2nd ed.). Boca Raton, Fl: CRC Press. McClements, D. J. (2007). Critical review of techniques and methodologies for characterization of emulsion stability. Critical Reviews in Food Science and Nutrition, 47(7), 611-649. McClements, D. J. (2011). Edible nanoemulsions: fabrication, properties, and functional performance. Soft Matter, 7(6), 2297-2316.

AC C

2504 2505 2506 2507 2508 2509 2510 2511 2512 2513 2514 2515 2516 2517 2518 2519 2520 2521 2522 2523 2524 2525 2526 2527 2528 2529 2530 2531 2532 2533 2534 2535 2536 2537 2538 2539 2540 2541 2542 2543 2544 2545 2546 2547 2548

ACCEPTED MANUSCRIPT 83

EP

TE D

M AN U

SC

RI PT

McClements, D. J. (2012). Nanoemulsions versus Microemulsions: Clarification of Differences, Similarities and Terminology. Soft Matter, 8 (6), 1719 - 1729. McClements, D. J., & Decker, E. A. (2000). Lipid oxidation in oil-in-water emulsions: Impact of molecular environment on chemical reactions in heterogeneous food systems. Journal of Food Science, 65(8), 1270-1282. McClements, D. J., & Decker, E. A. (2008). Lipids. In S. Damodaran, K. L. Parkin & F. O.R. (Eds.), Food Chemistry, 4th Edition (pp. 155-216). Boca Raton: CRC Press. McClements, D. J., Decker, E. A., Park, Y., & Weiss, J. (2009). Structural Design Principles for Delivery of Bioactive Components in Nutraceuticals and Functional Foods. Critical Reviews in Food Science and Nutrition, 49(6), 577-606. McClements, D. J., Decker, E. A., & Weiss, J. (2007). Emulsion-based delivery systems for lipophilioc bioactive components. Journal of Food Science, 72(8), R109-R124. McClements, D. J., Henson, L., Popplewell, L. M., Decker, E. A., & Choi, S. J. (2012). Inhibition of Ostwald Ripening in Model Beverage Emulsions by Addition of Poorly Water Soluble Triglyceride Oils. Journal of Food Science, 77(1), C33-C38. McClements, D. J., & Li, Y. (2010). Structured emulsion-based delivery systems: Controlling the digestion and release of lipophilic food components. Advances in Colloid and Interface Science, 159(2), 213-228. McClements, D. J., & Rao, J. (2011). Food-Grade Nanoemulsions: Formulation, Fabrication, Properties, Performance, Biological Fate, and Potential Toxicity. Critical Reviews in Food Science and Nutrition, 51(4), 285-330. Mei, L. Y., Choi, S. J., Alamed, J., Henson, L., Popplewell, M., McClements, D. J., & Decker, E. A. (2010). Citral Stability in Oil-in-Water Emulsions with Solid or Liquid Octadecane. Journal of Agricultural and Food Chemistry, 58(1), 533-536. Meyers, B., & Brewer, M. S. (2008). Sweet taste in man: A review. Journal of Food Science, 73(6), R81-R90. Miller, C. A. (1988). Spontaneous Emulsification Produced by Diffusion - a Review. Colloids and Surfaces, 29(1), 89-102. Mirhosseini, H., Tan, C. P., Aghlara, A., Hamid, N. S. A., Yusof, S., & Chern, B. H. (2008). Influence of pectin and CMC on physical stability, turbidity loss rate, cloudiness and flavor release of orange beverage emulsion during storage. Carbohydrate Polymers, 73(1), 83-91. Mirhosseini, H., Tan, C. P., Hamid, N. S. A., & Yusof, S. (2008). Effect of Arabic gum, xanthan gum and orange oil on flavor release from diluted orange beverage emulsion. Food Chemistry, 107(3), 1161-1172. Misharina, T. A., Terenina, M. B., Krikunova, N. I., & Medvedeva, I. B. (2010). Autooxidation of a Mixture of Lemon Essential Oils, Methyl Linolenoate, and Methyl Oleinate. Applied Biochemistry and Microbiology, 46(5), 551-556. Mitra, S., & Dungan, S. R. (1997). Micellar properties of quillaja saponin .1. Effects of temperature, salt, and pH on solution properties. Journal of Agricultural and Food Chemistry, 45(5), 1587-1595. Mortensen, A., & Skibsted, L. H. (1996). Kinetics of photobleaching of beta-carotene in chloroform and formation of transient carotenoid species absorbing in the near infrared. Free Radical Research, 25(4), 355-368. Nahon, D. F., Roozen, J. P., & DeGraaf, C. (1996). Sweetness flavour interactions in soft drinks. Food Chemistry, 56(3), 283-289.

AC C

2549 2550 2551 2552 2553 2554 2555 2556 2557 2558 2559 2560 2561 2562 2563 2564 2565 2566 2567 2568 2569 2570 2571 2572 2573 2574 2575 2576 2577 2578 2579 2580 2581 2582 2583 2584 2585 2586 2587 2588 2589 2590 2591 2592 2593 2594

ACCEPTED MANUSCRIPT 84

EP

TE D

M AN U

SC

RI PT

Nakamura, A., Takahashi, T., Yoshida, R., Maeda, H., & Corredig, M. (2004). Emulsifying properties of soybean soluble polysaccharide. Food Hydrocolloids, 18(5), 795-803. Nakamura, A., Yoshida, R., Maeda, H., Furuta, H., & Corredig, M. (2004). Study of the role of the carbohydrate and protein moieties of soy soluble polysaccharides in their emulsifying properties. Journal of Agricultural and Food Chemistry, 52(17), 5506-5512. Nakauma, M., Funami, T., Noda, S., Ishihara, S., Al-Assaf, S., Nishinari, K., & Phillips, G. O. (2008). Comparison of sugar beet pectin, soybean soluble polysaccharide, and gum arabic as food emulsifiers. 1. Effect of concentration, pH, and salts on the emulsifying properties. Food Hydrocolloids, 22(7), 1254-1267. Ogawa, S., Decker, E. A., & McClements, D. J. (2003). Production and characterization of O/W emulsions containing cationic droplets stabilized by lecithin-chitosan membranes. Journal of Agricultural and Food Chemistry, 51(9), 2806-2812. Orchard, T. S., Pan, X. L., Cheek, F., Ing, S. W., & Jackson, R. D. (2012). A systematic review of omega-3 fatty acids and osteoporosis. British Journal of Nutrition, 107, S253-S260. Ostertag, F., Weiss, J., & McClements, D. J. (2012). Low-energy formation of edible nanoemulsions: Factors influencing droplet size produced by emulsion phase inversion. Journal of Colloid and Interface Science, 388, 95-102. Ostlund, R. E. (2004). Phytosterols and cholesterol metabolism. Current Opinion in Lipidology, 15(1), 37-41. Phillips, G. O., & Williams, P. A. (2000). Gum arabic. In Handbook of Hydrocolloids (pp. 165): CRC. Phillips, G. O., & Williams, P. A. (2009). Handbook of Hydrocolloids (Second ed.). Boca Raton, FL: CRC Press. Pouton, C. W., & Porter, C. J. H. (2006). Formulation of lipid-based delivery systems for oral administration: Materials, methods and strategies. In Annual Meeting of the AmericanAssociation-of-Pharmaceutical-Scientists (pp. 625-637). San Antonio, TX. Qian, C., Decker, E. A., Xiao, H., & McClements, D. J. (2011). Comparison of Biopolymer Emulsifier Performance in Formation and Stabilization of Orange Oil-in-Water Emulsions. Journal of the American Oil Chemists Society, 88(1), 47-55. Qian, C., Decker, E. A., Xiao, H., & McClements, D. J. (2012). Physical and chemical stability of beta-carotene-enriched nanoemulsions: Influence of pH, ionic strength, temperature, and emulsifier type. Food Chemistry, 132(3), 1221-1229. Qian, C., & McClements, D. J. (2011). Formation of nanoemulsions stabilized by model foodgrade emulsifiers using high-pressure homogenization: Factors affecting particle size. Food Hydrocolloids, 25(5), 1000-1008. Quemada, D., & Berli, C. (2002). Energy of interaction in colloids and its implications in rheological modeling. Advances in Colloid and Interface Science, 98(1), 51-85. Rao, J., & McClements, D. J. (2012a). Food-grade microemulsions and nanoemulsions: Role of oil phase composition on formation and stability. Food Hydrocolloids, 29(2), 326-334. Rao, J., & McClements, D. J. (2012b). Impact of lemon oil composition on formation and stability of model food and beverage emulsions. Food Chemistry, In press. Rao, J. J., & McClements, D. J. (2010). Stabilization of Phase Inversion Temperature Nanoemulsions by Surfactant Displacement. Journal of Agricultural and Food Chemistry, 58(11), 7059-7066.

AC C

2595 2596 2597 2598 2599 2600 2601 2602 2603 2604 2605 2606 2607 2608 2609 2610 2611 2612 2613 2614 2615 2616 2617 2618 2619 2620 2621 2622 2623 2624 2625 2626 2627 2628 2629 2630 2631 2632 2633 2634 2635 2636 2637 2638

ACCEPTED MANUSCRIPT 85

EP

TE D

M AN U

SC

RI PT

Rao, J. J., & McClements, D. J. (2011). Food-Grade Microemulsions, Nanoemulsions and Emulsions: Fabrication from Sucrose Monopalmitate & Lemon Oil Food Hydrocolloids, 25(6), 1413-1423. Rao, J. J., & McClements, D. J. (2012c). Lemon oil solubilization in mixed surfactant solutions: Rationalizing microemulsion & nanoemulsion formation. Food Hydrocolloids, 26(1), 268-276. Reiner, S. J., Reineccius, G. A., & Peppard, T. L. (2010). A Comparison of the Stability of Beverage Cloud Emulsions Formulated with Different Gum Acacia- and Starch-Based Emulsifiers. Journal of Food Science, 75(5), E236-E246. Ribeiro, H. S., Ax, K., & Schubert, H. (2003). Stability of lycopene emulsions in food systems. Journal of Food Science, 68(9), 2730-2734. Ruxton, C. H. S., Reed, S. C., Simpson, M. J. A., & Millington, K. J. (2004). The health benefits of omega-3 polyunsaturated fatty acids: a review of the evidence. Journal of Human Nutrition and Dietetics, 17(5), 449-459. Ruxton, C. H. S., Reed, S. C., Simpson, M. J. A., & Millington, K. J. (2007). The health benefits of omega-3 polyunsaturated fatty acids: a review of the evidence. Journal of Human Nutrition and Dietetics, 20(3), 275-285. Ryan, L., O'Connell, O., O'Sullivan, L., Aherne, S. A., & O'Brien, N. M. (2008). Micellarisation of carotenoids from raw and cooked vegetables. Plant Foods for Human Nutrition (Formerly Qualitas Plantarum), 63(3), 127-133. Sadtler, V., Guely, M., Marchal, P., & Choplin, L. (2004). Shear-induced phase transitions in sucrose ester surfactant. Journal of Colloid and Interface Science, 270(2), 270-275. Sagalowicz, L., & Leser, M. E. (2010). Delivery systems for liquid food products. Current Opinion in Colloid & Interface Science, 15(1-2), 61-72. Sajjadi, S. (2006). Nanoemulsion formation by phase inversion emulsification: On the nature of inversion. Langmuir, 22(13), 5597-5603. Schieber, A., & Carle, R. (2005). Occurrence of carotenoid cis-isomers in food: Technological, analytical, and nutritional implications. Trends in Food Science & Technology, 16(9), 416-422. Shukat, R., Bourgaux, C., & Relkin, P. (2012). Crystallisation behaviour of palm oil nanoemulsions carrying vitamin E. Journal of Thermal Analysis and Calorimetry, 108(1), 153-161. Siddiqui, R. A., Shaikh, Sr., Sech, L. A., Yount, H. R., Stillwell, W., & Zaloga, G. P. (2004). Omega 3-fatty acids: Health benefits and cellular mechanisms of action. Mini-Reviews in Medicinal Chemistry, 4(8), 859-871. Sidhu, G., & Oakenfull, D. (1986). A mechanism for the hypocholesterolaemic activity of saponins. British Journal of Nutrition, 55(3), 643-649. Slavin, J. (2012). Beverages and body weight: challenges in the evidence-based review process of the Carbohydrate Subcommittee from the 2010 Dietary Guidelines Advisory Committee. Nutrition Reviews, 70, S111-S120. Sonneville-Aubrun, O., Babayan, D., Bordeaux, D., Lindner, P., Rata, G., & Cabane, B. (2009). Phase transition pathways for the production of 100 nm oil-in-water emulsions. Physical Chemistry Chemical Physics, 11(1), 101-110. Soupas, L., Juntunen, L., Lampi, A. M., & Piironen, V. (2004). Effects of sterol structure, temperature, and lipid medium on phytosterol oxidation. Journal of Agricultural and Food Chemistry, 52(21), 6485-6491.

AC C

2639 2640 2641 2642 2643 2644 2645 2646 2647 2648 2649 2650 2651 2652 2653 2654 2655 2656 2657 2658 2659 2660 2661 2662 2663 2664 2665 2666 2667 2668 2669 2670 2671 2672 2673 2674 2675 2676 2677 2678 2679 2680 2681 2682 2683 2684

ACCEPTED MANUSCRIPT 86

EP

TE D

M AN U

SC

RI PT

Stauffer, S. (1999). Emulsifiers. St Paul, MN: Eagen Press. Stringham, J. M., & Hammond, B. R. (2005). Dietary lutein and zeaxanthin: Possible effects on visual function. Nutrition Reviews, 63(2), 59-64. Sylvester, P. W., Wali, V. B., Bachawal, S. V., Shirode, A. B., Ayoub, N. M., & Akl, M. R. (2011). Tocotrienol combination therapy results in synergistic anticancer response. Frontiers in Bioscience-Landmark, 16, 3183-3195. Szuts, A., Budai-Szucs, M., Eros, I., Otomo, N., & Szabó-Révész, P. (2010). Study of gelforming properties of sucrose esters for thermosensitive drug delivery systems. International Journal of Pharmaceutics, 383(1-2), 132-137. Tadros, T., Izquierdo, R., Esquena, J., & Solans, C. (2004). Formation and stability of nanoemulsions. Advances in Colloid and Interface Science, 108-09, 303-318. Tan, C. (2004). Beverage emulsions. In F. S, L. K & S. J (Eds.), Food Emulsions (4th ed.). New York: Marcel Decker. Tan, C. T. (1998). Beverage flavor emulsion-a form of emulsion liquid membrane encapsulation. In E. T. Contis (Ed.), Food Flavors: Formation, Analysis and Packaging Influences (pp. 29). New York: Elsevier. Thakur, R. K., Villette, C., Aubry, J. M., & Delaplace, G. (2008). Dynamic emulsification and catastrophic phase inversion of lecithin-based emulsions. Colloids and Surfaces aPhysicochemical and Engineering Aspects, 315(1-3), 285-293. Traber, M. G., Frei, B., & Beckman, J. S. (2008). Vitamin E revisited: do new data validate benefits for chronic disease prevention? Current Opinion in Lipidology, 19(1), 30-38. Trubiano, P. C. (1995a). The role of specialty food starches in flavor emulsions. Flavor Technology, 610, 199-209. Trubiano, P. C. (1995b). The role of specialty food starches in flavor encapsulation. Flavor Technology, 610, 244-253. Tse, K. Y., & Reineccius, G. A. (1995a). Methods to predict the physical stability of flavor Cloud emulsion. In Flavor Technology (Vol. 610, pp. 172-182). Tse, K. Y., & Reineccius, G. A. (1995b). Methods to predict the physical stability of flavor Cloud emulsion. Flavor Technology, 610, 172-182. Tur, J. A., Bibiloni, M. M., Sureda, A., & Pons, A. (2012). Dietary sources of omega 3 fatty acids: public health risks and benefits. British Journal of Nutrition, 107, S23-S52. Ueno, T., Kiyohapa, S., Ho, C. T., & Masuda, H. (2006). Potent inhibitory effects of black tea theaflavins on off-odor formation from citral. Journal of Agricultural and Food Chemistry, 54(8), 3055-3061. Ueno, T., Masuda, H., & Ho, C.-T. (2004). Formation mechanism of p-methylacetophenone from citral via a tert-alkoxy radical intermediate. Journal of Agricultural and Food Chemistry, 52, 5677-5684. Ueno, T., Masuda, H., Muranishi, S., Kiyohara, S., Sekiguchi, Y., & Ho, C. T. (2003). Inhibition of the formation of off-odour compounds from citral in an acidic aqueous solution. In J. L. LeQuere & P. X. Etievant (Eds.), Flavour Research at the Dawn of the Twenty-First Century (pp. 128-131). van der Graaf, S., Schroen, C. G. P. H., & Boom, R. M. (2005). Preparation of double emulsions by membrane emulsification - a review. Journal of Membrane Science, 251(1-2), 7-15. van Setten, D. C., ten Hove, G. J., Wiertz, E. J. H. J., Kamerling, J. P., & van de Werken, G. (1998). Multiple-stage tandem mass spectrometry for structural characterization of saponins. Analytical chemistry, 70(20), 4401-4409.

AC C

2685 2686 2687 2688 2689 2690 2691 2692 2693 2694 2695 2696 2697 2698 2699 2700 2701 2702 2703 2704 2705 2706 2707 2708 2709 2710 2711 2712 2713 2714 2715 2716 2717 2718 2719 2720 2721 2722 2723 2724 2725 2726 2727 2728 2729 2730

ACCEPTED MANUSCRIPT 87

EP

TE D

M AN U

SC

RI PT

van Setten, D. C., van de Werken, G., Zomer, G., & Kersten, G. F. A. (1995). Glycosyl compositions and structural characteristics of the potential immuno-adjuvant active saponins in the Quillaja saponaria Molina extract Quil A. Rapid communications in mass spectrometry, 9(8), 660-666. Velikov, K. P., & Pelan, E. (2008). Colloidal delivery systems for micronutrients and nutraceuticals. Soft Matter, 4(10), 1964-1980. Waller, G. R., & Yamasaki, K. (1996a). Saponins used in food and agriculture (Vol. 405). New York, NY: Plenum Press. Waller, G. R., & Yamasaki, K. (1996b). Saponins used in traditional and modern medicine (Vol. 404). New York, NY: Plenum Press. Walstra, P. (1993). Principles of emulsion formation. Chemical Engineering Science, 48, 333. Walstra, P. (2003). Physical Chemistry of Foods. New York, NY.: Marcel Decker. Waraho, T., McClements, D. J., & Decker, E. A. (2011). Mechanisms of lipid oxidation in food dispersions. Trends in Food Science & Technology, 22(1), 3-13. Weiss, J., Herrmann, N., & McClements, D. J. (1999). Ostwald ripening of hydrocarbon emulsion droplets in surfactant solutions. Langmuir, 15(20), 6652-6657. Weng-Yew, W., & Brown, L. (2011). Nutrapharmacology of Tocotrienols for Metabolic Syndrome. Current Pharmaceutical Design, 17(21), 2206-2214. Williams, M. A. K., Fabri, D., Hubbard, C. D., Lundin, L., Foster, T. J., Clark, A. H., Norton, I. T., Loren, N., & Hermansson, A. M. (2001). Kinetics of droplet growth in gelatin/maltodextrin mixtures following thermal quenching. Langmuir, 17(11), 34123418. Wong, N. C. W. (2001). The beneficial effects of plant sterols on serum cholesterol. Canadian Journal of Cardiology, 17(6), 715-721. Woodall, A. A., Lee, S. W. M., Weesie, R. J., Jackson, M. J., & Britton, G. (1997). Oxidation of carotenoids by free radicals: relationship between structure and reactivity. Biochimica Et Biophysica Acta-General Subjects, 1336(1), 33-42. Wooster, T., Golding, M., & Sanguansri, P. (2008a). Impact of oil type on nanoemulsion formation and Ostwald ripening stability. Langmuir, 24(22), 12758-12765. Wooster, T. J., Golding, M., & Sanguansri, P. (2008b). Impact of Oil Type on Nanoemulsion Formation and Ostwald Ripening Stability. Langmuir, 24(22), 12758-12765. Wrolstad, R. E., & Culver, C. A. (2012). Alternatives to Those Artificial FD&C Food Colorants. In M. P. Doyle & T. R. Klaenhammer (Eds.), Annual Review of Food Science and Technology, Vol 3 (Vol. 3, pp. 59-77). Xianquan, S., Shi, J., Kakuda, Y., & Yueming, J. (2005). Stability of lycopene during food processing and storage. Journal of Medicinal Food, 8(4), 413-422. Yadav, M. P., Johnston, D. B., Hotchkiss, A. T., & Hicks, K. B. (2007). Corn fiber gum: A potential gum arabic replacer for beverage flavor emulsification. Food Hydrocolloids, 21(7), 1022-1030. Yamauchi, R., Miyake, N., Inoue, H., & Kato, K. (1993). Products Formed by Peroxyl Radical Oxidation of Beta-Carotene. Journal of Agricultural and Food Chemistry, 41(5), 708713. Yang, X. Q., Tian, H. X., Ho, C. T., & Huang, Q. R. (2011). Inhibition of Citral Degradation by Oil-in-Water Nanoemulsions Combined with Antioxidants. Journal of Agricultural and Food Chemistry, 59(11), 6113-6119.

AC C

2731 2732 2733 2734 2735 2736 2737 2738 2739 2740 2741 2742 2743 2744 2745 2746 2747 2748 2749 2750 2751 2752 2753 2754 2755 2756 2757 2758 2759 2760 2761 2762 2763 2764 2765 2766 2767 2768 2769 2770 2771 2772 2773 2774 2775

ACCEPTED MANUSCRIPT 88

EP

TE D

M AN U

SC

RI PT

Yang, X. Q., Tian, H. X., Ho, C. T., & Huang, Q. R. (2012). Stability of Citral in Emulsions Coated with Cationic Biopolymer Layers. Journal of Agricultural and Food Chemistry, 60(1), 402-409. Yang, Y., Leser, M. E., Sher, A. A., & McClements, D. J. (2013). Formation and stability of emulsions using a natural small molecule surfactant: Quillaja saponin (Q-Naturale (R)). Food Hydrocolloids, 30(2), 589-596. Yang, Y., Marshall-Breton, C., Leser, M. E., Sher, A. A., & McClements, D. J. (2012). Fabrication of ultrafine edible emulsions: Comparison of high-energy and low-energy homogenization methods. Food Hydrocolloids, 29(2), 398-406. Yang, Y., & McClements, D. J. (2013). Encapsulation of vitamin E in edible emulsions fabricated using a natural surfactant. Food Hydrocolloids, 30(2), 712-720. Yin, L. J., Chu, B. S., Kobayashi, I., & Nakajima, M. (2009). Performance of selected emulsifiers and their combinations in the preparation of beta-carotene nanodispersions. Food Hydrocolloids, 23, 1617-1622. Yoon, Y., & Choe, E. (2009). Lipid Oxidation and Stability of Tocopherols and Phospholipids in Soy-added Fried Products During Storage in the Dark. Food Science and Biotechnology, 18(2), 356-361. Zuccari, G., Carosio, R., Fini, A., Montaldo, P., & Orienti, I. (2005). Modified polyvinylalcohol for encapsulation of all-trans-retinoic acid in polymeric micelles. Journal of Controlled Release, 103(2), 369-380.

AC C

2776 2777 2778 2779 2780 2781 2782 2783 2784 2785 2786 2787 2788 2789 2790 2791 2792 2793 2794 2795 2796 2797

ACCEPTED MANUSCRIPT

RI PT

Figure 1: Schematic representation of different kinds of colloidal dispersions that can be used in the beverage industry: emulsions, nanoemulsions, and microemulsions.

SC

• Thermodynamically unstable • d > 100 nm • Optically opaque

M AN U

Emulsion

TE D

• Thermodynamically unstable • d < 100 nm • Transparent or slightly turbid

Microemulsion

AC C

EP

Nanoemulsion

• Thermodynamically stable • d < 100 nm • Transparent or slightly turbid

ACCEPTED MANUSCRIPT

SC

Kinetically Stable Nanoemulsion

RI PT

Figure 2: Schematic diagram of most common instability mechanisms that occur in food emulsions: creaming, sedimentation, flocculation, coalescence, Ostwald ripening and phase inversion.

TE D

M AN U

Gravitational Separation

AC C

EP

Sedimentation Creaming

Flocculation

Coalescence or Ostwald Ripening

Oiling Off

ACCEPTED MANUSCRIPT

SC

RI PT

Figure 3. The particles in beverage emulsions can be designed to have different functional performances by varying their physicochemical and structural properties, such as size, composition, charge and interfacial properties.

TE D

AC C

Charge -

+ +

+ -

-

-

+ -

-

+

-

EP

Composition

Interfacial properties

Negative

Neutral

+

+

Particle Size

M AN U

Droplet

Positive

ACCEPTED MANUSCRIPT

w(h)

RI PT

Energy Barrier

h Attraction

AC C

EP

TE D

M AN U

SC

Repulsion

Coalesced

1º Min

2º Min

Figure 4. Schematic representation of the interaction potential between two emulsion droplets. The droplets may be stable, weakly flocculated, strongly flocculated, or coalesced depending on the attractive and repulsive forces between them.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 5. Lipid droplets have a major impact on the texture, appearance and stability of emulsions

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Flavor Polarity

Figure 6. Initial flavor intensity of emulsions with fixed flavor content is strongly dependent on oil droplet content.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 7. Droplet growth due to Ostwald ripening of orange oil emulsions can be inhibited by adding a corn oil, which acts as a ripening inhibitor (McClements et al 2012)

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 8. Influence of pH on the chemical degradation rate of citral in aqueous solutions: the degradation rate increases with decreasing pH (Choi et al 2009).

ACCEPTED MANUSCRIPT

RI PT

Figure 9: Schematic representation of mechanical devices that can be used to produce beverages emulsions using the high-energy approach: high pressure valve homogenizer, microfluidizer and continuous ultrasonic homogenizer

M AN U

SC

Droplet disruption

Droplet disruption

AC C

EP

TE D

High Pressure Valve Homogenizer

Continuous Ultrasonic Homogenizer

Microfluidizer

Droplet disruption

ACCEPTED MANUSCRIPT

RI PT

Figure 10: Schematic representation of various methods that can be used to produce beverages emulsions using a low-energy approach: emulsion inversion point method, spontaneous emulsification and phase inversion temperature.

T > PIT

W/O

O/W/O

O/W

AC C

EP

TE D

Emulsion inversion point method

M AN U

SC

W/O Emulsion

Cool

T = PIT

Bicontinuous Microemulsion

Heat T < PIT

O/W Emulsion

Spontaneous emulsification method PIT method

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 11: Influence of homogenization pressure in a microfluidizer on the mean droplet diameter of oil-in-water emulsions stabilized by either β-lactoglobulin (BLG) or sodium dodecyl sulfate (SDS). The oil phase density was increased by adding increasing amounts of weighting agent (Qian, McClements 2011).

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 12: Influence of surfactant-to-oil ratio on the formation of 20 wt% oil-in-water emulsions stabilized by a non-ionic surfactant (Tween 80) prepared using the spontaneous emulsification method (Sabeti, Fang, and McClements 2012). The oil phase consisted of 80% vitamin E acetate and 20% MCT.

ACCEPTED MANUSCRIPT

Figure 13: Schematic diagram of formulation-composition map for a typical surfactant-oil-water system. This map can be used to understand the formation of emulsions and nanoemulsions suing various phase inversion methods.

RI PT

W/O/W

M AN U

SC

W/O

Transitional

TE D

O/W

O/W/O

AC C

EP

Catastrophic

• HLD – Hydrophilic-Lipophilic Deviation • Relative affinity of surfactant for water and oil phases (surfactant & system dependent) • WOR – Water-to-Oil Ratio • Depends on system composition

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 14: Change in turbidity when a coarse lemon oil emulsion (10% lemon oil, 13% Tween 80, 77% water) is heated and cooled.. A very fine transparent nanoemulsion is formed when the system is heated, and then cooled below its phase inversion temperature. (Rao and McClements 2010)

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 15: Influence of oil type on mean droplet diameter in emulsions produced using the emulsion inversion point (EIP) method by titrating water into an oil-surfactant mixture (Ostertag, Weiss and McClements, 2012).

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 16: Influence of pH and storage time on the chemical degradation of 0.05% citral dispersed in citrate buffer solutions (Choi et al 2009). The citral chemically degrades rapidly at acidic pH values.

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 17: Influence of oil droplet concentration on the chemical degradation of citral dispersed within MCT oil-in-water emulsions (Choi et al 2009). The citral chemically degrades more rapidly in an aqueous than an oil environment, and therefore the less oil present the faster the degradation rate.

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 18: Influence of oil droplet concentration on the partitioning of citral within MCT oil-in-water emulsions (Choi et al 2009). The mass fraction of citral in the aqueous phase decreases as the oil concentration increases.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 19: Influence of emulsifier concentration on the mean particle diameter produced using a microfluidizer using either whey protein isolated (WPI), modified starch (MS) or gum arabic (GA) are emulsifiers under standardized conditions.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 20: Influence of pH on the particle size and charge produced using a microfluidizer using either whey protein isolated (WPI), modified starch (MS) or gum arabic (GA).

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 21: Influence of ionic strength on the particle size and charge produced using a microfluidizer using either whey protein isolated (WPI), modified starch (MS) or gum arabic (GA).

ACCEPTED MANUSCRIPT

Table 1: Global revenue from non-alcoholic beverage sales by major beverage manufacturing companies in 2010 (Beverage

35.1

PepsiCo

21.4

Kraft Foods

8.8

Dr. Pepper Snapple Group

5.6

AC C

EP

TE D

Coca-Cola Co.

SC

Annual Sales (Billions of Dollars)

M AN U

Company

RI PT

Executive. October 2011, “Worldwide 100”, Pages 44-48).

ACCEPTED MANUSCRIPT

Table 2. Summary of major colloidal interactions operating between oil droplets in beverage emulsions. It is assumed that the

RI PT

droplets are coated by the same type of emulsifier. Sign

Magnitude

Range

Factors Affecting

Van der Waals

Attractive

Intermediate

Intermediate

Always present

Steric

Repulsive

Strong

Short

Thickness and chemistry of interface

SC

Interactions

Repulsive

Strong-to-Weak

Long-to-Short

pH and ionic strength

Depletion

Attractive

Weak-to-Medium

Short

Amount & type of non-

M AN U

Electrostatic

adsorbed polymer

Short

TE D

Attractive

Strong

Strong

Long

EP

Hydrophobic

Attractive

AC C

Bridging

Amount & type of adsorbing polymer Surface hydrophobicity

ACCEPTED MANUSCRIPT

Table 3. Physicochemical properties of oil phase, water phase, and weighting agents commonly used in the formulations of

reported.

Density (kg m-3)

Restriction Level (ppm)

-

≈1050

-

-

≈1000

-

Yes Yes

840-850 900-920

-

No

1240-1330

15

Yes

1060

Defined by manufacturer

1080

Not approved for use in the United States 100

No

1146

300

Structure

Natural

SC

Components

Oil Phases Flavor Oils Vegetable Oils

Mainly Terpenes Mainly Triglycerides

Weighting Agents

Ester Gum Sucrose Acetate Isobutyrate

EP

Dammar Gum

Brominated Hydrocarbon Chain of Triacylglycerols Exudate from Caesalpinaceae and Dipterocarpaceae shrubs Hydrophobic Polymer

AC C

Brominated Vegetable Oil

TE D

Low calorie

M AN U

Aqueous Phases Sugared

RI PT

beverage emulsions (Chanamai & McClements, 2000; McClements, 2005). The restriction levels based on USA regulations are also

Esterification of Sucrose with Acetic and Isobutyric Anhydrides

ACCEPTED MANUSCRIPT

SC

HLB Value 16 15 6 2 16 15.6 14.9 15 >9 8.6 6.7 4.7 4.3

AC C

EP

TE D

M AN U

Surfactant Sucrose monopalmitate Sucrose monostearate Sucrose distearate Sucrose polystearate Tween 20 (Polyoxyethylene Sorbitan Monolaurate) Tween 40 (Polyoxyethylene Sorbitan Monopalmitate) Tween 60 (Polyoxyethylene Sorbitan Monostearate) Tween 80 (Polyoxyethylene Sorbitan Monoooleate) Q-Naturale (Quillaja saponin) Span 20 (Sorbitan monolaurate) Span 40 (Sorbitane monopalmitate) Span 60 (Sorbitan monostearate) Span 80 (Sorbitan monooleate)

RI PT

Table 4. HLB numbers of selected small molecule surfactants and co-surfactants that could be used to produce beverage emulsions. Information from various sources: McClements (2005), National Starch, and Rahn-Group.com.

ACCEPTED MANUSCRIPT

Source

Main Structure Type

Major Monomer(s)

Algal

Linear

Carrageenan

Algal

Linear/Helical

Sulfated Galactose & Anhydrogalactose

Chitosan

Crustaceans, Invertebrates

Linear

Acacia Sap

Branched Coil Domains on Protein Scaffold

Carob seeds

Linear

Linear

Modified Starch

Waxy Maize

Linear

Glucose with some nonpolar branches

Plant Cell Walls Xanthomonas campestris exudate

Highly Branched Coil

Glucuronate (backbone)

AC C

Wood Pulp

Xanthan Gum

Negative

Neutral

Galactose and mannose

Modified Celluloses

Pectin

Positive

Galactose

TE D

Locust bean gum

Negative

Glucosamine

EP

Gum Arabic

Negative

M AN U

Alginate

Mannuronic & Guluronic Acids

Charge

SC

Name

RI PT

Table 5. Summary of molecular characteristics of some common food-grade polysaccharides. Adapted from Matalanis et al (2011).

Linear/Helical (High MW)

Glucose and modified Glucoses

Neutral or Negative Negative Negative Negative

Glucose (backbone)

ACCEPTED MANUSCRIPT

Table 6. Examples of the composition and calorie content of some commercial beverage products currently on the market. Emulsifier

Weighting Natural Agent Claim

50

GA

None Listed

No

0

GA

None Listed

No

Cane Sugar

70

GA

EG

Yes

Cane Sugar

70

GA

EG

Yes

HFCS

80

GA

EG

No

HFCS

80

GA

EG

No

RTD

HFCS

80

GA

EG

No

RTD

HFCS

80

M-FS

MCT, SAIB

No

RTD

HFCS

120

M-FS

EG, BVO

No

APM, Ace-K

0

GA

EG, BVO

No

RTD

Fructose, Cane Sugar

120

GA

EG

No

Coca-Cola

RTD

Fructose, Cane Sugar

120

GA

EG

No

Coca-Cola

RTD

Fructose, Cane Sugar, Sorbitol

120

GA, M-FS

EG

No

Coca-Cola

RTD

Erythritol, Rebiana

0

GA

EG

No

Beverage Sweetener Type System

Half and Half Iced Tea Lemonade

Arnold Palmer

AriZona

RTD

HFCS, Sucralose, Ace-K

Half and Half Iced Tea Lemonade

Arnold Palmer Zero

AriZona

RTD

Sucralose

Cherry Citrus

BodyArmor

Body Armor Nutrition

RTD

Tropical Mandarin

BodyArmor

Body Armor Nutrition

RTD

Lemon Lime

Xion4 (Powerade)

Coca-Cola

RTD

Orange

Xion4 (Powerade)

Coca-Cola

RTD

Strawberry Lemonade

Xion4 (Powerade)

Coca-Cola

Sour Melon

Xion4 (Powerade)

Coca-Cola

Orange Soda

Fanta

Coca-Cola

Original Citrus

Fresca

Coca-Cola

RTD

Energy - Tropical Citrus

Vitamin Water

Coca-Cola

Multi-V - Lemonade

Vitamin Water

Essential Orange-Orange

Vitamin Water

Squeezed Lemonade

Vitamin Water Zero

TE D

EP

AC C

SC

Manufacturer

M AN U

Brand

RI PT

Calories (per 8 oz. Ready to Drink)

Product

ACCEPTED MANUSCRIPT

Sunkist

Dr. Pepper Snapple

RTD

HFCS

110

M-FS

EG

No

Orange

Diet Crush

Dr. Pepper Snapple

RTD

APM, Ace-K

0

GA

EG

No

Citrus Soda

Sun Drop

Dr. Pepper Snapple

RTD

HFCS

120

GA

EG, BVO

No

Pineapple Orange Guava

Nantucket Nectar's

Dr. Pepper Snapple

RTD

Sucrose

120

GA

None Listed

Yes

Orange Mango

Nantucket Nectar's

Dr. Pepper Snapple

RTD

Sucrose

120

GA

None Listed

Yes

Strawberry Kiwi

Snapple

Dr. Pepper Snapple

RTD

Sucrose

100

GA

EG

Yes

Lemonade

Farmer's Daughter

The Farmer's Cow

RTD

Cane Sugar

120

GA

EG

Yes

Strawberry Lemonade

Farmer's Daughter

The Farmer's Cow

RTD

Cane Sugar

110

GA

EG

Yes

Lemonade

MiO

Kraft Foods

LC

Sucralose, Ace-K

0

GA

SAIB

No

Orange Tangerine

MiO

Kraft Foods

LC

Sucralose, Ace-K

0

GA

SAIB

No

Sparkling Lemonade

N/A

Lorina

RTD

Sucrose

90

GA

EG

No

Pink Citrus Lemonade

N/A

Lorina

RTD

Sucrose

120

GA

EG

No

Lemonade

G2 of Gatorade G-Series

PepsiCo

RTD

Sucrose

80

GA

EG, SAIB

No

Lemon-Lime

G2 of Gatorade G-Series

PepsiCo

RTD

Sucrose

80

GA

EG

No

Orange

G2 of Gatorade G-Series

PepsiCo

RTD

Sucrose

80

GA

EG, BVO

No

SC

M AN U

TE D

EP

AC C

RI PT

Orange Soda

ACCEPTED MANUSCRIPT

RTD

HFCS

110

GA

BVO

No

N/A

PepsiCo

RTD

APM, Ace-K

0

GA

BVO

No

Nature's Blends

Poland Spring

RTD

Cane Sugar

50

GA

EG

Yes

Orange Cranberry Tangerine

Skinny Water

Skinny Nutritional

RTD

Sucralose, Ace-K

0

GA

EG

No

Old-Fashioned Lemonade

N/A

Zeigler's Beverages

RTD

Sucrose

100

GA

EG

No

SC

Key:

Ace-K

Acesulfame Potassium

APM

Aspartame

BVO

Brominated Vegetable Oil

EG

Ester Gum

GA

Gum Acacia

HFCS

High Fructose Corn Syrup

Liquid Concentrate

LC

Medium Chain Triglycerides

MCT

M-FS RTD SAIB

M AN U

Lemon

TE D

Diet Mtn Dew

RI PT

PepsiCo

EP

N/A

ModifiedFood Starch

AC C

Mtn Dew

Ready to Drink Sucrose Acetate Isobutyrate

ACCEPTED MANUSCRIPT

Claim

Justification

“0 Calories”

Less than 5 cal per RACC.

RI PT

Table 7. Claims and justifications that can be made on beverage products.

Source 21 CFR 101.60(b)

“All Natural”

M AN U

components.

SC

Undefined, but regarded as a product lacking any synthetic

N/A

Any substance that imparts flavor that is not derived from its natural “Artificially Flavored”

21 CFR 101.22 (a) (1)

source.

For juice made from concentrate, calculate use-percentage from the

21 CFR 101.30(j),

Brix table in 21 CFR 101.30(h)(1) as the basis for 100% juice.

21 CFR 101.30(h)

Contains 20% or more of the DV per RACC.

21 CFR 101.54(b)

TE D

“Contains % Juice”

“Excellent Source of

Vitamin” “Low Calorie Soda” “Low Sodium”

Contains 10%-19% of the DV per RACC.

21 CFR 101.54(e)

40 cal or less per RACC.

21 CFR 101.60(b)

140 mg or less per RACC.

21 CFR 101.61

AC C

“Good Source of

EP

Vitamin”

ACCEPTED MANUSCRIPT

21 CFR 101.22 (a) (1), “Natural and Artificial Contains Natural and Artficial Flavors.

RI PT

Flavor”

21 CFR 101.22 (a) (3), 21 CFR (g) 101.22 (3)

The flavor constituents function in the food as flavor, not solely as a “Naturally Flavored”

21 CFR 101.22 (a) (3)

No Sugar or Sugar-Containing Ingredients are added during processing.

M AN U

“No Added Sugar”

SC

source of nutrition are derived from its natural source in nature.

21 CFR 101.60(c) (2)

Contains no colarant source outside the principle flavor components of

“No Artificial Colors”

21 CFR 101.22(k)(1) (2),

beverage. Natural color is considered artficial color when used for 21 CFR 74

TE D

colorant purposes.

“No Pulp”

“Reduced Sugar”

EP

21 CFR 101.65 (b)(2) 21 CFR 101.65(b)(1)

Cannot Contain Preservatives.

AC C

“No Preservatives”

21 CFR 101.65(b)(1)

Cannot Contain HFCS.

“No HFCS”

21 CFR 101.65 (b)(2) 21 CFR 101.65(b)(1)

Cannot Contain Pulp.

At least 25% less sugars per RACC than an appropriate reference food.

21 CFR 101.65 (b)(2) 21 CFR 101.60 (c)