You are on page 1of 22

Cellulose Derivatives

In making cellulose derivatives (Chapter 8), the polymer is steeped in a strongly


alkaline solution and “aged” in a manner that promotes oxidation of hydroxyl groups
by dissolved oxygen from air.

From: Carbohydrate Chemistry for Food Scientists (Third Edition), 2019

Related terms:

Hydrocolloid, Emulsification, Yogurt, Fat Replacers, Functional Food, Edible Film,


Sweetener, Casein, Milk Protein

View all Topics

Antimicrobial Food Packaging Based


on Biodegradable Materials
V. García Ibarra, ... A. Rodríguez-Bernaldo de Quirós, in Antimicrobial Food Packag-
ing, 2016

29.4.1.3 Cellulose
Cellulose derivatives such as methylcellulose, hydroxypropyl cellulose, and car-
boxymethyl cellulose (CMC) are being extensively used to produce films due to
their suitable properties (Sanla-Ead et al., 2012). Cellulose alone or in combination
with other matrixes has been employed in food packaging applications. In a study
carried out by Barbiroli et al. (2012) two AM proteins, lactoferrin and lysozyme,
were incorporated into a cellulose-based material. The active system was applied to
thin slices of raw meat. The results suggested that lysozyme or lysozyme/lactoferrin
seems to be the best combination for this particular application.

Sebti et al. (2007) reported the preparation of different films based on chitosan com-
bined or not with hydroxypropyl methylcellulose polymer (HPMC), among others.
The chitosan-based films combined or not with HPMC presented AM properties.

> Read full chapter


Functionality of Hydrocolloids in Bat-
ter Coating Systems
Marc A. Meyers, Andrew Grazela, in Batters and Breadings in Food Processing
(Second Edition), 2011

MC AND HPMC
The cellulose derivatives MC and HPMC have all of the previously described ma-
jor functional properties or benefits that make them attractive for use in batters
(Greminger and Savage 1973; Dow Chemical 1984, 1987, Dow Chemical 1988–d;
Meyers 1988a,c, 1989). They also have certain properties, such as thermal gelation,
that are unique. This property, along with film-forming properties, makes these
gums useful in formulating edible barriers for moisture retention and oil reduction.

Figures 7.7 and 7.8 show studies performed using Methocel (a trademark of The
Dow Chemical Company) premium food gums, which are commercially available
grades of MC and HPMC. The graphs indicate that these cellulosic derivatives are
able to reduce oil and increase moisture retention in battered and breaded fried
foods. When these particular gums were added into the water phase as a prehydrated
solution, a significantly higher barrier for reducing oil was found. This performance
is shown in Table 7.6 (Meyers 1988a,1988c, 1989).

Fig. 7.7. Oil absorption vs. frying time. A commercial low-viscosity batter system was
used, with hydroxypropyl methylcellulose (HPMC) added at a 0.55% level (w/w, wet
batter basis; 1.8% w/w, dry batter basis). A two-pass battered and breaded system
was used, and the samples were fried for 2 min at 188°C. Methocel is a trademark
of The Dow Chemical Company.
Fig. 7.8. Moisture retention vs. frying time. A commercial low-viscosity batter system
was used, with hydroxypropyl methylcellulose (HPMC) added at a 0.55% level (w/w,
wet batter basis; 1.8% w/w, dry batter basis). A two-pass battered and breaded system
was used, and the samples were fried for 2 min at 188°C. Methocel is a trademark
of The Dow Chemical Company.

TABLE 7.6. Oil Reduction and Moisture Retention Properties of Methylcellulose


(MC) and Three Common Substitution Ranges of Hydroxypropyl Methylcellulose
(HPMC)a

Methoxyl Substi- Hydroxypropoxyl Fat Reduced (%)b Moisture In-


tution (%) Substitution (%) creased (%)b
MC 27.5–31.5 0.00 18.4–36.0 10.0–25.5
HPMC 28.0–30.0 7.0–12.0 25.6–48.2 20.0–26.3
27.0–30.3 4.0–7.5 31.5–55.4 20.0–28.7
19.0–25.0 4.0–12.0 20.6–44.2 17.0–25.4

a Solutions were made using the polymer and the water fraction of the batter
formulation and were allowed to hydrate for three days at 4°C before the
remaining ingredients were added. Polymer concentration was 0.5% w/w,
wet-batter basis.
b Percentages based on the total weight of the finished fried product.

These cellulosic derivatives can also improve product yield and lower the use of batter
solids (resulting in batters with a high water-solids ratio and high WHC) (Meyers
1988a,1988c, 1989; Duxbury 1989). With these gums, batters can be formulated to
provide a particular function or a combination of functions at varying degrees of
effectiveness.

> Read full chapter


Biopolymers-embedded nanoemul-
sions and other nanotechnological ap-
proaches for safety, quality, and stora-
bility enhancement of food products:
active edible coatings and films
Hadar Arnon-Rips, Elena Poverenov, in Emulsions, 2016

6.2.1 Polysaccharides-Based LbL


Alginate, cellulose derivatives, pectin, chitosan, and carrageenan represent the most
utilized polysaccharides in the field of LbL EFs and ECs. An exactly defined chemical
structure of a polysaccharide’s repetitive monomer unit allows the prediction and as
a consequence, the design of EFs’ and ECs’ properties.

Poverenov’s group has published three papers on polysaccharide-based LbL edible


coatings (Arnon et al., 2014, 2015; Poverenov et al., 2014a). Two of these studies
examined carboxymethyl-cellulose (CMC)/chitosan bilayered ECs on citrus fruit and
the third study examined an alginate/chitosan EC on fresh-cut melons. Chitosan, a
positively charged polysaccharide, served as the external layer in all three studies
due to its eminent antimicrobial activity, while the negatively charged CMC or
alginate constituted the inner layer due to their satisfying adhesive properties. These
formulations allowed for the bilayer coatings to gain the individual advantages from
both comprising polymers.

In the first study Arnon et al. (2015) performed a series of systematic experiments
with the aim of finding polysaccharide-based ECs that convey enhanced quality,
improved storage duration, and had an attractive appearance of citrus fruit. Such
polysaccharides-based coatings may provide a natural alternative to synthetic waxes
that are currently utilized on citrus fruit (Bai and Plotto, 2011). After screening a
series of different polysaccharides on mandarin fruit, CMC in a fixed concentration
of 1.5% w/v was chosen as an inner coating layer due to its best adhesion properties
in addition to good gas and water vapor permeability. Chitosan was applied as an
external active layer. Chitosan was found to cause an improvement in fruit texture
and gloss. Glycerol and fatty acids were added to the CMC coating formulation
instead of the second chitosan layer, in order to compare their effect with the LbL
approach. However, these auxiliary additives did not result in a desired improvement
of coating properties. Moreover, in several cases they caused deterioration in the
coating’s properties. Various chitosan concentrations were examined in order to
formulate the optimal LbL coating. The obtained results showed that the higher the
chitosan concentration used, the more gloss was seen in coated fruits, as well as
an improvement in firmness and a ripening rate inhibition (Fig. 10.3). On the other
hand it should be noted that excessive chitosan content causes an increase in ethanol
concentrations. An accumulation of high ethanol concentration is not desired, since
it may cause off-flavors. This is especially important for mandarins, which are the
most off-flavor sensitive among all the citrus fruit.

Figure 10.3. The effect of chitosan concentration in the LbL EC on (a) fir-
mness/weight loss, (b) gloss, (c) CO2/EtOH accumulation, and (d) progress of ripening
in mandarin fruit. This was compared to the effect of single layered coatings,
synthetic polymer-based commercial wax (PE-CW), and uncoated fruit (control).
Measurements were taken after 10 days at 20°C and represent means of repli-
cations ± SE. Different letters indicate significantly different values (at p≤0.05)
according to a Tukey-Kramer HSD test (Arnon et al., 2015).

Bilayered coating formulations of 1.5% (w/v) CMC and 1% (w/v) chitosan were
selected as the optimal coating content. In the second study Arnon et al. (2014)
have examined the effects of this coating formulation on the postharvest quality of
oranges, grapefruit and two types of mandarins at simulated storage conditions.
The coatings demonstrated good performance and showed that can serve as a
considerable alternative to the currently used synthetic waxes.

The third study published by Poverenov’s group involved the implementation of a


bilayered polysaccharide-based coating composed of polyanionic alginate and poly-
cationic chitosan on fresh-cut melons (Poverenov et al., 2014a). The application of
ECs to minimally processed fruits faces some technical problems. These are related
to the difficult adhesion of materials to the cut fruit hydrophilic surface. In addition,
fresh-cut fruits are usually very prone to microbial spoilage and texture deterioration.
The alginate/chitosan LbL coating was found to possess beneficial properties from
both ingredients. The bilayer combination allowed for a good adhesion to the melon
matrix by the inner alginate layer, along with good antimicrobial activity provided
by the outer chitosan layer (Fig. 10.4).

Figure 10.4. Effect of edible coating on the microbial quality of melons after 10 days
of storage at 6oC.

This antimicrobial protection allows reducing bacteria, yeast, and fungi counts. Both
layers contributed to the melons’ texture enhancement; alginate due to a presence
of Ca+2 ions in its formulation and chitosan by avoiding microbial damages that
affect fruit entirety. The coating slowed down tissue texture degradation, so that after
14 days of storage, only LbL samples maintained an appreciable firmness. This is
unlike uncoated melons or fresh-cut melons coated with a single layer of alginate or
chitosan. Surprisingly, unlike the previously described studies on citrus fruits coated
by a bilayered coating, the bilayered coating exhibited enhanced gas exchange prop-
erties that exceeded those of both monolayer coatings and even of the noncoated
control. As a result, the bilayered coating prevented an increase in headspace CO2
and ethanol concentrations, which are synonymous to hypoxic stress and off-flavor
development as observed in other samples, especially in alginate-coated melons.
Polysaccharide-based EFs and ECs prepared by the LbL electrostatic deposition
approach can consist of more than two layers, as was demonstrated by two papers
published by Vincente’s group (Medeiros et al., 2012a; Souza et al., 2015). In the
first study, nanomultilayer EFs were prepared, consisting of five alternating layers of
polyanionic pectin and polyactionic chitosan (Medeiros et al., 2012a). This formula-
tion was first characterized in terms of water vapor and oxygen and CO2 permeability.
Further on, the nanomultilayer system was applied on whole “Tommy Atkins” man-
goes and the layers’ adsorption was confirmed by changes in the coated fruits’ skin’s
contact angle. After 45 days in storage, coated mangoes presented a better external
appearance and a less dehydrated surface. No fungal growth was observed, with
a much more preserved flesh in comparison to the uncoated fruit. These findings
suggest that a combination of chitosan’s antimicrobial and gas barrier properties,
along with pectin’s low oxygen permeability, were possibly efficient in the reduction
of gas flow and on the extension of the mangoes’ shelf life. In the second study, a
nanomultilayer EC consisting of five layers of polyanionic alginate and polycationic
chitosan were prepared and applied on fresh-cut mangoes (Souza et al., 2015). The
fruit quality was evaluated after 14 days of storage at 8°C. The LbL coated fresh-cut
mangoes exhibited a high value of titratable acidity and lower values of soluble
solids, mass loss, pH and browning rate. In addition, the LbL alginate/chitosan EC
improved the fruits’ microbial quality during storage and extended their shelf life up
to 8 days in 8°C when compared to uncoated fruit (<2 days).

> Read full chapter

Polysaccharides
James N. BeMiller, in Carbohydrate Chemistry for Food Scientists (Third Edition),
2019

Thixotropic flow
Thixotropic flow is another type of non-Newtonian, shear-thinning flow. The vis-
cosity of thixotropic solutions (like that of pseudoplastic solutions) decreases as the
rate of shear increases, but it changes in a time-dependent manner rather than in-
stantaneously. Also, like pseudoplastic fluids, thixotropic fluids regain their original
viscosity after cessation of shear, but again only after a measurable time interval
(which can range from fractions of a second to hours), rather than instantaneously
(Fig. 5.13). This behavior is due to a gel → sol → gel transition. A thixotropic solution
at rest is a weak, often pourable, gel. In other words, a thixotropic solution has
properties of a gel (at rest) and properties of a fluid when a shear force is applied.
Alginates (Chapter 14).are examples of hydrocolloids whose solutions can exhibit
thixotropic behavior.

Figure 5.13. Idealized representation of thixotropic rheology showing the time de-
pendence of the change in viscosity with a change in shear rate. Compare with
pseudoplastic rheology (Fig. 5.9).

Solutions of certain cellulose derivatives (especially certain types of CMCs) may


exhibit thixotropic behavior. The degree of thixotropy is a function of the degree
of substitution (DS), the uniformity of substitution, and the DP of the polymer
(Chapter 8). Cellulose molecules associate easily and strongly with each other, so
nonuniformly derivatized cellulose molecules that have stretches of unmodified --
d-glucopyranosyl units can form limited intermolecular associations called junction
zones (section on Gels in this chapter). These interactions are responsible for the
weak, easily broken gel structure. Energy may be required to break the junction zones
and start flow. Hysteresis12 loops are characteristic of thixotropic flow (Fig. 5.14).
Because the pulp in fruit juices provides a slight thixotropic characteristic, a dry mix
for a fruit drink should mimic this behavior.

Figure 5.14. One type of graphical presentation of the hysteresis loop of thixotropic
solutions: viscosity versus shear rate.
> Read full chapter

Fat Replacers
Thomas P. O'Connor, Nora M. O'Brien, in Reference Module in Food Science, 2016

Cellulose
Several types of cellulose and cellulose derivatives can be used as fat mimetics,
usually in combination with other hydrocolloids. Microcrystalline cellulose (MCC)
is a non-caloric fat replacer. It mimics fat in aqueous systems by contributing
to body, mouthfeel and viscosity. Typically, 60–70% of the cellulose microcrystals
are less than 0.2 μm long. These microcrystals form an insoluble dispersion in
water. Other ingredients (e.g., gums) hold the insoluble microcrystals together in
a network which gives a creamy mouthfeel and body similar to that found in full-fat
products. Other forms of cellulose such as powdered cellulose, methylcellulose and
hydroxypropylmethylcellulose also have application as fat mimetics in baked goods,
frozen desserts, sauces and salad dressings.

> Read full chapter

Cellulose and Cellulose-Based Hydro-


colloids
James N. BeMiller, in Carbohydrate Chemistry for Food Scientists (Third Edition),
2019

Modified cellulose products


The principle behind converting cellulose into water-soluble products (hydrocol-
loids) is to separate the molecules in a cellulose fiber, then derivatize some of their
hydroxyl groups so that the molecules cannot reassociate through intermolecular
hydrogen bonding over enough of their length to bring about insolublization.
Swelling of cellulose fibers to separate the cellulose molecules is accomplished by
treating a highly purified wood pulp (or cotton linters) with 18% sodium hydroxide
solution. This treatment forms “alkali cellulose”. In alkali cellulose, many of the
hydroxyl groups (-OH) have been converted into alkoxy groups (-O−), which are quite
nucleophilic. (The alkoxy groups also give the molecules negative charges which
repel each other and swell the fibers, destroying some of their crystallinity and
making the molecules more accessible to derivatizing reagents.) (Note that a much
more alkaline system is used than is used for the derivatization of starches.) If an
oxidant, such as oxygen, is present, some hydroxyl groups are oxidized to carbonyl
groups, and under the alkaline conditions, a beta-elimination reaction (Chapter 4)
occurs and the cellulose is depolymerized (that is, reduced in average molecular
size). So various reduced viscosity types are made by aging the alkali cellulose (that is,
holding it until the desired average MW is achieved). All modified cellulose products
used in foods are cellulose ethers made by reaction of alkali cellulose with an alkyl
halide (to make CMC and MC) or an epoxide (to make a hydroxyalkylcellulose,
namely, HPC) (Fig. 8.3).

Figure 8.3. Representation of the introduction of carboxymethyl, methyl, and hy-


droxypropyl ether groups into a cellulose molecule. In the substituted molecule
(bottom one with R groups), there are two substituent groups at an O2 position
and one at an O6 position. There are three substituent groups per nine hydroxyl
groups (three glucopyranosyl units) and, thus, a DS of 1.0. The diagram indicates
that a polysaccharide composed only of neutral hexopyranosyl units (like cellulose) is
a molecule of only primary and secondary hydroxyl groups available for substitution.

Variables in cellulose derivatives are (1) the amount of etherification, (2) the nature
of the substituent group, (3) the average MW of the derivatized cellulose molecules,
(4) the distribution of the substituent groups on O2, O3, and O6 of the -d-glucopy-
ranosyl units, and (5) the distribution of the substituent groups along the cellulose
chain.

The average MW (average DP) of a cellulose derivative determines the viscosity


type. The largest molecules produce the highest viscosity (Chapter 5). For example,
viscosities of 1.0% solutions of CMC at room temperature can range from less than
10 mPa s (cps) for a low-MW product to about 10,000 mPa s for a high-MW product.
The nature and number of substituent groups determines solubility. Distribution of
substituent groups determines the rheology type (of importance only with CMC).

General characteristics of water-soluble cellulose derivatives are as follows:

• Wide range of viscosity types available

• High-viscosity types available

• High purity

• pH stable

• Tasteless

• Odorless

• Uniform quality

• Form clear solutions

• Film forming

• Noncaloric

Carboxymethylcelluloses
Carboxymethylcellulose (CMC), the sodium salt of the carboxymethyl ether of cel-
lulose,10 is widely used as a hydrocolloid. For production of CMC, alkali cellulose is
reacted with the sodium salt of chloroacetic acid. (Alkali cellulose is shown below
as being 100% ionized, but may not be. In the case of cellulose, a much stronger
solution of sodium hydroxide is used, so there are undoubtedly more hydroxyl
groups in the alkoxy (R–O ) form than in the case of starch.

Most food-grade sodium CMC products have degree of substitution (DS) values
in the 0.7–0.8 range, but some products may have a DS as low as 0.4. (For water
solubility, CMC must have a DS of at least 0.4.) The usual CMCs of DS 0.7–0.8
hydrate rapidly because they are quite ionic. Therefore, to make a smooth solution
(that is, one without lumps), one of the means of dispersing it described in Chapter
5 must be followed. As the DS of CMC increases, salt tolerance, hygroscopicity, and
alcohol tolerance increase and its thixotropic nature (see below) decreases.

Substituent groups are found at O2, O3, and O6 of d-glucopyranosyl units in the
approximate proportion of 2.1:1.0:1.6. The degree of uniformity of derivatization
along the cellulose chain determines the behaviors of its solutions. Nonuniformly
derivatized CMC is a good example of a hydrocolloid that produces solutions with
thixotropic rheology. Nonuniform distribution leaves unsubstituted stretches of
molecules available for junction zone formation with unsubstituted regions of other
CMC molecules, producing thixotropy. As explained in Chapter 5, thixotropic fluids
have the characteristics of a weak gel at rest. Why a CMC with nonuniform substi-
tution behaves this way and undergoes shear thinning is diagrammed in Fig. 8.4.
Nonuniformly derivatized CMC produces solutions with thixotropic rheology (that
is, with hysteresis as apposed to instantaneous, reversible shear thinning [pseudo-
plasticity]) (Chapter 5) because (1) the molecules with shorter contacts (contacts held
together by the fewest hydrogen bonds) disengage from junction zones first, while
those with longer contacts require more energy and time for pulling away from
associations of molecules, and (2) the relatively short “naked regions” require time to
“find each other” and reestablish a junction after shear is removed. So there is a time
dependency to both shear thinning and the subsequent rethickening when shear is
removed. Uniform derivatization produces molecules with no long unsubstituted
stretches. Because they do not contain “naked regions” that can bind to the “naked
regions” of other CMC molecules, products with uniform substitution will form
smooth, stable solutions. The “naked regions” of nonuniformly substituted CMC
allow binding to powdered cellulose and MCC. Nonuniformly substituted CMC also
interacts with certain hydrocolloids such as galactomannans (Chapter 9). Maximum
viscosity is produced when CMC and guar gum are used together in a CMC to guar
gum ratio of about 1:3.

Figure 8.4. A diagrammatic explanation of why a solution of nonuniformly substi-


tuted carboxymethylcellulose is a weak gel at rest and undergoes shear thinning
when stirred, pumped, or swallowed (that is, is thixotropic). The diagram depicts a
situation where sufficient shear is applied to (A) break all associations and (B) to
align the dissociated molecules in a fully extended, linear arrangement. Neither will
completely occur at low rates of shear.

CMC is available in a wide range of viscosity types, and as already discussed, solution
viscosity depends mainly on the average MW of the water-soluble polysaccharide
(Chapter 5). Solutions of higher MW (higher viscosity) products also exhibit greater
pseudoplasticity or thixotropy. Because it is a highly ionic hydrocolloid, CMC's sol-
ubility and the viscosity of its solutions are affected by salts. As explained in Chapter
5 (Fig. 5.13), CMC should be dissolved in water before other solutes are added (not
dispersed in a salt solution) to take full advantage of its properties. Effects of salts
on solutions of CMC are a function of the type (DS and viscosity type) of the CMC,
the type and concentration of the salt, and the pH. In general, monovalent cations
form soluble CMC salts; divalent cations produce hazy dispersions; and salts of
trivalent cations are insoluble (that is, aluminum ions will bring about precipitation,
but aluminum ions are not encountered in foods).

Because CMC consists of long, fairly stiff molecules bearing negative charges,
its molecules in solution are stretched out due to electrostatic repulsion of chain
segments (that is, chain folding is restricted because any folding would bring
the carboxylate groups closer together where their negative charges would repel
each other). In addition, because the molecules repel each other, monodisperse,11
highly viscous, stable solutions result. Lowering the pH to less than 4 represses
ionization of the carboxyl groups so that some lose their charge (-COO− → -COOH).
Molecular association then occurs and viscosity increases for a time, but viscosity
cannot be maintained long term at pH values of less than 4 because of hydrolysis
(depolymerization). At pH 3, insolubilization (precipitation) occurs due to extensive
loss of the negative charge, changing the anionic polymer to a neutral polymer,
which allows chain associations to occur. CMC undergoes hydrolysis with loss of
viscosity when acidic solutions of it are retorted.

CMC is widely used in food products to absorb and hold water, to control crystal
growth, to thicken, as a binder, to increase shelf life, and to provide desired texture
or body. Its largest volume use is in the preparation of dry pet foods that form their
own gravy when warm water is added. Its second largest use is in the preparation of
ice cream, sherbets, and other frozen desserts. It is the primary stabilizer in most ice
cream products, where it is used to prevent growth of ice crystals (see Chapter 13).
Keeping ice crystals small maintains the smooth, creamy consistency of the product.
CMC also controls sugar crystal size in fondants.

CMC is used when proteins must be stabilized, such as in yogurt, fruit, soy, and
other acidic drinks containing protein. CMC, a polyvalent anion, can stabilize protein
dispersions, especially near their isoelectric pH value where they are least soluble,
because the protein molecules will have multiple positive charges and can bind
to CMC molecules. Using CMC, milk products can be stabilized against casein
precipitation when milk is acidified because CMC forms stable soluble complexes
with casein at pH values in the range 3–6 where casein is insoluble.

Because of its rapid hydration, it affects hydration of other dry mix components. It is
both a binder and an extrusion aid in the preparation of pet foods and other extruded
products. In baked goods, like cakes, CMC is added to adjust the consistency
of the batter, to increase product volume, to improve the quality of the finished
product, to provide moisture retention, and to prolong freshness. Some of the many
applications of CMC are given in Tables 8.2 and 8.4.
Table 8.4. Typical applications of carboxymethylcelluloses

Product types Functions


Beverage mixes Alcohol tolerance
Cake, doughnut, and related mixes Batter thickenerHumectant (improves
texture and extends shelf life)Increases
volume (film former)

Cheese spreads Protective colloid

Dietetic foods ThickenerBodying agent

Dressings Thickener

Dry pet food Makes gravy when water is added

Dry-powder fruit drink mixes and salad Suspending aidRapid hydration


dressings

Extruded products LubricantBinderFilm formerProcessing


aid

Fondants Inhibits sugar crystal growth

Frozen and dried egg white Protein stabilizer

Hot cocoa mixes Thickener

Ice cream and other frozen dessert prod- Inhibits ice crystal growthImproves
ucts mouthfeel, body and textureStabilizes ca-
sein

Icings and frostings Inhibits sugar crystal growthPrevents


moisture loss.

Meat emulsions TexturizerBinder

Milk products Protein stabilizer

Pie fillings Improves texturePrevents syneresis

Puddings Inhibits sugar crystal growth

Sauces Suspending aid

Syrups Transparent thickener for low-calorie


pancake syrup
Toppings Holds moisturePrevents syneresis

There is no limit to the concentration of CMC that may be used in food products,
but it should be used with good manufacturing practices in mind. Ingredient labels
on products containing CMC may read sodium carboxymethylcellulose, sodium
carboxymethyl cellulose, carboxymethylcellulose, carboxymethyl cellulose, CMC,
sodium CMC, or cellulose gum.

Methylcelluloses and hydroxypropylmethylcelluloses


When alkali cellulose is treated with methyl chloride, the protons of some of the
hydroxyl groups of cellulose are replaced by methyl groups (that is, methyl ether
groups are introduced). The products are methylcelluloses (MC). Most members of
this class of hydrocolloids also contain hydroxypropyl ether groups (cellulose-O-CH-
2-CHOH-CH3). Hydroxypropylmethylcelluloses (HPMC) are made by reacting alkali

cellulose with both propylene oxide and methyl chloride. The DS of commercial MC
range from 1.1 to 2.2. The hydroxypropyl MS (moles of substitution [Chapters 4
and 7Chapter 4Chapter 7]) levels in commercial HPMCs are in the range 0.02–0.3.
At this low degree of derivatization, it is unlikely that much, if any, formation of
poly(propylene oxide) [poly(propylene glycol)] chains occurs. Both members of this
hydrocolloid family (that is, both MCs and HPMCs, are generally referred to simply
as MCs) (Scheme 8.3).

MC products are available in a wide range of viscosity types, with the viscosity of 2%
solutions varying from less than 10 to more than 250,000 mPa s. MC products are
cold water soluble because the protrusions along the chains prevent intermolecular
reassociation. They are not readily hydrolyzed by cellulase and, hence, are somewhat
resistant to microbial attack.

A unique and most important property of the MC and HPMC is that of reversible
thermal gelation. Their solutions gel when heated; the gels reliquefy when cooled.
While a few ether groups spread along cellulose molecules enhance water solubility
by prevention of intermolecular associations, they also decrease chain hydration by
replacing water-binding hydroxyl groups with less polar ether groups. The ether
groups, while responsible for water solubility by blocking interchain associations,
restrict hydration of the chains to the point that MC are on the borderline of water
solubility. (Even so MC hydrate rapidly.) In solutions at room temperature, MC and
HPMC chains are hydrated and there is little polymer–polymer interaction, even
through the hydrophobic ether groups. However, when an aqueous solution is heat-
ed, the water molecules solvating the polymer molecules achieve sufficient kinetic
energy that many dissociate from the chain, allowing intermolecular van der Waals
associations (hydrophobic associations) between the ether groups and gelation to
occur when the incipient gelation temperature (gel point) is reached. Gel strength
continues to increase to a maximum as the temperature is held above the gelling
temperature, but the process is completely reversible. The gelation temperature and
gel strength are determined by the specific type of MC. Reducing the temperature
once again brings about rehydration and solubility. Thus, the thermal gelation is
reversible. This property is the basis for many of the applications of MC and HPMC
(Table 8.5).

Table 8.5. Typical applications of methyl- and hydroxypropylmethylcelluloses

Product types Functions


Acidic foods pH stability
Bakery products Gas entrainment (film formation)
Batters and coatings Provides a barrier to fat absorption via thermal
gelation and film formation, provides hot cling
(adhesion)
Deep-fried foods Reduction of oil absorption via thermal gelation
and film formation
Extruded products Lubricant, binder
Frozen foods, desserts, and whipped toppings Provides freeze-thaw stability, maintains tex-
ture and viscosity, inhibits phase separation and
syneresis
High-sugar products Compatibility
Meat analogues Adhesion and binding, lubricant
Pie fillings Reduces boilover, prevents water migration from
the filling to the crust
Salad dressings Emulsifier and emulsion stabilizer (inhibits phase
separation)
Structured foods Adhesion binding
Whipped toppings Foam stabilizer, provides freeze-thaw stability

An interesting aspect of this property is that MC and HPMC are insoluble in hot
water because these and other polysaccharides are insoluble under conditions at
which they form gels; so MC (the generic term that includes HPMC) can be dispersed
in hot water without thickening or clumping. Then, the stirred dispersion can be
cooled to give a clear, smooth solution. In fact, the rate and degree of solubility are
inversely related to the temperature of the water.

Surface activity is another property of MC (generic) products. They share their


emulsifying and emulsion stabilization properties with gum arabic (Chapter 16),
propylene glycol alginate (Chapter 14), and starch octenylsuccinate (Chapter 7). Be-
ing derivatives of linear molecules, they also form films, and MC (generic) products
add lubricity during processing. Low-MW MC (generic) interact synergistically with
hydroxypropylstarch (Chapter 7).

MC (generic) products can be used to reduce the amount of fat in food products
via two mechanisms: (1) they impart fatlike properties so that the fat content of a
product can be reduced, and (2) they reduce absorption of fat by products being
fried. When foods to be battered and deep fried, such as onion rings, are formu-
lated with a MC product in the batter, fat absorption is reduced (because the gel
structure produced by thermal gelation provides a barrier to penetration of fat and
oil) and adhesion is improved (because MC products form continuous films and
have adhesive properties). MCs also function as binders in certain structured foods
because of their thermal gelation, and MC products impart richness (creaminess),
provide structure and body, and generate and stabilize foams. In baked products,
their thermal gelation gives greater gas retention during baking, and they aid in
moisture retention, improve tenderness, and extend shelf life.

Other attributes of MC (generic) products are the provision of hot cling (via thermal
gelation), reduction in boilover (via thermal gelation), prevention of moisture mi-
gration, maintenance of texture (inhibition of phase separations, such as syneresis)
during freeze-thaw cycling, both emulsification and emulsion stabilization, and
tolerance to high concentrations of sugar. The ability to provide lubricity is also
useful in making products by extrusion.

A major use of HPMC products is in nondairy whipped toppings. Its interfacial


activity stabilizes foams and imparts better whipping characteristics, and its emulsi-
fying property prevents phase separation. It provides freeze-thaw stability because it
becomes more and more soluble as the temperature is lowered. MC creates a denser
foam as compared with HPMC.

Acceptable label designations for MC are methylcellulose, methyl cellulose, and


modified cellulose; modified vegetable gum is used occasionally. Acceptable label
designations for HPMC are hydroxypropylmethylcellulose, hydroxypropyl methyl-
cellulose, hydroxypropyl methyl cellulose, modified cellulose, and propylene glycol
ether of MC; carbohydrate gum is used occasionally.

Hydroxypropylcelluloses (HPC) are made by reacting alkali cellulose with propylene


oxide only. Food-grade HPC products are water soluble as long as the water temper-
ature is below about 40°C (105°F). (Like the MC and HPMC, HPC is insoluble in hot
water). As HPC solutions made at a lower temperature are heated to a temperature
above about 40°C (105°F), the HPC molecules precipitate rather than form a gel,
but the principle is the same as in gelation of HPMC and MC. Because HPC
molecules are nonionic, they are compatible with most salts at low concentrations,
but higher concentrations of some salts (for example, 5%–10%) and sucrose (more
than 10%) lower the precipitation temperature. Hydoxypropylcelluloses are on the
edge between water solubility and water insolubility, so anything that competes with
them for the water molecules needed to hydrate the polymer molecules and keep
them in solution will effect precipitation.

It requires much more derivatization to make cellulose molecules water soluble than
it does to stabilize starch molecules, so while the MS of hydroxypropylated starch
does not often exceed 0.20, commercial HPC products have MS values of more
than 2.0. In fact, the legal limit for food use is MS 4.6, which is on average 4.6
hydroxypropyl groups per glucosyl unit. A molecule of HPC with MS 4.6 will contain
62% by weight of hydroxypropyl groups. (This is an excellent example of why MS,
rather than DS, is used when hydroxypropylation is involved.) To achieve MS values
greater than 3, some propylene oxide molecules have to react with the newly added
hydroxyl groups on the added hydroxypropyl ether groups (that is, chaining has to
occur). In fact, the chaining reaction begins long before MS 3.0 is reached because
the hydroxyl group on the hydroxypropyl ether group is more reactive than any of the
hydroxyl groups on glucosyl units and, as already pointed out, the hydroxyl group
on C3 is not very reactive.

Like the more highly substituted HPMC, HPC products are thermoplastic. Like other
cellulose derivatives, HPC can be cast into transparent, water-soluble, edible films.

The label designation may be HPC or hydroxypropyl cellulose.

Ethylmethylcellulose
Cellulose containing both ethyl and methyl ether groups is an approved food
ingredient in the European Community.

> Read full chapter

Cellulose Polymers in Microencapsula-


tion of Food Additives
Dave Wallick, in Microencapsulation in the Food Industry, 2014

17.2.2 Solubility
The substitution type and level of a particular nonionic cellulose derivative deter-
mines to a large extent its aqueous solubility. Cellulose, though a highly hydroxylated
polysaccharide, is swellable but not soluble in water. This is ascribed to high levels
of intramolecular and, to a lesser extent, intermolecular hydrogen bonding between
hydroxyl groups in and between glucopyranosyl rings within a polysaccharide chain
and between adjacent polysaccharide chains in bundled fiber structures of cellulose.

Many commercially available cellulose ethers, such as MC, HPMC, and CMC, are
water-soluble polymers. Commercial MC products have an average DS ranging
from 1.5 to 2.0; commercial HPMC products have an average DS for methyl group
substitution of 0.9 to 1.8 and the MS of hydroxypropyl groups ranges from 0.1
to 1.0. Altering the amounts of methyl and hydroxypropyl substitution also affects
the solubility properties of MC and HPMC cellulose ethers. Decreasing the level
of substituent groups below a DS of 1.4 gives products whose solubility in water
decreases. Increasing the substitution above an MS of 2.0 improves solubility in
polar organic solvents.

MC and HPMC possess the rather unusual property of solubility in cold water and
insolubility in hot water, so that when a solution is heated, a three-dimensional gel
structure is formed (Sarkar, 1995; Sarkar and Walker, 1995). By modifying production
techniques and by altering the ratios of methyl and hydroxypropyl substitutions,
it is possible to produce products whose thermal gelation temperature ranges
from 50 to 90°C (122–194°F) and whose gel texture ranges from firm to rather
mushy. Hydroxypropyl cellulose (HPC) also exhibits cold water solubility and thermal
gelation. It becomes insoluble above about 42°C and is soluble in a broad range of
polar organic liquids (Desmarais and Wint, 1975).

Aqueous solubility of ethylcellulose (EC) depends on its degree of ethylation (DS) and
also on the distribution of ethyl substituent groups on the glucopyranosyl rings. EC
with DS 0.7 to 1.7 is water soluble and exhibits thermal gelation. Selective synthesis
of 3-mono-O-ethylcellulose has been reported (Koschella et al., 2006), resulting in a
positional isomer not readily found in EC prepared by conventional reaction of ethyl
chloride with alkali cellulose. EC from conventional preparation processes typically
possesses about equal proportions of 2- and 6-positional isomers with low partial
DS at position 3 (Koschella et al., 2006). 3-Mono-O-ethylcellulose is soluble in both
aprotic-dipolar organic media and in water, and exhibits thermoreversible gelation
at 58.5°C, about 30°C higher than the cloud point temperature of conventional
water-soluble EC.

Ethylcellulose, once described in pharmaceutical formulation as “probably the most


widely used water-insoluble polymer in film coating due to its good film forming
properties that enable tough, flexible coatings to be produced” (Narisawa et al.,
1994), is beginning to find application in microencapsulation applications in food.
Commercially representative ethylcellulose, of DS >2.0, dissolves in polar organic
solvents such as ethanol or ethyl acetate, and in aqueous-solvent systems (i.e., 85%
ethanol-water or higher). Theoretical values for ethylcellulose–solvent interaction
parameters have been calculated for 112 solvents, and such parameters can be
used to identify useful solvents or solvent–nonsolvent systems for preparation of
microcapsules using coacervation (Robinson, 1989). Commercial EC (DS >2.0) is
thermoplastic, softens at 130°C, flows at 140 to 160°C, and can be used for prepa-
ration of films by melt extrusion (Koschella et al., 2006). Careful manipulation of EC
solubility properties in various solvents has been used to develop coacervation coat-
ing systems. In particular, non solvent addition, emulsion-solvent evaporation, and
temperature reduction-poor solvent methods have been employed to achieve very
effective EC coatings in microencapsulation systems (Nixon, 1985; Chemtob et al.,
1986; Nixon and Wong, 1990; Sveinsson and Kristmundsdottir, 1992; Amperiadou
and Georgarakis, 1995; Jones and Pearce, 1995; Song et al., 2005; Choudhury and
Kar, 2009). Select examples of these coating methods will be indicated described in
Section 17.3. Finally, aqueous dispersions of ethylcellulose are sold into the food
industry and are used widely in the pharmaceutical industry, but have not generally
been adopted by food formulators for microencapsulation coating applications.

Cellulose esters exhibit quite different solubility profiles than do cellulose ethers. For
example, because the phthalate substituent of cellulose acetate phthalate (CAP) is
ionizable in solution above its pKa value, this cellulose derivative is insoluble in acid
media at pH 5 or below, but is soluble at pH higher than 6. This profile is generally
representative of the class of cellulose esters, though particular pKa values for select
esters may vary. Such solubility behavior provides useful cellulose ester coatings for
pH-activated release of microencapsulated active ingredients, such as nutrients and
vitamins, in the gastrointestinal (GI) tract.

> Read full chapter

Processes and Applications for Edi-


ble Coating and Film Materials from
Agropolymers
Hyun Jin Park, ... Ho Jae Bae, in Innovations in Food Packaging (Second Edition),
2014

Cellulose
Cellulose is the most abundant organic renewable resource in the plant kingdom,
and cellulose derivatives (e.g., methylcellulose, hydroxypropylcellulose) have excel-
lent film-making properties. The first cellulose film, cellophane, was developed in
1908, and has been tried as an edible film and coating since the 1980s. Cellu-
lose-based films are very efficient oxygen and hydrocarbon barriers, and their
water-vapor barrier properties may be improved by the addition of lipids. Cellu-
lose-based films have been investigated for controlling the migration of moisture,
gas, and hydrocarbons in various types of foods such as meat products, fruits,
and vegetables. Ayranci et al. (1997) reported that the water vapor permeability of
films decreased with increasing molecular weight of hydroxypropylmethylcellulose
(HPMC). A similar trend has been observed with MC, but only above a molecular
weight of 41,000. WVP values decrease sharply with increasing molecular weight of
PEG up to about 1000 and only a slight increase above this level.
> Read full chapter

Reduced and zero calorie lipids in food


W.E. Artz, ... S.L. Hansen, in Modifying Lipids for Use in Food, 2006

18.2.4 Carbohydrate fatty acid esters


The carbohydrate-fat combination fat substitutes include those derived from poly-
dextrose, sugar alcohols, altered sugars, starch derivatives, cellulose and gums. They
can also be made from components from rice, wheat, corn, oats, tapioca or potato,
and they can replace from 50 to 100 % of the fat in foods (Glueck et al., 1994).

The synthesis and analysis of carbohydrate fatty acid esters have been reported by
several groups (Akoh and Swanson, 1989a,b; Drake et al., 1994; Rios et al., 1994). Car-
bohydrate fatty acid polyesters with a degree of substitution (DS: number of hydroxyl
groups esterified with long-chain fatty acids) of 4–14 are lipophilic, non-digestible,
non-absorbable, fat-like molecules with the physical and chemical properties of
conventional fats and oils and are referred to as low-calorie fat substitutes (Akoh,
1994a; 1995). Swanson has published work from his laboratory on carbohydrate
fatty acid esters synthesized from a variety of carbohydrate sources under a range
of catalytic conditions (Akoh and Swanson, 1990; Akoh, 1994a,b; 1995), including
glucose, sucrose, raffinose, stachyose and verbascose fatty acid esters (Akoh
and Swanson, 1989a). The synthesis of both trehalose octa-oleate and of sorbitol
hexa-oleate (Akoh and Swanson, 1989b) have been reported.

Oleic acid esters of erythritol, pentaerythritol, adonitol and sorbitol were prepared by
transesterification with an excess of methyl oleate to form complete esters (Mattson
and Volpenhein, 1972). The esters formed were erythritol tetra-oleate, pentaerythri-
tol tetra-oleate, adonitol penta-oleate and sorbitol hexa-oleate. These esters were not
susceptible to in vivo lipolysis by lipolytic enzymes of rat pancreatic juice, suggesting
potential application as low-calorie oils (Mattson and Volpenhein, 1972; Akoh and
Swanson, 1989a). Chung et al. (1996) also reported the preparation of a sugar alcohol
fatty acid ester made with sorbitol.

Enzymatic methods for the synthesis of carbohydrate fatty acid esters have been
discussed in detail by Riva (1994). One of the most promising enzymes tested,
particularly for fatty acid esterification of the alkylated glycosides, was a lipase from
the yeast Candida antarctica, which had been immobilized on macroporous resin
beads. Mutua and Akoh (1993) have reported the synthesis of glucose and alkyl
glycoside fatty acid esters in organic solvents using Candida antarctica as a catalyst.
> Read full chapter

Edible Coating and Film Materials


Yachuan Zhang, ... Derek Mclaren, in Innovations in Food Packaging (Second Edi-
tion), 2014

Carbohydrates have been intensively studied to develop edible coatings and films. In
the last decades, starch, chitosan, cellulose derivatives, pectin, and galactomannans
has been evaluated in their coating and film forming ability for applications in
the food packaging area. This chapter reviews the coating and film matrices, the
formation methods, physicochemical properties, and applications in food preserva-
tion of the chitosan, cellulose derivatives, pectin, and galactomannans coatings and
films. Tests on fruits, vegetables, and fish fillets have shown that the shelf life of
these food products has been extended. Also, researchers are spending considerable
time and effort to improve the water vapor barrier properties, which is the major
limitation to these coatings and films, through the use of polysaccharides and lipids,
nanotechnology, and mechanical modification, etc.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like