Sei sulla pagina 1di 5

Encyclopedia of Food Chemistry: Fat replacers

Michael A Rogers, University of Guelph, Guelph, ON, Canada


© 2018 Elsevier Inc. All rights reserved.

Introduction 1
Carbohydrate-Derived Fat Replacers 1
Protein-Derived Fat Replacers 3
Lipid-Derived Fat Replacers 3
Small Molecule Fat Replacers 4
Conclusions 4
References 4

Glossary
Fat extender used to reduce the concentration of fats by replacing them with a standard lipid combined with other ingredients.
Fat mimetic used to replace fat however it requires a high-water content for to attain its desired functionality.
Fat Substitute used to replace fat on a weight-by-weight basis but is resistant to hydrolysis by digestive enzymes.
Low-calorie fat used to substitute fat with a synthetic fat with one or more of the fatty acids on the glycerol backbone being
replaced with something other than a fatty acid.

Introduction

Reduced fat products are common place in the food landscape of today; consumers often choose these products to control or restrict
caloric intake or to reduce the perceived threat of coronary heart disease (CHD). Messages centering on lipids, especially around
saturated and trans unsaturated, has led a large cohort of consumers to shift their diet towards consuming more low-fat options.
However, the paradigm that all saturated and trans fats have deleterious biological consequences has shifted once again and it is
now being suggested that saturated fat and trans fats from ruminants may not be implicated in CHD as once believed (Chavan
et al., 2016; de-Souza et al., 2015). In practice, because of recommendations to limit saturated fats to no more than 10% and trans
unsaturated fats to no more than 1% of a person’s dietary energy, consumers often choose low-fat options. Although mandates exist
to reduce hardstock fats from foods, a simple fat reduction is associated with undesirable changes to texture and sensory perception
of fat products (Liu et al., 2016b). Up to now, there has been no single fat replacer identified that is able to mimicking the diverse
functionalities of fat across a wide range of foods therefore fat replacers are targeted to specific applications (Table 1).
Replacement of the fat component of foods is often difficult because they have multi-functional roles in foods as they influence
chemical, physical and sensory attributes of foods as well as their processing characteristics. Simply reducing the content of fat
causes quality losses that decrease the acceptance of foods (Peng and Yao, 2017). Therefore, depending on the role that fat has
in determining the quality of these product it will affect the complexity of the system that must be implemented to replace that
lipid fraction (Jones, 1996). Fat replacer is a broad term to describe any technology used to reduce or eliminate fat and can be
subdivided into the following categories (Jones, 1996):
Fat Substitute: used to replace fat on a weight-by-weight basis but is resistant to hydrolysis by digestive enzymes;
Fat mimetic: used to replace fat however it requires a high-water content for to attain its desired functionality;
Low-calorie fat: used to substitute fat with a synthetic fat with one or more of the fatty acids on the glycerol backbone being replaced
with something other than a fatty acid.
Fat extender: used to reduce the concentration of fats by replacing them with a standard lipid combined with other ingredients.
Fat replacers can also be sub-divided based on the primary reason for the removal of fat. Caloric reducers replace a fraction of all
lipids, while hardstock fat replacers only replace a fraction of the trans unsaturated and saturated fats (Fig. 1) (Akoh and Decker,
1995; Rogers, 2011). Caloric reducers include olestra (Bergholz, 1992), non-digestible carbohydrates (Peng and Yau, 2017),
proteins (Singer, 1996), while trans unsaturated and saturated fat replaces are often small molecules such as sorbitan tristearate,
phospholipids, monoglycerides, and modified fatty acids (Rogers et al., 2014, 2009; Wang and Rogers, 2015).

Carbohydrate-Derived Fat Replacers


The molecular diversity of carbohydrates gives rise to various structural and physicochemical properties allowing formulation
chemists to tailor the physical attributes that need replaced upon the removal of fat (de Souza et al., 2015; Peng and

1
2 Encyclopedia of Food Chemistry: Fat replacers

Table 1 Classification of fat replacers by nutrient source and function properties

Type of fat substitute Nutrient source (energy density) Functional properties Use in food

Derived from carbohydrate Water-soluble polymer of Bulking and retaining moisture A wide range of foods, including baked
Polydextrose dextrose (1 cal g1) goods, confections, and frozen
desserts
Modified food starch A variety of starch sources Modifying texture, gelling, Processed meats, salad dressing,
(1–4 cal g1) thickening, and stabilizing baked goods, frozen desserts, etc.
Dextrin and maltodextrins A variety of starch sources Modifying texture and bulking Baked goods, dairy products, salad
(4 cal g1) dressing, sauces, and spreads
Gums and pectin Xanthan, guar, locust bean, Retaining moisture, mouthfeel Wide range of products, including
carrageenan, alginates and and modifying texture baked goods, sauces, and salad
fruit dressings
13-Glucan Various plant sources Modifying mouthfeel, texture, Dairy products
(virtually noncaloric) and pourability
Derived from protein Soluble fiber extracted from Adding body and texture Baked goods and a variety of other
oats (1–4 cal g1) food products
Microparticulated protein Denatured or Modifying mouthfeel Dairy products, spreads, and bakery
and modified whey microparticulated products
protein from egg/milk
(1–4 cal g1)
Derived from fat Sucrose polyester with Modifying texture and mouthfeel Savory snacks (stable for fried foods)
Olestra triglycerides (noncaloric)
Caprenin and salatrim Caprylic, capric, behenic acid, Simulating properties of Confections, baked goods, and
and glycerine, or triglyceride cocoa butter dairy foods
of short- and long-chain fatty
acids (5 cal g1)
Mono- or diglycerides Derived from vegetable oil Adding moisture and modifying Baked goods, vegetable dairy replacers
and emulsified with water texture and mouthfeel
(9 cal g1)

Reproduced with permission from Chavan, R.S., Khedkar, C.D., Bhatt, S., 2016. Fat replacer. In: Caballero, B., Finglas, P. M., Toldrá, F. (Eds.), Encyclopedia of Food and Health.
Academic Press, Oxford, pp. 589–595.

Figure 1 Primary strategies to substitute fats in edible food applications. Reproduced from Rogers, M.A., 2011. Novel lipid substitutes.
In Moo-Young, M. (Ed.), Comprehensive Biotechnology, second ed., vol. 4. Elsevier, pp. 603–616. Reproduced with permission from Elsevier.

Yau, 2017). Furthermore, methods are available to modify carbohydrates; for example, substituted starch using octenyl succinate or
cross-linked starches are commonly used as a fat replacer because of the indigestibility and further caloric reduction. When a fraction
of fat is removed, it is most commonly replaced with carbohydrates such as starch, maltodextrin, gums, or dietary fiber with or
without additional water.
Starch is typically used in its non-gelatinized form because the starch granules are on the same order of magnitude as fat crystals
or fat emulsion droplets which emulates the mouthfeel imparted by lipid droplets (McClements et al., 2017). Because of the size
and physical attributes of starch granules, they are commonly crossed linked to prevent swelling and leaching of the amylose and in
numerous causes amylopectin is debranched (Klaochanpong et al., 2017). Starch based gels have been proven capable of replacing
numerous attributes which fats contribute including: increased water-holding capacity, gel firmness and viscosity and these changes
are often perceived as an increase in creaminess in liquid formulation such as low fat salad dressing (Peng and Yau, 2017).
Encyclopedia of Food Chemistry: Fat replacers 3

Starch can also be partially hydrolyzed using enzymes or acid to form maltodextrin. Maltodextrins with dextrose equivalents
(DE) between 2 and 4 are most commonly used as fat replacers between 1 and 5 wt% because of their ability to gel and thicken
(McPherson and Seib, 1997). At 1 to 5 wt% maltodextrin, 25 to 35 wt% of the fat can be replaced in cookies while maintaining
a full-bodied texture and desirable mouthfeel (Chavan et al., 2016). By selecting the appropriate DE the melting point of the
maltodextrin gel can be tailored to simulate the oral melt down properties of fat (Peng and Yau, 2017). Not only does the DE effect
the ability to mimic the physical properties of fats, but so does the source of the maltodextrin (Peng and Yau, 2017). Maltodextrin
fat replacers are commonly employed in dairy products, confectionary, frozen desserts, cereal baked goods, and meat products
because they form soft, spreadable, thermo-reversible gels with melt-in-the-mouth properties. However, maltodextrins increase
the glycemic load of the food and post meal glycemia, which are less desirable for health (Hofman et al., 2016). There are also
physical limitations in its application due to the amylopectin which has a tendency to retrograde, causing setback in low-fat salad
dressings and their poor freeze–thaw stability and unreliable heat and acid stability (Chavan et al., 2016).
Non-digestible polysaccharides may also be utilized as fat replacers. Microcrystalline cellulose (MCC) is partially depolymerized
cellulose synthesized from a-cellulose precursor. MCC is a modified form of cellulose that is composed mainly of anhydroglucose
units with b(1–4) glycosidic bonds (Chavan et al., 2016). The majority of colloidal grades have microcrystals less than 0.2 mm in
length, that when added to an aqueous medium disperse rather than hydrate (Lucca and Tepper, 1994). MCC is typically processed
with adjuvants such as guar gum which act as a ‘glue’ to hold the insoluble cellulose crystals in a 3D network that is thixotropic
mimicking the body and mouthfeel of lipids. Other non-digestible polysaccharides such as pectin, b-glucan, inulin, locust bean
gum have all shown promise to be used as fat replacers.
Pectin is composed of a galacturonan backbone which may be methyl-esterified giving rise to low and high-methoxy pectin. Pectin
is able to gel and thicken foods; however, because it forms strong gels, there are challenges in incorporating high concentrations of
pectin into foods (Min et al., 2010). The functionality of pectin is influence by the source, which gives rise to variations in molecular
weight and degree of methyl esterification. High methoxy pectin gels in the presence of high sugar concentrations in acidic conditions,
while low methoxy pectin requires the addition of divalent ions such as calcium in basic conditions (Durand et al., 1990). Slendid is
a citrus pectin that when combined with calcium is able to replace up to 100% of the fats in a variety of processed foods (“Low-calorie
fat replacer is based on pectin,” 1991). Recent work with pectin fat replacers has focused on reducing the amount of fat in processed
meats, salad dressings and ice cream (Maneerat et al., 2017; Tekin et al., 2017).
Inulin is a fructosyl-fructose-linked oligomeric carbohydrate with a polymerization degree z10. Inulin is a fructosyl-fructose-
linke d oligomeric carbohydrate with a polymerization degree z10, inulin is a moderately water soluble fructosyl-fructose-linked
oligomeric carbohydrate with a polymerization degree >10 that is resistant to human digestion (Zahn et al., 2010). Inulin forms
a short spreadable gel network, when added at high concentrations, that acts as fat replacer in spreads and fillings, dairy products,
desserts and dressings (Franck, 2007; Guven et al., 2005). In the case of adding inulin to muffins, only 50% of the fat could be
replaced with similar crumb firmness as the full fat counterpart (Zahn et al., 2010).

Protein-Derived Fat Replacers


The ability to mimic fat functionality by microparticulated proteins is attributed to their spherical shape and small size that are
comparable in size to oil droplets found in food emulsions. There are numerous microparticulated protein based fat replacers
that are derived from whey, egg, and zein proteins (Cheftel and Dumay, 1993). Often, the microparticulated protein is applied
in combination with a stabilizing polysaccharide such as carboxymethylcellulose with zein protein and xanthan with egg white
protein (Liu et al., 2016b). Specifically, microparticulated proteins effectively replace the sensory perception of oil droplets in liquid
and semi-solid food matrices (Liu et al., 2016a). Microparticulated whey protein has numerous application in dairy-based products
where it has been shown to provide a smooth, creamy mouthfeel in low-fat applications such as low fat cheeses, yoghurts, ice creams
(Cheftel and Dumay, 1993; Liu et al., 2016a, 2016b). In the case of microparticulated whey protein, the attributes of these
ingredients are dependent on the size of the particles. Small particles, contribute to the creaminess, while large particle impart
a roughness which suppress creaminess (Liu et al., 2016a).
Microparticulated proteins induce a creamy oral sensation when they are dispersed into an aqueous phase because of the
‘ball-bearing’ lubrication properties of microparticulated whey proteins in liquid and semi-solid food matrices (Liu et al.,
2016a). The ‘ball-bearing’ lubrication model suggests that creaminess arises when the microparticulated protein is dispersed in
the aqueous phase allowing them to roll freely over one another under an applied shear. Although microparticulated proteins
increase the creaminess by reducing the friction via the ‘ball-bearing’ model, the underlying mechanism differs considerably
from how oil droplets impart creaminess. These differences may induce changes in sensory perception of creaminess.

Lipid-Derived Fat Replacers


The most commonly thought of lipid-derived fat replacer is olestra, which is a mixture of hexa-, hepta- and octa-esters of sucrose. Six
to eight long-chain fatty acids are covalently linked to sucrose by chemical transesterification. Gastric lipases and the bacteria lipases
produced in the colon are incapable of cleaving the fatty acids from the sucrose ester and as such the fatty acids are not absorbed by
the microvilla but instead are excreted in the stool (Jandacek, 1982). Since olestra is lipophilic, nondigestible, nonabsorbable, it can
interfere with the absorption of other lipophilic components in the diet, consumed at the same time as olestra. An advantage of
olestra over other fat replacers is that the physical properties can be tailored depending on the fatty acids esterified to sucrose.
4 Encyclopedia of Food Chemistry: Fat replacers

For example, if highly unsaturated fatty acids are esterified to sucrose than the resulting olestra will be fluid, conversely, if long chain
saturated fatty acids are used than it will be solid (Ognean et al., 2006). Olestra has very similar organoleptic and thermal properties
compared to triglycerides with similar fatty acids.
Caprenin is a triglyceride that contains glycerol esterified to two medium chain fatty acids (caprylic and capric) and a long chain
fatty acid, behenic acid. The reduction in calories occurs because behenic acid is only partially absorbed, while capric and caprylic
acids are readily metabolized and as such caprenin provides only 5 kcal/g (Akoh, 1998). Because of the fatty acid composition, the
functional properties are similar to cocoa butter allowing it to be used in soft candy and confectionary coatings. Salatrim also relies
on the fact that short chain fatty acids provide fewer calories per unit weight and that long chain fatty acid, stearic acid, is only partly
absorbed resulting in 55% fewer calories than traditional fats (Akoh, 1998). Salatrim is a generic name for triglycerides that contain
at least one short chain fatty acid and at least one long chain fatty acid, typically stearic acid. The physical properties of salatrim can
be manipulated by chaining the amounts of short chain and long chain fatty acids allowing the melting points and hardness to be
manipulated. It is sold as Benefat 1. Its primary application is as a replacer for cocoa butter.

Small Molecule Fat Replacers


Small molecule fat replacers are used to replace the hardstock fats, i.e. trans and saturated fats, while maintaining the concentration
of unsaturated oils and the original physical properties of the fat blend. These liquid oil gels, termed oleogels, molecular gels and/or
organogels, are similar to fats in that they can undergo the sol–gel transition numerous times simply by reheating and recooling the
system (Weiss, 2014). Highly effective oleogelators form gels in vegetable oils, at concentrations as low as 0.5 wt% (Rogers, 2017).
Oleogels self-assemble, in vegetable oils, via non-covalent interactions (i.e. hydrogen bonding, p-p stacking, electrostatic and van
der Waals interactions) forming fibrillar or platelet crystals (Patel and Dewettinck, 2016; Rogers et al., 2009). These intermolecular
interactions favor 1-dimensional (1D) fibrillar growth leading to the formation of a 3 dimensional (3D) networks that is capable
of entrapping the liquid oil phase. Few applications of oleogels in foods exist because it is difficult to identify new food grade,
inexpensive gelators; however, research in this area is increasing exponentially. Although there is a plethora of reported small
molecular gelators, here we will focus only on two systems – the waxes and the phytosterol/g-oryzanol system.
Although numerous waxes have been studied as potential oleogelators to replace hardstock fats, candelilla wax (CDW),
carnauba wax (CBW), rice bran wax (RBW) and beeswax (BW) are of the greatest interest as food-grade waxes for use as edible
oleogels (Rogers et al., 2014). Oleogels comprised of wax and vegetable oils have varying fractions of n-alkanes, fatty alcohols,
and fatty acids. Wax esters have a tendency to form platelet-like or needle-like crystals in edible oils at low concentrations that
are very effective at gelling vegetable oil (Toro-Vazquez et al., 2013, 2010). CDW is obtained from the leaves of Euphorbia cerifera
and comprises n-alkanes, between 29 and 33 carbons and structures safflower and canola oil at concentrations less than 2 wt%
(Blake et al., 2014). CDW forms the most elastic gel (i.e., highest hardness value) while BW is the lowest (Lim et al., 2016).
b-Sitosterol and g-oryzanol are unable to gel vegetable oil when added individually, however they are able to self-assemble
efficiently enough to entrain oil when combined at a 60:40 wt% g-oryzanol/b-sitosterol ratio. At a 1:1 M ratio they self-
assembled into hollow tubules, that have a diameter between 67 and 80 Å, are between 8 and 12 Å thick and have a length which
can exceed 1000 Å, at concentrations as low as 2% (Bot et al., 2012). The ratio of g-oryzanol to b-sitosterol and vegetable oil effects
the self-assembled structures which in turn alters their ability to entrain liquid oil. For this system to gel it requires the molecules to
form synergistic hydrogen bonding, they must contain a ring system, and alkyl residue (Bot and Agterof, 2006). Not only does
b-sitosterol form effective gels with g-oryzanol but so do other plant sterols such as ergosterol, stigmasterol, cholesterol, cholestenol
(Bot et al., 2012). In sunflower oil, the gels are transparent and by changing the concentration, the mechanical and rheological
properties can be fine-tuned to mimic saturated fats. These oleogels are food grade and phytosterols lower blood cholesterol levels
making these attractive alternatives to colloidal fat crystal networks.

Conclusions

The physical attributes that fats impart to foods are very diverse. In emulsions they provide enhanced viscosity, mouthfeel, and body
to the food, while in solid fat matrices they provide an elastic network essential for the food structure. Depending on the type of
food, which fat is being removed from, there are numerous fat replacers that can be used depending on the physical attributes that
need replaced. Fat replacers that work well in emulsions, such as microparticulated whey proteins, will not work as a hardstock fat
replacers, therefore it is important to identify a suitable fat replacer for the specific application or food that is being formulated.
Although numerous fat replacers exist to create low-fat options, there are very few oleogelators that are allowed in foods that
can effectively structure unsaturated oils. However, this is a rapidly growing area of research as there is a constant desire by
manufactures to eliminate trans fats while also reducing the concentration of saturated fats and maintaining unsaturated oil
concentration.

References

Akoh, C.C., 1998. Fat replacers. Food Technol. 52, 47–52.


Akoh, C.C., Decker, E.A., 1995. Lipid-based fat substitutes. Crit. Rev. Food Sci. Nutr. 35 (5), 405–430.
Bergholz, C.M., 1992. Safety evaluation of olestra, a nonabsorbed, fatlike fat replacement. Crit. Rev. Food Sci. Nutr. 32 (2), 141–146.
Encyclopedia of Food Chemistry: Fat replacers 5

Blake, A.I., Co, E.D., Marangoni, A.G., 2014. Structure and physical properties of plant wax crystal networks and their relationship to oil binding capacity. J. Am. Oil Chemists Soc.
91 (6), 885–903.
Bot, A., Agterof, W.G.M., 2006. Structuring of edible OIls by mixtures of g-oryzanol with b-sitosterol or related phytosterols. J. Am. Oil Chem. Soc. 83, 513–521.
Bot, A., Gilbert, E.P., Bouwman, W.G., Sawalha, H., den Adel, R., Garamus, V.M., Venema, P., van der Lindend, E., Floter, E., 2012. Elucidation of density profile of self-assembled
sitosterol þ oryzanol tubules with small-angle neutron scattering. Faraday Discuss. 158, 223–238.
Chavan, R.S., Khedkar, C.D., Bhatt, S., 2016. Fat replacer. In: Caballero, B., Finglas, P.M., Toldrá, F. (Eds.), Encyclopedia of Food and Health. Academic Press, Oxford,
pp. 589–595.
Cheftel, J.C., Dumay, E., 1993. Microcoagulation of proteins for development of “creaminess”. Food Rev. Int. 9, 473–502.
de Souza, R.J., Mente, A., Maroleanu, A., Cozma, A.I., Ha, V., Kishibe, T., Uleryk, E., Budylowski, P., Schünemann, H., Byene, J., Anand, S.S., 2015. Intake of saturated and trans
unsaturated fatty acids and risks of all cause mortality, cardiovascular disease, and type 2 diabetes: a systematic review and meta-analysis of observational studies. BMJ 351.
Durand, D., Bertrand, C., Clark, A.H., Lips, A., 1990. Calcium-induced gelation of low methoxy pectin solutions d thermodynamic and rheological considerations. Int. J. Biol.
Macromol. 12 (1), 14–18.
Franck, A., 2007. Technological functionality of inulin and oligofructose. Br. J. Nutr. 87 (S2), S287–S291.
Guven, M., Yasar, K., Karaca, O.B., Hayaloglu, A.A., 2005. The effect of inulin as a fat replacer on the quality of set-type low-fat yogurt manufacture. Int. J. Dairy Technol. 58 (3),
180–184.
Hofman, D.L., van Buul, V.J., Brouns, F.J.P.H., 2016. Nutrition, health, and regulatory aspects of digestible maltodextrins. Crit. Rev. Food Sci. Nutr. 56 (12), 2091–2100.
Jandacek, R.J., 1982. The effect of non-absorbable lipids on the intestinal absorption of lipiophiles. Drug Metab. Rev. 13, 695–714.
Jones, S.A., 1996. Fundamental issues. In: Roller, S., Jones, S.A. (Eds.), Handbook of Fat Replacers. CRC Press, New York, NY, pp. 3–24.
Klaochanpong, N., Puncha-arnon, S., Uttapap, D., Puttanlek, C., Rungsardthong, V., 2017. Octenyl succinylation of granular and debranched waxy starches and their application in
low-fat salad dressing. Food Hydrocoll. 66 (Suppl. C), 296–306.
Lim, J., Hwang, H.-S., Lee, S., 2016. Oil-structuring characterization of natural waxes in canola oil oleogels: rheological, thermal, and oxidative properties. Appl. Biol. Chem. 60,
17–22.
Liu, K., Stieger, M., van der Linden, E., van de Velde, F., 2016a. Effect of microparticulated whey protein on sensory properties of liquid and semi-solid model foods. Food Hydrocoll.
60 (Suppl. C), 186–198.
Liu, K., Tian, Y., Stieger, M., van der Linden, E., van de Velde, F., 2016b. Evidence for ball-bearing mechanism of microparticulated whey protein as fat replacer in liquid and
semi-solid multi-component model foods. Food Hydrocoll. 52 (Suppl. C), 403–414.
Low-calorie fat replacer is based on pectin, 1991. Chem. Eng. News Archive 69 (41), 26.
Lucca, P.A., Tepper, B.J., 1994. Fat replacers and the functionality of fat in foods. Trends Food Sci. Technol. 5, 12–19.
Maneerat, N., Tangsuphoom, N., Nitithamyong, A., 2017. Effect of extraction condition on properties of pectin from banana peels and its function as fat replacer in salad cream.
J. Food Sci. Technol. 54 (2), 386–397.
McClements, D.J., Chung, C., Wu, B.-C., 2017. Structural design approaches for creating fat droplet and starch granule mimetics. Food Funct. 8 (2), 498–510.
McPherson, A.E., Seib, P.A., 1997. Preparation and properties of wheat and corn starch maltodextrins with a low dextrose equivalent 1. Cereal Chem. 74, 424–430.
Min, B., Bae, I.Y., Lee, H.G., Yoo, S.-H., Lee, S., 2010. Utilization of pectin-enriched materials from apple pomace as a fat replacer in a model food system. Bioresour. Technol. 101
(14), 5414–5418.
Ognean, C.F., Darie, N., Ogenan, M., 2006. Fat replacers - review. J. Agroalimentary Process. Technol. 12, 433–442.
Patel, A.R., Dewettinck, K., 2016. Edible oil structuring: an overview and recent updates. Food Funct. 7, 20–29.
Peng, X., Yao, Y., 2017. Carbohydrates as fat replacers. Annu. Rev. Food Sci. Technol. 8 (1), 331–351.
Rogers, M.A., 2011. Novel lipid substitutes. In: Moo-Young, M. (Ed.), Comprehensive Biotechnology, second ed., vol. 4. Elsevier, pp. 603–616.
Rogers, M.A., 2017. Advances in edible oleogel technologies – a decade in review. Food Res. Int. 97, 307–317.
Rogers, M.A., Strober, T., Bot, A., Toro-Vazques, J.F., Stortz, T., Marangoni, A.G., 2014. Edible oleogels in molecular gastronomy. Int. J. Fo Gastron. Food Sci. 2, 22–31.
Rogers, M.A., Wright, A.J., Marangoni, A.G., 2009. Oil organogels: the fat of the future? Soft Matter 5 (8), 1594–1596.
Singer, N.S., 1996. Microparticulated proteins as fat mimetics. In: Roller, S., Jones, S.A. (Eds.), Handbook of Fat Replacers. CRC Press, Boca Raton, FL.
Tekin, E., Sahin, S., Sumnu, G., 2017. Physicochemical, rheological, and sensory properties of low-fat ice cream designed by double emulsions. Eur. J. Lipid Sci. Technol. 119 (9).
1600505-n/a.
Toro-Vazquez, J.F., Mauricio-Pérez, R., González-Chávez, M.M., Sánchez-Becerril, M., Ornelas-Paz, J.J., Pérez-Martíneza, J.D., 2013. Physical properties of organogels and water
in oil emulsions structured by mixtures of candelilla wax and monoglycerides. Food Res. Int. 54, 1360–1368.
Toro-Vazquez, J.F., Morales-Rueda, J., Mallia, V.A., Weiss, R.G., 2010. Relationship between molecular structure and thermo-mechanical properties of candelilla wax and amides
derived from (R)-12-Hydroxystearic acid as gelators of safflower oil. Food Biophys. 5 (3), 193–202.
Wang, T.-M., Rogers, M.A., 2015. Biomimicry – an approach to engineering oils into solid fats. Lipid Technol. 27, 175–178.
Weiss, R.G., 2014. The past, present, and future of molecular gels. What is the status of the field, and where is it going? J. Am. Chem. Soc. 136, 7519–7530.
Zahn, S., Pepke, F., Rohm, H., 2010. Effect of inulin as a fat replacer on texture and sensory properties of muffins. Int. J. Food Sci. Technol. 45 (12), 2531–2537.

Potrebbero piacerti anche