You are on page 1of 5

www.mrs.

org/publications/bulletin

The Materials
Science of Chocolate
Peter Fryer and Kerstin Pinschower
Introduction
Chocolate is a common confectionery material throughout the world that has seen generally increasing production trends over the last 10 years.1 Making chocolate requires an understanding of how the consumer perceives it. The preferred type of chocolate varies from country to country; for example, common U.S. and U.K. chocolate tastes are mutually incomprehensible, while the rest of Europe hates both of them! The different tastes and uses for chocolate reflect the histories of the industry in different places. The taste of chocolate is partially determined by the chemistry of the product; typical formulations of chocolate are shown in Table I. Whatever type or taste of chocolate, however, the taste experienced by the consumer also depends critically on the micrometer-scale structure of the chocolate, which can consist of crystals and particles ranging from 10 m to 120 m in diameter, depending on the product.2,3 Taste depends on the release of flavor compounds to the mouth and nose, while perceived texture is a function of the way in which the material melts and breaks up in the mouth. This is a materials-science problem; making chocolate involves solving problems that are familiar in other areas of science. Industrial chocolate processing is well developed and includes several complex operations for the development of flavor and texture. Chocolate is first mixed and ground to give a mixture of the correctly sized particles. The process of conching then involves the mixing and shear of the chocolate under controlled conditions and results in the removal of volatile components and adjustments in moisture content and viscosity. This process results in chocolate with the correct flavor profile. The production and structuring of the solid material prior to molding involves a further complex step, that of tempering, in which the chocolate is heated and sheared prior to its final solidification. The complexity of chocolate arises from the polymorphic nature of its constituent fats, which can come in at least five crystal forms. Cocoa butter is chemically a multicomponent mixture of triglycerides and trace compounds.4 Approximately 85% of the composition consists of just three triglycerides: POP ( 20%), POS ( 40%), and SOS ( 25%), where palmitic (P), oleic (O), and stearic (S) acids are the fatty acids attached to the glycerol base. The exact composition depends on factors such as growing conditions and therefore can vary between batches, especially from different geographic regions.5 Milk fats typically consist of one long chain and two short chains. Milk fat is one of the few fats that are compatible with cocoa butter, with which it builds a continuous phase. Addition of milk fats gives milk chocolate, which is sweeter and cheaper than dark chocolate. Examples of the conditions under which the various polymorphs arise are given

in Table II. The melting points are those given by Wille and Lutton,6 who also defined the polymorph numbering system (IVI), which labels the forms in order of increasing melting point. This has now become a standard. However, in chocolate the melting points also depend on composition; for example, the addition of milk solids lowers the melting temperature by up to 10 C, as will be shown later. Extensive work has been carried out to quantify the physical behavior of the triglycerides. Five or six crystalline forms have been identified using experimental techniques such as observing crystals melting under a microscope and, more recently, differential scanning calorimetry (DSC) and x-ray diffraction (XRD) (or a combination of these). Six forms were identified by Wille and Lutton using XRD. They found it difficult, however, to assign x-ray nomenclature to data based on crystal spacing of a multicomponent mixture. The precise number of polymorphs has been the subject of debate; Wille and Lutton considered that Forms V and VI represented distinct but very closely related crystalline structures, and that Form I was a fleeting state that might be a phase mixture. Form III has also been thought to be a phase mixture; Merken and Vaeck,7 supported by SchlichterAronhime and Garti,8 believed Form III to be a mixture of Forms II and IV. Hernqvist9 suggested, on the basis of x-ray data, that only five forms existed, while Talbot10 identified five structured forms and one mixed form, supporting the ideas of Merken and Vaeck. However, the existence of the six polymorphic forms of cocoa butter was confirmed by Loisel,11 who used DSC and XRD as a function of temperature at different scanning rates. Form I showed liquid-crystal form, with some unordered material. This form trans-

Table I: Typical Chocolate Formulations.


Component Cocoa mass (cocoa solids and cocoa butter) Added cocoa butter Sugar Lecithin Flavor compounds (e.g., salt and vanillin) Whole milk powder Total fats Source: From Reference 27. Typical Percentage of Component in Milk Chocolate Dark Chocolate Bittersweet Chocolate 11.8 39.6 60.7

20.0 48.7 0.4 0.1 19.1 31.5

11.8 48.1 0.4 0.1

2.6 36.3 0.3 0.2

36.4

35.4

MRS BULLETIN/DECEMBER 2000

25

The Materials Science of Chocolate

Table II: Overview of Cocoa-Butter Polymorphs.


Polymorph Form I Conditions under which Polymorph Arises Rapid cooling of melt. (Successive polymorphs are then obtained sequentially by heating at 0.5 C min 1.) Cooling of melt at 2 C min. Rapid cooling of melt, followed by storing from several min up to 1 h at 0 C. The form is stable at 0 C for up to 5 h. Solidification of melt at 510 C. Transformation of Form II by storing at 510 C. Solidification of melt at 1621 C. Transformation of Form III by storing at 1621 C. Solidification of melt. Transformation of Form IV. Solvent crystallization. Transformation of Form V (4 months at room temperature). Melting Point ( C) 17.3 Comments

Form II

23.3

Form III Form IV Form V

25.5 27.3 33.8 Forms after tempering, has good gloss and texture; most desirable form. Bloomed chocolate.

Form VI

36.3

Source: From Reference 28.

formed to the more stable Form II on heating. Form III was the most difficult to identify. The lower-melting polymorphs generally form by cooling to a lower temperature and at a faster rate. The lower-melting forms then transform to higher-melting forms with time.12 Form VI is generally associated with fat bloom, an unpleasant-looking appearance of the chocolate surface (shown in Figure 1). It is widely believed that good temper prevents fat bloom to some extent, as does the presence of free milk fat.13 Bricknell and Hartel14 used XRD to study bloom, using different sugars. In all cases, Form V transformed into Form VI, but this was not the cause of visual bloom of the sort seen in Figure 1. The microstructure of the sucrose in the chocolate was considered critical: chocolate made with amorphous sugar did not bloom, while that made with crystalline sugar did. In recent years, importance has been attached to the relevance of the microstructure of the chocolate, which determines many of the physical properties of the material. Narine and Marangoni3 have studied the microstructure of cocoa butter and other fats and identified fractal dimensions from both rheological data and image analysis taken using polarized-light microscopy. The fractal dimension obtained from microscopy is taken as a measure of the spatial distribution of crystals.15 The fractal dimension of the crystal network is considered to be an indicator of hardness; the

normalized elastic modulus for a number of fats is on the same curve. The authors suggest that the fractal dimension is related to conditions of nucleation: systems with a sharp nucleation step in heat flow as the sample crystallizes from the melt will have higher fractal dimensions and a more ordered system. The complexity of chocolates polymorphic system and the crystals that are produced make this an active area of food materials research. The key practical problem for the chocolate maker and for the process engineer is that the thermodynamically stable form (Form VI) is not the one that the consumer finds attractive; making chocolate thus involves ensuring that the final product is in the right form to be sold and eaten. The complexity of the chemistry is such that it is crucial to get the right crystal form. As will be shown later, this involves precise control of the temperaturetimeshear history of the material in a tempering process. If the chocolate is heated to too high a temperature, and not cooled in the correct way, it will set in an untempered form. The types of problems that can result are discussed in the following sections.

coats an inner confection of some kind; and filled chocolates (bonbons, sweets), which consist of a molded chocolate shell surrounding a center. Figure 2 shows schematically the stages involved in making filled chocolates. The first stage is to fill a cold mold with hot melted chocolate and then invert it to leave a shell of the desired thickness. The center material is then poured in to fill the shell. Another layer of chocolate is added above the filling to form what will become the base of the sweet, and finally, the mold is inverted, and the sweet falls out of the mold. A range of problems can occur during this process: incorrect control of the initial process results in a shell that is too thick or too thin, the hot center material can melt or burn through the shell and cause a leak of the filling, the base can fall off if it is not properly welded to the rest of the sweet, or the final product may not demold. Many of these problems are due to the polymorphic nature of the fats. Controlled crystallization (tempering) is thus a vital step in chocolate production. Tempered chocolate (largely in Form V) shrinks on cooling and comes away readily from the mold; untempered chocolate does not. Untempered material is also softer; excessive local heating from the cream filling, or around the join between the two chocolate layers, can cause detempering and weak regions in the final product. Chocolates may then suffer from a process analogous to weld decay in metals. In addition to all of the manufacturing problems that result from the production of the untempered form, this type of chocolate is not attractive to the consumer. Figure 1 shows two bars; the tempered form is glossy to the eye, while the untempered product is spotted and marked, showing strong fat bloom. Untempered chocolate also tastes gritty; the melting point of some of the crystals is above the temperature of the mouth (unlike the tempered form, which melts in the mouth), making its mouthfeel less pleasant. The problems of chocolate manufacture are those of chemical engineering and materials science: to devise a processing method that delivers a product with the right structure.

Tempering Processes and Microstructure


Different types of chocolate are made in different ways. The tempering process is common to all production methods, however. The aim of tempering is to generate sufficient seed crystals of Form V that can subsequently act as points on which the fats can crystallize. In practice, tempering

Making Chocolate: Casting and Solidification


Different sorts of chocolate products are produced commercially: bars or blocks, in which chocolate is poured into molds and set; enrobed products, in which chocolate

26

MRS BULLETIN/DECEMBER 2000

The Materials Science of Chocolate

Figure 1. Good and bad chocolate: The well-tempered chocolate on top is glossy and has come away easily from the mold. The untempered chocolate sticks in the mold and shows a spotted and marked surface typical of strong fat bloom.

Figure 2. Schematic molding process for filled chocolates. (a) The mold is filled with chocolate. (b) Excess chocolate is removed by inversion of the mold. (c) The filling is added. (d) The base of the sweet is added. (e) The completed sweet is removed by inverting the mold again.

involves four steps: (1) complete melting of the chocolate to about 50 C, which removes most or all of the crystalline material; (2) cooling to the point of crystallization; (3) maintaining a holding temperature for crystallization for about a minute; and (4) reheating to melt out unstable crystals. During this process, temperature control is very critical, and the shear applied within the process is also a crucial parameter. The tempering time varies with recipes, equipment, and the purpose of the final product.9 In industry, this is done in a range of different ways. An automatic tempering machine usually consists of multistage heat exchangers through which the chocolate passes. The correctly tempered chocolate is then subjected to the best possible conditions for cooling. This is usually achieved in a series of cooling tunnels using a combination of conduction, convection, and radiant heating. The cooling rate depends on temper, type, and thickness of the chocolate, but it also largely determines the quality: too rapid a cooling rate yields a material that is kinetically trapped into

the wrong form. Figure 3 shows the effective (specific) heat capacity for milk chocolate cooled at different rates, obtained using DSC. This effective (specific) heat capacity includes both specific heat and phase-change effects; the different cooling rates shift the main peak by about 15 C, showing the difference between the tempered material produced at low cooling rates and the untempered material produced at faster rates. We have done some work16 on trying to understand the combination of process steps that produce tempered chocolate. Using a Couette mixer, which consists of a pair of concentric cylinders with the inner one rotating, it is possible to generate a uniform shear on chocolate placed between the two cylinders. The temperature of the device can be controlled and varied; this allows a controlled temperature timeshear pattern to be given to the chocolate. Figure 4 shows the process variables in tempering: the temper time, melt temperature, and shear rate can be varied. As an example of the effects of process conditions on tempering, Figure 5 shows

the DSC trace for milk chocolate processed at different shear rates. The phasechange temperatures are lower than those shown in Table II, since milk chocolate, which has a lower melting point because of the effect of added milk fat on the cocoa butter, was used in all of the experiments. At lower shear rates (027/s), the material melts at 1314 C, showing that it is untempered; at higher shear rates, different forms of chocolate result. The peak shifts to the right with increasing shear; when 53/s is applied, the chocolate melts at 18 C, with a sharp peak. The broader, double peak in the intermediate area shows the existence of different polymorphs in the chocolate. The same type of results can be produced by holding chocolate at different temperatures for different lengths of time: it was found that for this type of chocolate, one needs to hold the material at 22 C for 600 s before warming to melt unstable crystals in order to ensure the temper. Times below 500 s result in badly tempered material and below 300 s, untempered material. The process is very temperature-sensitive: subsequent reheating to a temperature of

MRS BULLETIN/DECEMBER 2000

27

The Materials Science of Chocolate

Temperature (C)
50
at Capacity (kJ/kg C)

40

30

20

10

-10

9 8 7 6 5 4 3 2
under unsuitable conditions, such as at too high a temperature.17 The industrial need is to be able to control the cooling rates of commercial chocolate products such that the wrong crystal form does not result. It has been shown18 that conventional CFD (computational fluid dynamics) modeling programs, using the effective (specific) heat capacity versus temperature plot of Figure 3, can satisfactorily predict the cooling rates of real chocolate. Within the cooling-rate range of commercial chocolate, the data set of Figure 3 can predict the temperaturetime range on cooling; that is, given the temperature and local cooling rate, the specific heat can be put into the model and used predictively. In practice, the surface of Figure 3 results from a composite of thermodynamic and kinetic effects; the heat measured by DSC reflects the extent of the crystallization reaction. As a result of the kinetic effects, rapid cooling results in a lower degree of crystallization at low temperatures than the lower cooling rates. Rapid shifts in the cooling rate thus cannot be accommodated using this approach; a kinetic model for the crystallization of chocolates would be useful. Some approaches to this have been proposed. Rousset19 studied the kinetics of POSSOS crystallization in terms of nucleation and growth kinetics and showed transformations on time scales of both seconds and hours. The effects of solidification were modeled using a conventional materials-science model.20 Van Malssen21 used x-ray powder diffraction to identify kinetic effects at all time scales between seconds and weeks: the solidification takes much longer than the few minutes in the factory process. Some kinetic parameters were also obtained by Metin and
Figure 5. Effective (specific) heat capacity for milk chocolate at different shear rates: all of the chocolates are held at 22 C for 400 s and then rewarmed to 32 C. (From Reference 16.)

) Effective (Specific) He /min

C te (

g Ra oolin C

Figure 3. Effective (specific) heat capacity of milk chocolate as a function of temperature and cooling rate. Data obtained using differential scanning calorimetry. (From Reference 16.)

Figure 4. Model tempering process. Process variables include temper time, melt temperature, and shear rate. (From Reference 16.)

30.5 C instead of 29.5 C is enough to create a material that is not properly tempered. The data suggest that the processes of tempering can be quite well defined.

10

However, even well-tempered and wellprocessed chocolate can transform into the thermodynamically stable but undesirable Form VI if it is stored for a long time or

28

MRS BULLETIN/DECEMBER 2000

The Materials Science of Chocolate

Hartel,22 who used the Avrami equation to model phase transformation, calculating the extent of crystallization from the measured heat evolution in a DSC operated isothermally. Avrami exponents could be fitted to the data, but accurate nucleation and growth data are still lacking.

the application of materials-science techniques in this area will continue.

Acknowledgments
Work at the University of Birmingham on chocolate processing has been sponsored by Cadbury Ltd. and by the Biotechnology and Bioscience Research Council (BBSRC) of the United Kingdom. We wish to acknowledge our co-workers, Andrew Stapley and Heather Tewkesbury.

Discussion and Conclusion


The majority of chocolate is produced via conventional routes, that is, the process is based on mixing, grinding, and conching. The latter process, conching, is necessary for the chemical reactions of flavor development. Due to the polymorphism of the fats, the subsequent tempering stage plays an important role. The need to temper limits the types of chocolate processes that can be designed. Better understanding of the materials science of chocolate, leading to ways of rapidly generating the right crystal forms, would be useful. For example, Beckett23 identifies a range of novel approaches that may in the longer term come into wider-scale commercial use: application of ultrasound (ca. 20 kHz) is found to improve tempering of plain chocolates. Use of a high-shear/lowtemperature crystallizer, such as the device developed by Windhab,24 has been shown to produce tempered chocolate at 4 C in high shear after about 10 s. A number of attempts to extrude chocolate have also been made: Mackley25 and Beckett et al.26 report a novel method for cold extrusion of chocolate in which tempering can be carried out rapidly. Without kinetic understanding, the development of these processes is largely empirical. An understanding of the relationship between the process history and the final quality of the material is critical in any new process. It is possible to model tempering using approximations: a better kinetic model, which could be used predictively, would be a significant advance. The work described here has used a variety of materials-science approaches to study a food-processing problem. We hope that

References
1. C. Nuttall and W.A. Hart, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 443. This book is an invaluable guide to the complexity of chocolate making and the science and engineering involved. 2. G. Ziegler and R. Hogg, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 124. 3. S.S. Narine and A.G. Marangoni, Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 59 (1999) p. 1908. 4. T.R. Davis and P.S. Dimick, J. Am. Oil Chem. Soc. 66 (1989) p. 1488. 5. S. Chaiseri and P.S. Dimick, J. Am. Oil Chem. Soc. 66 (1989) p. 1771. 6. R.L. Wille and J. Lutton, J. Am. Oil Chem. Soc. 43 (1966) p. 491. 7. G.V. Merken and S.V. Vaeck, Lebensm.-Wiss. Technol. 13 (1980) p. 314. 8. J. Schlichter-Aronhime and N. Garti, in Crystallization and Polymorphism of Fats and Fatty Acids, edited by N. Garti and K. Sato (Marcel Dekker, New York, 1988) p. 363. 9. L. Hernqvist, in Crystallization and Polymorphism of Fats and Fatty Acids, edited by N. Garti and K. Sato (Marcel Dekker, New York, 1988) p. 97. 10. G. Talbot, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 216. 11. C. Loisel, G. Keller, G. Lecq, C. Bourgaux, and M. Ollivon, J. Am. Oil Chem. Soc. 75 (1998) p. 425. 12. T. Arishima and K. Sato, J. Am. Oil Chem. Soc. 66 (1989) p. 1614. 13. S.J. Haylock and T.M. Dodds, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 66. 14. J. Bricknell and R.W. Hartel, J. Am. Oil Chem. Soc. 75 (1998) p. 1609.

15. S.S. Narine and A.G. Marangoni, J. Am. Oil Chem. Soc. 76 (1999) p. 7. 16. A.G.F. Stapley, H. Tewkesbury, and P.J. Fryer, J. Am. Oil Chem. Soc. 76 (1999) p. 677. 17. D.J. Cebula and G. Ziegleder, Fett Wiss. Technol. 95 (1993) p. 340. 18. H. Tewkesbury, PhD dissertation, University of Birmingham (1999). 19. P. Rousset, M. Rappaz, and E. Minner, J. Am. Oil Chem. Soc. 75 (1998) p. 857. 20. A. Jacot, M. Swiekost, J. Rappaz, M. Rappaz, and D. Mari, J. Phys. 4 (1996) p. 203. 21. K. van Malssen, A. van Langevelde, R. Peschar, and H. Schenk, J. Am. Oil Chem. Soc. 76 (1999) p. 669. 22. S. Metin and R.W. Hartel, J. Am. Oil Chem. Soc. 75 (1998) p. 1617. 23. S.T. Beckett, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 405. 24. E.J. Windhab, S. Bolliger, and T. Wagner, presented at 4th Eur. Rheology Conf., September 49, 1994, Seville, Spain. 25. M.R. Mackley, UK Patent No. 9,226,477.5 (1993). 26. S.T. Beckett, M.A. Craig, R.J. Gurney, B.S. Ingleby, M.R. Mackley, and T.C.L. Parsons, Trans. Inst. Chem. Eng. 72 (1994) p. 47. 27. K. Jackson, in Industrial Chocolate Manufacture and Use, edited by S.T. Beckett (Blackwell Science, Oxford, 1999) p. 342. 28. J. Schlichter-Aronhime, S. Sarig, and N. Garti, J. Am. Oil Chem. Soc. 65 (1988) p. 1140. s

os

s ib

i l i t y a co m

m u

MRS ONLINE
N E W ! M R S P U B L I C AT I O N S A L E R T
mrs-pubs-alert-subscribe@mrs.org
MRS announces a new service for both its members and the materials research communityMRS Publications Alert. This FREE listserv will provide advance table-of-contents listings for the MRS Bulletin and Journal of Materials Research. Subscribers will receive each alert prior to the availability of the print version. So sign up today. Access the MRS Web site at www.mrs.org or subscribe by e-mail.
ty
o f sci e n t i f i c

ni

un

it y

o f s ci e n ti f i c

po

MRS BULLETIN/DECEMBER 2000

ss

i bi

l i t y a co m

www.mrs.org/publications/bulletin

29

You might also like