You are on page 1of 7

Journal of Food Engineering 78 (2007) 11671173 www.elsevier.

com/locate/jfoodeng

Water transport in bread during baking


Muriel J. Wagner
a

a,b

, T. Lucas

a,*

, D. Le Ray a, G. Trystram

, F-35044 Rennes Cedex, France CEMAGREF, Food Processes Engineering Research Unit, 17 av. de Cucille b ENSIA, Food Engineering Research Unit, 1 avenue des Olympiades, F-91744 Massy Cedex, France Received 11 June 2005; accepted 19 December 2005 Available online 6 March 2006

Abstract The main aim of this study was to obtain data on local water content in bread during baking. Breads were baked for dierent times between 2 min and 60 min at 182 C. At the end of each baking run, samples were taken at dierent levels in the bread to assess water content proles. The results showed an increase in water content in the bread core at the very beginning of baking. However the core water content was the same as in raw dough after 7 min of baking. These results were compared to previously published results, and the mechanisms of evaporationcondensation which are responsible for such local increases are dened. 2006 Elsevier Ltd. All rights reserved.
Keywords: Transport; Water content; Bread; Evaporationcondensation; Baking

1. Introduction Measuring and understanding water distribution during baking is fundamental to better control of nal bread quality. Water content and its distribution govern textural properties such as softness of crumb, crispness of the crust and shelf-life. Water also plays an important role in the major physical changes (e.g. expansion of bubbles) and chemical changes (e.g. starch gelatinisation) that take place during bread making. Heat transport towards the centre during baking induces water transport and therefore a change in the water distribution. Two directions of water transport occur concomitantly during the process: towards the thermal centre of the product and towards the surface. As soon as the product is placed in the oven, water evaporates from the warmer region, absorbing latent heat of vaporisation and the surface layers start drying. Beneath this drying region, water vapour diuses through the interconnected pores towards the surface, under the inuence of

Corresponding author. Fax: +33 223482115. E-mail address: tiphaine.lucas@cemagref.fr (T. Lucas).

the water vapour concentration gradient. A concomitant liquid water gradient is formed and ensures the diusive transfer of water from the core to the surface. As the diusive ow of liquid water from the core is less rapid than evaporation ow at the surface, a drying zone is developed, which slowly increases in thickness and forms the crust. Both Zanoni and Peri (1993) and Lostie et al. (2002) characterized this transport; bread crust thickness was evaluated at dierent baking times (Zanoni & Peri, 1993), and water content in a sponge cake crust was measured after sampling at dierent baking times and the drying rate was calculated (Lostie, Peczalski, Andrieu, & Laurent, 2002). Moreover, the temperature gradient inside the bread induces a partial water vapour pressure gradient, via the saturating water vapour pressure. Beneath the drying region, water vapour migrates through the gas phase towards the thermal core of the bread, where the partial water vapour pressure is at its lowest. Such transport occurs according to the Watt-principle or heat pipe. Because the water migrates counter to the thermal gradient, it is accompanied by condensation. This is why this mechanism is also called evaporationcondensation. Henry proposed the theory of evaporation and condensation in 1939 (cited by

0260-8774/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.jfoodeng.2005.12.029

1168

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

De Vries, Sluimer, & Bloksma, 1989). This mechanism can be divided into four stages (De Vries et al., 1989): 1. Firstly water evaporates on the warmer side of the gas cell, absorbing latent heat of vaporisation. 2. Vapour migrates through the gas phase, induced by the vapour concentration gradient inside the cell. 3. Vapour then condenses on the colder side of the gas cell, setting free its latent heat. 4. Liquid water is transported by diusion through the dough membrane to the warmer side of the next gas cell, where the series of processes can start again. When the gas phase becomes continuous, the fourth stage no longer exists. This mechanism takes place as long as a temperature gradient remains (and water remains!) and the core temperature has not reached 100 C. Thorvaldsson and Skjo ldebrand (1998) reported that this mechanism occurs only when the gas phase is continuous, in other words when the dough has already been transformed into crumb, with pore interconnectivity (only stages 13). In addition, on its way towards the centre, the water vapour is believed to be stopped by the interface between the crumb (open structure) and the dough (closed structure) and not to move any deeper. The authors stressed the high probability that this water is absorbed by starch gelatinisation taking place just beneath this crumb/dough interface. Only when the core has changed into crumb can the water move freely and water content rise in the centre. The principle of evaporation and condensation was proposed by Sluimer (cited by De Vries et al., 1989) to explain rapid heat transport during baking. De Vries et al. (1989) demonstrated that conduction alone could not explain the rapidity of heat transport during baking. In theory, the thermal diusivity in a foam equals the thermal diusivity of the continuous phase, multiplied by a factor equal to or smaller than unity, taking into account the presence of insulating elements (gas bubbles) in the continuous phase. If conduction is the only mechanism, heat transport in a gas free dough would then be as rapid as or faster than transport in fermented dough pieces. De Vries et al. (1989) recorded the temperature during baking in the centre of gas free and fermented doughs with the same external dimensions, and showed that heat transport in fermented dough was much faster than in gas free dough. Such results could be only explained by the contribution of the evaporation condensation mechanism to faster heat transport. In practical terms, although the overall water loss is easy to assess (Hasatani, Arai, Harui, & Itaya, 1992; Lostie et al., 2002), the determination of water content proles in crumb during baking requires the setting up of a method for accurate cutting of the loaf into samples while still hot and deformable, and the minimization of water losses by evaporation from cut surfaces. De Vries et al. (1989) determined water content in the centre of the bread directly after baking, and after cooling to ambient temperature. The bread was sliced and a sample was taken from the loaf cen-

tre to determine the water content after oven drying. The number of replicates or repetitions was not given. The water content measured after cooling (43.79%) was very close to the initial water content in dough (43.69%), whereas an increase of 3.5% was measured directly after baking. Constant water contents were also observed in three samples taken at the core 10 min after removal from the oven (Marston, 1976). Czuchajowska, Pomeranz, and Jeers (1989) waited over an hour after baking before removing crumb samples and the nal water content was 23% lower than the initial water content in dough (between 46.39% and 48.88%). The latter results demonstrate that further changes leading to water redistribution occur during the cooling phase, rendering delicate the detection of an increase in water content dicult due to the evaporationcondensation process. Zanoni and Peri (1993) cut bread slices into 27 pieces directly after baking and samples were immediately placed in separate polyethylene bags, frozen in liquid nitrogen and stored at 20 C. Water content was measured in triplicate at dierent locations in the loaf and at dierent baking times between 0 and 30 min. Only the crust was aected by dehydration, whereas the value of water content in the crumb beneath the crust was the same as in raw dough (44%). Even when sampling took place just after removal from the oven, the results were ambiguous with either constant or increasing (De Vries et al., 1989) water content at core. In addition to the conventional, destructive method of sampling and oven-drying, other continuous water content measurement techniques have been proposed. The microwave technique can be used to measure water content through the dependence of an electromagnetic eld on the dielectric and power dissipation properties of water. Another continuous method based on near infrared (NIR) reectance was proposed by Thorvaldsson and Skjo ldebrand (1996). To measure the local content of a sample, light was sent into the sample through an optic bre and the intensity of the reected light was measured. The instrument is also known to be sensitive to the structure and temperature of the sample, and dependence on temperature can be incorporated into the calibration. Structural dependence makes quantitative measurements in the sample dicult for inhomogeneous materials. The bre optic method is better suited for dynamic studies when the probe is placed in a xed environment during the entire measurement period and this is why the authors did not interpret ndings during the oven rise. Thorvaldsson and Skjo ldebrand (1998) showed an increase in water content in the core immediately after the local temperature had reached 70 5 C (i.e. after expansion was more or less complete). The water content of the dough was 41.89% and the average water content in the centre at the end of baking (35 min at 225 C) was 49.6% 5.4. According to the above analyses, water content measurement must ideally be determined directly after baking, in order to minimise experimental bias. Moreover, avail-

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

1169

able data mostly characterise the end of bread baking and are often contradictory. Attempts to monitor dynamically the local water content in bread during baking has failed to show any increase in water content at the core (Zanoni & Peri, 1993) or characterise the rst part of the baking process (Thorvaldsson & Skjo ldebrand, 1998). The aim of this study was to present additional water content data over the entire duration of the baking process. Ideally, a non-destructive technique should be used. However, previous studies using NIR (Thorvaldsson & Skjo ldebrand, 1998) or MRI (Wagner et al., 2003) showed that such methods had to be improved before being able to convert the out-coming signal into water content. We therefore chose to cut bread samples at dierent baking times, while emphasising the number of replicates and repetitions. The present ndings were compared to those reported in the literature, and also to previously formulated theories of evaporationcondensation. This last objective was supported by parallel measurements of temperature inside the bread. 2. Materials and methods 2.1. Sample preparation A recipe for white bread was used in the study with the following quantities and ingredients: 300 g wheat our, 225 g water, 6 g saccharose, 6 g salt, 9 g compressed yeast, 6 g rape seed oil. Flour was stored at 20 C between experiments and defrosted at 4 C 24 h before experiments. Two series of water content measurements were performed at a six months interval. During the rst series, the water content of our was 13.75% and in the second series it was 14.46%, making the initial water content in dough vary slightly. The ingredients were mixed in a mixer adapted from a Chopin alveograph at 35 rpm for 5 min and at 90 rpm for 12 min. The nal temperature of the dough was 26 1 C. Four dough samples were used for the determination of initial dry matter. The dough was placed in a rectangular glass mould (250 50 60 mm), that was coated with Teon to reduce adherence. The dough was proved for 1h10 in a proving chamber with saturated air at 27 C. The lateral walls of the mould were externally insulated with a ceramic layer before baking. In the case

of temperature measurements, optic bres were placed vertically at dierent levels inside the dough (y-axis) before baking. 2.2. Experimental protocol An oven specically designed for an MR imager was used as described in detail by Wagner et al. (2003) and operating conditions were: air temperature 182 3 C and mean air speed 3 m/s. Thermal insulation on the wall sides of the mould insured one-dimensional heat transfer along the vertical axis (y-axis) and mass transfer was expected to behave similarly (except at the interface on the lateral sides). The mould was placed in the oven at t0 and temperatures were recorded over 45 min. For measurement of water content, the baking process was interrupted at times (t) 2, 4, 7, 15, 20, 25, 30, 40 and 60 min. Three slices 1 cm thick were cut at positions 80, 100, 120 mm along the z-axis, while the loaf was still hot. A vertical piece of only 1.5 cm thick was sampled at the mid-width of each slice. Each piece was cut out into seven samples as shown in Fig. 1(a) for determination of dry matter and the vertical position of each sample was measured. Particular attention was paid to rapidity of cutting and weighing of samples by all three operators. As shown in Fig. 1(b), only three regions could be distinguished at t < 7 min: top, centre and bottom. The top and bottom regions were already crumb and no crust could be distinguished: three samples along the z-axis were removed from each of these regions. The core was still liquid dough, from which three samples were taken. At t = 7 min, four samples were taken along the z-axis: three in the same regions as described above and one sample from the crust, which was a thin dehydrated layer. Baking times of 4, 7 and 20 min were repeated three or four times and 40 min was repeated twice. 2.3. Measurements and expression of results 2.3.1. Temperature Temperatures in the loaf were measured with optic bres (diameter 1 mm, accuracy 1 C). Fibres were positioned at dierent levels in the loaf: 0, 5, 20, 30 and 40 mm from the bottom of the mould and along the vertical axis (y-axis). Temperatures were collected using a data acquisition

50 mm top 6 mm

y i+1

core 60 mm 55 mm 39 mm bottom 10 mm

yi

x 0

50 mm

Fig. 1. Sample positioning depending on the baking time: (a) for t > 10 min; (b) for t < 10 minvalues given apply to t = 4 min.

1170

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

Luxtron 790 connected to a PC and recorded every 5 s. Temperatures at 20 mm were repeated four times. 2.3.2. Height The total height of the loaf (Y expressed in millimetres) was measured with a graduated ruler (1 mm) before cutting, then samples were cut out and placed on graph paper to measure their respective heights (Dy). The heights of the top and bottom crusts were measured using a slide calliper. The vertical position yi attributed to each sample was centred at mid-height, with the bottom of the loaf as reference (y = 0). Given y1, the height of the sample near the base of the loaf was y1 Dy 1 2 1

y* = 0.05, 0.49 and 0.99, and their respective water content was compared to those at positions y* = 0.03, 0.48 and 0.97. For t = 4 min, the positions of the three regions were y* = 0.09, 0.47 and 0.94 and their respective water content was compared to those at positions y* = 0.14, 0.48 and 0.97. For t = 7 min, the positions of the four regions were y* = 0.08, 0.44, 0.86 and 0.96 and their respective water content was compared to those at positions y* = 0.14, 0.48, 0.85 and 0.97. 2.3.3. Dry matter and water content Samples were weighed (m), placed in a drying oven at 104 C for 24 h and weighed again (mdry). The results were expressed in grams of water per 100 g of dough at t0 or per 100 g of bread at t: m m  dry x 100 4 m The water content values obtained at the same level in the three dierent slices (along the z-axis) were averaged at given baking times, accompanied by the corresponding standard deviation. 3. Results Fig. 2 presents water content values of samples at dierent levels inside the loaf during baking. As previously reported (e.g. Zanoni & Peri, 1993), dehydration aected only the supercial layers (crust) at the top and bottom, whereas the crumb water content remained constant and equal to that measured in the unbaked dough. However, at 4 min (all three runs) and at 7 min (two runs out of three), the water content at the core (dough) was signi-

For 2 6 i 6 7: y i y i1 Dy i1 Dy i 2 2 2

Because the bread did not reach its maximum height during baking (especially at the beginning), the heights in each loaf were standardized to compare water content proles with each other: yi y 3 i Y It must be emphasized, however, that such comparison assumes uniform expansion of the bread. Seven common regions were thus dened, whose average corresponding mid-heights were y* = 0.03 0.01, 0.14 0.03, 0.33 0.02, 0.48 0.03, 0.67 0.02, 0.85 0.01 and 0.97 0.02. For t = 2 min, the positions of the three regions were

50 45 40

Water content (%)

35 30 25 20 15 10 0
y* = 0.03

51 50 49 0 3 6 9 12 15

10
y*= 0.14

15
y*= 0.33

20
y*= 0.48

25

30
y*= 0.67

35
y*= 0.85

40

45
y*= 0.97

50

55

60

Time (min)
initial water content average

Fig. 2. Water content during baking at dierent levels y i inside the loaf (data acquired during the rst series of baking runs). Water content measurement was repeated three times at 7 and 20 min and twice at 40 min.

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

1171

cantly higher than the initial water content (0.41.3%, see Table 1). No signicant increase in water content at the core could be measured before 4 min (see t = 2 min in Table 1). The order of magnitude of the increase in water content was slightly lower than values reported in the literature (8% (De Vries et al., 1989) to 18.7% on average (Thorvaldsson & Skjo ldebrand, 1998)). This can be explained by the dierence in bread geometry as the thermal gradient, i.e. the driving force for water transport towards the core, is more intense and persistent for a larger-sized product 580 g of dough in a rectangular tin of 9 25 6 cm dimensions (Thorvaldsson & Skjo ldebrand, 1998), and 870 g of dough in a cylinder of 5 cm radius (De Vries et al., 1989). Another possible explanation is the baking time selected for observation as the water content at the core is expected to increase continuously with baking (provided that a thermal gradient persists) and higher water content should be obtained at the core at the end of baking. Values were obtained after 35 and 40 min by Thorvaldsson and Skjo ldebrand (1998) and De Vries et al. (1989), respectively, instead of 47 min in this study.
Table 1 Water content during baking at dierent baking times Baking time t (min) Series number Run number Region in the loaf slice

The present study measured the increase in water content at the core at the very beginning of baking. At this stage, the core was still liquid (dough)see Section 2. Fig. 3(a)(c) present the temperature proles in each region at 2, 4 and 7 min of baking, respectively. The temperatures in the top and bottom regions increased whereas the temperatures in the core region did not exceed 6065 C (this temperature range was calculated after linear interpolation between experimental data and can thus be slightly overestimated). Such temperatures are often reported at the onset of starch gelatinisation (Bloksma, 1990; Engelsen, Jensen, Pedersen, Norgaard, & Munck, 2001; Zanoni, Peri, & Bruno, 1995) which is consistent with a still liquid core (dough). This combination of local water content and temperature measurements conrmed that the evaporation condensation mechanism takes place even when the pores are not connected (closed porous structure) as previously proposed by De Vries et al. (1989). Moreover, it suggests that water moving towards the core is not (or not fully) mobilized by the area under gelatinization as previously stated by Thorvaldsson and Skjo ldebrand (1998). No real increase in water content was observed at 2 min although

Water content (% wet basis) At t0 (dough) x r 0.1 At t x 32.6 50.2 47.9 30.0 50.3 46.0 28.9 49.4 47.0 37.0 49.9 46.3 42.9 51.0 48.2 41.1 51.4 49.3 45.5 50.6 48.1 41.1 49.4 46.4 r 0.1 0.8 0.7 0.4 1.2 0.5 0.0 0.7 1.4 0.2 1.0 0.5 0.2 0.8 0.6 0.3 0.3 0.8 0.1 0.6 1.7 0.2 1.3

Top Core Bottom Top Core Bottom Top Core Bottom Top Core Bottom Top Core Bottom Top Core Bottom Top Core Bottom Top Core Bottom

49.6

0.6

49.6

0.0

0.7

49.3

0.1

0.1

49.0

0.1

0.9

50.2

0.0

0.8

50.1

0.0

1.3

50.2

0.1

0.4

49.0

0.1

0.4

Mean values refer to three samples along the z-axis; r is the corresponding standard deviation. D is the dierence in average water content between t and t0 (dough).

1172
60 50

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

crumb

40 30 20 10 0
crumb liquid core

a
60 50

25

35

45

55

65

75

85

95

30 20 10 0

b
60

25

35

45

55

65

75

85

95

Height (mm)

40 30 20 10 0
liquid core

crumb

50

crumb

crust

liquid core

40

times, since the latter were clearly not based on repetitions (De Vries et al., 1989; Thorvaldsson & Skjo ldebrand, 1998). Moreover, other studies (Zanoni & Peri, 1993) observed no increase in water content at the end of baking (30 min). It is probable that the evaporationcondensation mechanism still persisted in our operating conditions, but was not measurable because water losses by evaporation from cut, hot surfaces counterbalanced the possible increase by evaporationcondensation. Indeed, at t 6 7 min the core temperature did not exceed 65 C (Fig. 3(a)(c)) and water losses could have been small enough to allow the measurement of any increase in water content. The porous structure is expected to be closed before the dough/crumb transformation and this may also contribute to limited bias when sampling at early baking times. Apart from experimental bias due to destructive sampling, another explanation for not observing any increase in water content could be the severe reduction in the driving force for water transport, i.e. the temperature gradient. The temperature gradient at t = 15 min did not exceed 30 C in the lower part of the loaf (Fig. 3(d)), whereas it reached 4050 C for t 6 7 min. This is typical of small-sized products and again could explain the dierences from previously ndings. 4. Conclusion Our results showed an increase in water content at the core during the rst 7 min of baking and with the results from De Vries et al. (1989) and Thorvaldsson and Skjo ldebrand (1998), conrmed the occurrence of the evaporationcondensation mechanism during bread baking. This increase was observed in regions where the temperature did not exceed 65 C, tending to conrm the possible occurrence of the evaporationcondensation mechanism even in a closed porous structure (dough). When comparing our values to values reported in the literature, increases in water content appeared to vary in amplitude and with the baking time, such variations mainly being explained by the dierences in loaf thickness. Thickness acts on the driving force for water transport by evaporationcondensation i.e. the temperature gradient; it is expected to modify both its slope (or intensity) and its time duration. Given the number of transformations taking place during baking and their interactions, this hypothesis can only be validated by comparing a theoretical model of baking including water transport with the evaporation condensation mechanism. This will be our next stage of investigation. As data free of bias are required for full validation of such a model, and bias from sampling cannot be quantied, the development of non-destructive, quantitative methods is also required in this research area. Acknowledgements This work was supported by grants from the French Ministry of Research (RARE-Canal-salve, 01 P0838-01 P 0839-01 P 0840).

Height (mm)

Height (mm)

c
60 50

25

35

45

55

65

75

85

95

Height (mm)

30 20 10 0

25

35

45

55

65

75

85

95

Temperature (C)

Fig. 3. (a)(d): Temperature proles at 2 min (a), 4 min (b), 7 min (c) and 15 min (d) of baking inside the loaf. The temperature value at 20 mm is the mean of four repetitions and is accompanied by the corresponding standard deviation. The three regions distinguished at early baking times are specied on the right of the gure. Refer to the text for further details.

signicant thermal gradients had already developed (Fig. 3(a)). These could therefore be interpreted as inertia eects. The increase in water content was not signicant at t > 7 min. For instance, at 40 min, one baking run out of two presented, a 0.23% increase for y* = 0.48 compared to the initial water content. It can hardly be concluded that the last results are not consistent with previously reported ndings for long baking

final height

40

crumb

crumb

M.J. Wagner et al. / Journal of Food Engineering 78 (2007) 11671173

1173

References
Bloksma, A. H. (1990). Dough structure, dough rheology and baking quality. Cereal Foods World, 35(2), 237244. Czuchajowska, Z., Pomeranz, Y., & Jeers, H. C. (1989). Water activity and moisture content of dough and bread. Cereal Chemistry, 66(2), 128132. De Vries, U., Sluimer, P., & Bloksma, A. H. (1989). A quantitative model for heat transport in dough and crumb during baking. In Cereal Science and Technology in Sweden, Proceedings of an International Symposium (pp. 174188). Sweden: Lund University. Engelsen, S. B., Jensen, M. K., Pedersen, H. T., Norgaard, L., & Munck, L. (2001). NMR-baking and multivariate prediction of instrumental texture parameters in bread. Journal of Cereal Science, 33(1), 5969. Hasatani, M., Arai, N., Harui, H., & Itaya, Y. (1992). Eect of drying on heat transfer of bread during baking in oven. Drying Technology, 10(3), 623639.

Lostie, M., Peczalski, R., Andrieu, J., & Laurent, M. (2002). Study of sponge cake batter baking process. I: Experimental data. Journal of Food Engineering, 51(2), 131137. Marston, P. E., & Wannan, T. L. (1976). Bread baking: the transformation from dough to bread. Bakers Digest, 2428. Thorvaldsson, K., & Skjo ldebrand, C. (1996). Method and instrument for measuring local water content inside food. Journal of Food Engineering, 29, 111. Thorvaldsson, K., & Skjo ldebrand, C. (1998). Water diusion in bread during baking. Lebensmittel Wissenschaft und Technologie, 31, 658663. Wagner, M., Lucas, T., Davenel, A., Broyart, B., Collewet, G., & Trystram, G., (2003). Study of bread baking process: MRI experimental data. In Proceedings ICEF9, Montpellier, Fr, p. 496. Zanoni, B., & Peri, C. (1993). A study of the bread-baking process I: A phenomenological model. Journal of Food Engineering, 19(4), 389398. Zanoni, B., Peri, C., & Bruno, D. (1995). Modelling of starch gelatinisation kinetics of bread crumb during baking. Lebensmittel Wissenschaft and Technologie, 28(3), 314318.

You might also like