Vous êtes sur la page 1sur 108

Drying Technology

An International Journal

ISSN: 0737-3937 (Print) 1532-2300 (Online) Journal homepage: http://www.tandfonline.com/loi/ldrt20

Drying and Rehydration of Pasta

Takenobu Ogawa & Shuji Adachi

To cite this article: Takenobu Ogawa & Shuji Adachi (2017): Drying and Rehydration of Pasta,
Drying Technology, DOI: 10.1080/07373937.2017.1307220

To link to this article: http://dx.doi.org/10.1080/07373937.2017.1307220

Accepted author version posted online: 16


May 2017.

Submit your article to this journal

Article views: 3

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=ldrt20

Download by: [The UC San Diego Library] Date: 20 May 2017, At: 14:45
Drying and Rehydration of Pasta

Takenobu Ogawa1, Shuji Adachi2,


1
Division of Food Science and Biotechnology, Graduate School of Agriculture, Kyoto
University, Uji, Japan
2
Division of Food Science and Biotechnology, Graduate School of Agriculture, Kyoto
University, Sakyo-ku, Japan

Address Correspondence Shuji Adachi,Kitashirakawa, Oiwake-machi, Sakyo-ku, Kyoto


606-8502, Japan. E-mail: adachi@kais.kyoto-u.ac.jp

Abstract

Pasta is dried at the production stage and consumed after rehydration by cooking.

Because the water migration behavior during drying and rehydration largely affects the

quality of pasta, a better understanding of this behavior helps to efficiently manufacture

and cook pasta of good quality. However, the key mechanism controlling water migration

inside pasta is not fully understood. This review aims to provide an overview of the

phenomena occurring during the drying and rehydration processes. In addition, the effects

of drying and rehydration conditions on pasta quality are discussed. Knowledge of these

effects would be useful for rational determination of the conditions for producing pasta

with desired qualities.

KEYWORDS: pasta; drying; rehydration; quality

INTRODUCTION

The phenomena occurring during an industrial food-making process are not fully

understood, and the process is often designed and operated based on a great deal of

1
experience. Drying is one of the most common processes for improving the shelf life of

food. The primary objective of food drying is to ensure longer quality preservation by

decreasing the moisture content of the food to a level that minimizes microbial spoilage.

Prior to use or consumption of dried foods, they are usually sorbed or rehydrated to

improve the taste and digestibility. That is, the water molecules in food are removed and

added during the drying and rehydration processes, respectively. Because the water

migration behavior during these processes largely affects the quality of dried and

rehydrated foods, a better understanding of the water migration kinetics helps to

efficiently manufacture dry food of good quality and cook it to obtain a good texture,

taste, and digestibility. However, the key mechanism controlling the water migration

inside food is not fully understood.

Pasta consists of mainly starch and protein, which are major components of food. Pasta is

also a porous material. Therefore, the knowledge obtained from pasta can provide useful

information for the design of other food-making processes. Advantageously, the

properties of pasta are easy to measure and analyze owing to its macroscopically

homogeneous structure. The word pasta, which is Italian for dough, is generally used

for the products fitting the Italian style of extruded foods, such as spaghetti or lasagna.

Pasta is a healthy food because it is relatively low in fat and high in carbohydrates, with a

good composition of protein. The main ingredients of pasta are principally durum wheat

semolina and water. Milling of durum wheat (Triticum durum), which is the hardest

wheat, produces a coarse particle called semolina. Durum semolina is ideal for making

pasta because of its hardness, intense yellow color, and nutty taste.[1] A proportion of

2
water (18%25%) is added to dry raw durum semolina at 3540C and the mixture is

kneaded for 1020 min to produce fresh dough with an average moisture content of

30%32%.[1] The durum semolina dough is then extruded through a die using a vacuum

extruder to produce pasta.[1,2] Although dies made of bronze have traditionally been used,

dies made of Teflon have recently been introduced to elongate the lifetime of the die by

reducing wear, obtain a smoother pasta surface, and improve the appearance of dried

pasta.[35] Pasta prepared using Teflon and bronze dies have smooth and rough surfaces,

respectively. Pasta prepared using bronze dies has higher porosity, lower density, lower

breaking strength, and a larger effective diffusion coefficient of water during drying than

that prepared using Teflon dies.[6,7]

Pasta is usually distributed in the dry state in order to improve its storage stability and

transportation efficiency. Although some dried foods, such as instant noodles, are

processed using superheated steam, with the increasing demand for ready-to-eat

products,[810] pasta is dried using hot air. Dried pasta has a moisture content of ca. 11%

on a wet basis, which is suitable for preservation. Knowledge of the moisture sorption

isotherm, which describes the relationship between the moisture content and equilibrium

relative humidity, is useful for understanding the phenomena occurring during the drying

or rehydration process of food.[11] This knowledge allows optimization of drying times

and energy utilization, as well as evaluation of the storage stability of food products. The

moisture content is also related to some of deteriorative mechanisms, such as microbial

growth, enzymatic reactions, non-enzymatic browning, and lipid oxidation.[12,13]

Therefore, as reviewed by Al-Muhtaseb et al.,[14] the moisture sorption isotherms of many

3
food products, for example, starchy foods (e.g., corn, potato, wheat flour, and rice), high

protein foods (e.g., chicken, egg, milk, and cheese), fruits (e.g., banana, apple, apricot,

and raisin), and vegetables (e.g., green pepper, lentil, tomato, onion, sugar beet root,

carrot, and celery), have been experimentally determined. Many models have been

proposed for the dependence of the equilibrium moisture content on the relative humidity,

and van den Berg and Bruin analyzed 77 models.[15] These models can be further

categorized into several groups: kinetic models based on the monolayer sorption theory

(e.g., Langmuir model), kinetic models based on the multilayer sorption theory (e.g.,

BET and GAB models), and empirical and semi-empirical models (e.g., Peleg and Oswin

models).

A typical drying curve for pasta is concave, i.e., the moisture content rapidly decreases

during the early stage of drying, and gradually decelerates to become very low at the later

stage.[16] The drying characteristic curve is usually divided into preheating, constant

drying-rate, and decreasing drying-rate periods. Because the preheating and constant

drying-rate periods are usually very short compared with the whole drying period during

the production of dried pasta, the preheating and constant drying-rate periods have been

ignored and the decreasing drying-rate period is assumed from the beginning of drying.

Many theoretical and empirical models, most of which are based on Ficks law of

diffusion,[17-20] have been reported for describing water transfer and its kinetics during the

decreasing drying-rate period. Fourteen types of empirical or semi-empirical equations

have been utilized to describe drying curves,[21] including the Newton,[22] Page,[2326]

modified Page (two types),[27,28] Henderson and Pabis,[27], logarithmic,[29] two term,[30]

4
two-term exponential,[31] Wang and Singh,[32] Thompson et al.,[33] diffusion

approximation,[34] Verma et al.,[35] modified Henderson and Pabis,[36] and Midilli and

Kucuk.[37] In spite of the assumption of a decreasing drying-rate period from the

beginning of drying, good agreement between the predicted and experimental results has

generally been obtained with these models. For drying udon (Japanese noodle), it was

reported that the initial drying rate is crucial to prevent crack formation, which brings

about a remarkable decrease of the udon quality.[38] This result indicates the importance

of precisely predicting the drying behavior during the early stage, in which a large

amount of water evaporates from the samples surface.

Multiple factors contribute to pasta quality, and their roles are not yet fully understood.[39]

Protein content and composition play important roles in pasta quality. The protein content

is generally recognized as a primary factor, and gluten strength as a secondary

factor.[4043] The drying temperature also affects pasta quality. Although pasta is

traditionally dried at low temperature, the drying temperature has been increasing with

the advancement of technology. High-temperature (HT) and very-high-temperature

(VHT) drying methods have been rapidly accepted, and ultra-high-temperature (UHT)

drying has also emerged. Owing to such increases in the drying temperature, many

researchers have investigated the effects of drying temperature on pasta properties.[4458]

Consideration on energy balance or efficiency during drying would be important although

it is not referred to in this review.

5
As mentioned above, pasta mainly consists of starch and gluten. HT drying causes

protein denaturation, which can be assessed by gluten solubility in acetic acid solution,[45]

and promotes inter- and/or intramolecular disulfide cross-linking of gluten, resulting in

increased rigidity of the protein network.[46,47] The strengthening of the protein network

increases pasta firmness.[44] Starch damage generally brings about lower cooking quality.

A certain amount of starch is damaged during the production of raw pasta by extrusion,

but most damage is increased in the drying stage, where starch is subjected to the action

of amylolytic enzymes.[48] The level of starch damage is reported to be higher for HT

pasta because of the higher activity of such enzymes in HT pasta drying than in the VHT

process.[48]

The intrinsic properties of the raw materials largely affect the cooking quality at the low

temperatures used for the traditional production process. However, a high drying

temperature can modify the material properties and improve the cooking quality,

regardless of the raw material properties.[49,50] It is reported to increase the amylose

content in cooked pasta, resulting in a decrease of the amount of amylose eluted into the

cooking water.[51] This decreased elution results in high pasta quality.[39] The extent of

amylose elution is closely related to pasta stickiness[52] and elasticity, which is mainly

related to the drying temperature.[42]

Color is always a crucial factor governing the choice of pasta by consumers. The color is

positively affected by the drying temperature via the inhibition of enzymatic activity. A

lower drying temperature is expected to considerably change the color of pasta.[53]

6
Increased pasta yellowness is caused by carotenoid bleaching by partial inactivation of

endogenous lipoxygenase, and the increase in red-brown color comes from the formation

of melanoidin in the advanced Maillard reaction at a low product moisture content.[54,55]

A high drying temperature positively affects cooking quality and product color, reduces

microorganism content, and shortens drying time.[56-58] The drying temperature is also

reported to have a positive effect on certain textures of cooked rice.[59] Thus, high

temperature might be preferred in industrial pasta manufacturing, although pasta dried at

low temperature is still distributed in many countries. These facts suggest that the optimal

drying conditions depend on the properties needed to satisfy consumer demands.

Therefore, the effects of drying conditions on pasta properties should be systematically

assessed.

Cracks generated inside pasta mar its aesthetic appearance and result in a reduction of

product value. In order to prevent crack generation and produce a high-quality product,

the temperature and humidity are changed stepwise with time during the production of

dry pasta (called programmed drying). The gradual stepwise transitions in temperature

and humidity during programmed drying are also important to control product quality.[60]

These drying processes involve three variables, temperature, humidity, and duration, and

their combination. Although the temperature effects on various pasta properties have been

intensely investigated, the effects of humidity have scarcely been examined. The lack of

studies on humidity-dependent properties could come from the fact that the drying rate is

mainly governed by temperature. We reported the effects of temperature and relative

7
humidity on the drying rate during a constant drying-rate period, together with the overall

mass transfer coefficient during a decreasing drying-rate period, for a cylindrical tablet of

durum semolina dough using thermogravimetry, and demonstrated that the parameters

determining the drying characteristics are affected mainly by temperature rather than

humidity.[16] Thus, the suitable range of humidity for programmed-drying conditions is

partially limited by the drying temperature, as crack generation should be prevented in

practice. Confidentiality issues are another possible reason, particularly for industrial

collaboration studies, as the prevention of cracking is crucial for the industrial production

of dry pasta. Therefore, the humidity conditions have not been explicitly stated. On the

other hand, as humidity affects the properties of starch, changes in the starch properties

can alter the texture and sauce retention capacity, which are largely related to amylose

leaching from starch granules, as well as the rehydration characteristics during cooking,

which are governed by the gelatinization properties of starch.

Because dried pasta is consumed after rehydration by cooking to recover its properties, it

is important to understand the process occurring during the rehydration of dried pasta.

This process is a complicated mass transport process governed by several migration

mechanisms of water into pores.[61] Empirical and theoretical equations can be used to

describe the transient behavior of rehydration process.[62] Because less effort is required

for the development of empirical equations, such equations can be useful for predicting

and optimizing the rehydration kinetics.[63] Six types of empirical or semi-empirical

equations are often utilized to describe rehydration kinetics:[64] the exponential

equation,[65] Peleg's model,[66] first-order kinetics,[67] Becker's model,[68] Weibull

8
distribution function,[69] and normalized Weibull distribution function.[70] Peleg's model

and the Weibull distribution function can predict the moisture content of pasta accurately

for any rehydration time.[71] However, empirical equations treat the rehydration process

as a black box, by varying specific input setup parameters, measuring output quantities,

and deriving adequate correlations. Therefore, the coefficients of the equation should be

determined by varying the specific input setup parameters in detail, such as the

physicochemical characteristics of pasta manufactured under various conditions,

rehydration temperature, and concentration of NaCl in an immersion solution. On the

other hand, theoretical equations are not convenient for practical purposes.[64,71] However,

they provide insight into the mechanistic relevance of an observed phenomenon.[61] Most

theoretical equations describing rehydration kinetics are based on Fick's first and second

laws of diffusion, where the difference in the moisture content of pasta is a driving force

for water migration [61,7275], although other phenomena such as capillary flow and

swelling are involved in the rehydration process. The rehydration process of fresh penne

pasta at 2080C was characterized by two effective diffusion coefficients using the

second law of diffusion.[71] In addition to water diffusion, starch crystalline domain

melting, macromolecular matrix relaxation, and residual deformation release occur

during pasta rehydration.[76] However, in some opinions, the actual process of moisture

migration is not diffusion-controlled and several other mechanisms have been proposed,

such as water imbibition, capillarity, and flow in porous media.[61,77]

Although the moisture distribution is obtained by numerically solving the equations, a

proposed model is typically validated by measuring the average moisture content of an

9
entire sample[78] because of the absence of an adequate method for obtaining the precise

moisture distribution of the sample and for verifying the numerical results. Owing to the

absence of an adequate method, it is difficult to discern or interpret the mechanism

controlling water migration during rehydration. Rehydration curves express transient

changes in the average moisture content and are obtained by numerically solving the

various models. The curves are of the hyperbolic type and satisfactorily fit the

experimental data.[62,71,75,76,78] However, agreement between the empirical and theoretical

curves is insufficient for verifying the appropriateness of the models. The measurement

of a precise moisture distribution is necessary for verification of the model.

Nuclear magnetic resonance (NMR) and magnetic resonance imaging (MRI) are

powerful tools for measuring the moisture distribution in foods, and have been applied

for measuring the moisture distribution in rice grains during cooking,[79] noodles during

drying or rehydration,[80-83] and cheese during brining.[84] However, the accuracy of these

techniques is insufficient to verify the numerically calculated distribution because of the

following limitations. First, the minimum moisture content measureable by these

techniques is relatively high. Although dried pasta begins rehydrating from a moisture

content of approximately 0.11 kg-H2O/kg-dry matter (d.m.), fast water proton relaxation

does not allow moisture contents of less than 0.67 kg-H2O/kg-d.m. to be measured.[82]

Second, these techniques have low spatial resolution. The resolution of MRI

measurements is 65 m at best.[79] Therefore, only approximately 12 data points can be

obtained for pasta with a radius of 0.8 mm. Third, these measurements are time

consuming. The MRI technique requires a few minutes to measure the moisture

10
distribution of a sample, but the moisture distribution of samples, such as pasta, change

within this time. The final limitation is the very high price of NMR and MRI. Therefore,

it is necessary to develop a method without these limitations for determining the

mechanism controlling water migration in dried foods.

Dried pasta is yellowish or yellowish-brown and becomes lighter as the moisture content

increases. Based on the color change of pasta during rehydration, a new method using a

digital camera was developed to precisely measure the moisture content (0.1

kg-H2O/kg-d.m. or higher) of pasta.[85] The cross-sectional moisture distribution of a

sample was measured by the method. Digital cameras with high pixel densities are

currently available and provide high-resolution images that enable measurement of the

precise moisture distribution during the rehydration process.

In this paper, transient behavior of the drying and rehydration processes of pasta and the

factors affecting the kinetics are discussed.

DRYING OF PASTA

Moisture Sorption Isotherm Of Durum Semolina

A moisture sorption isotherm, which represents the relationship between the water

activity and the moisture content of a sample at a specific temperature, reflects the

interaction between water molecules and the sample.[8688] The temperature dependence

of the moisture sorption behavior provides information on the thermodynamic properties.

The isosteric heat can be determined from the moisture sorption isotherms based on the

11
Clausius-Clapeyron equation. Knowledge of the differential heat of sorption is useful for

designing equipment to be utilized in drying processes.[89,90] Drying is a combined heat

and mass transfer process. The drying air temperature is 3040C in a traditional process

for drying pasta, and the maximum drying temperature in an industrial process is

8090C to shorten the drying time. Therefore, the moisture sorption isotherm of durum

semolina over a wide range of temperatures is necessary in order to reasonably design the

industrial pasta drying process.

Figure 1 shows the moisture desorption isotherms for durum semolina at various

temperatures.[91] The isotherms were sigmoidal at every temperature and were

categorized as type II according to Brunauer et al.[86] These results were similar to those

reported previously.[11,87] The smaller amount of sorbed water at higher temperatures

indicated that the sorption of water onto the semolina flour was exothermic. A slight

hysteresis was observed between sorption and desorption at low temperatures.

Both the sorption and desorption isotherms could be separately expressed by the

Guggenheim-Anderson-de Boer (GAB) equation:

abcaw
X (1)
(1 caw )(1 caw bcaw )

where X is the amount of water sorbed, aw is the water activity, and a, b, and c are

constants. The constant a corresponds to the amount of water for monolayer coverage, b

is a measure of the interaction between the adsorbate (water) and solid material (flour),

and c is a correction coefficient. The a and b values became smaller at higher

12
temperatures, while c scarcely depended on the temperature. The equilibrium moisture

content of durum semolina can be evaluated under any temperature and relative humidity

conditions from the temperature dependencies of these parameters.

The isosteric heat, q, is an indication of the interaction force between a water molecule

and a sorption site on durum semolina. The q value at a specific amount of water sorbed,

X, can be estimated based on the Clausius-Clapeyron equation:[92]

d ln aw
q R (2)
d(1/T ) X

where R is the gas constant and T is the absolute temperature. The plots for estimating the

q values at various X values from both the sorption and desorption isotherms were linear,

indicating that Eq. (2) is applicable for estimating the q value. The larger q values

obtained at lower moisture contents indicated that water molecules interact more strongly

with durum semolina at lower moisture contents. The plots for the desorption process lie

over those for the sorption process, which indicates that the energy consumed during

desorption of a water molecule sorbed onto durum semolina is greater than the energy

liberated during water sorption.

Equation (3) has also been used for cereals to express the relationship among the amount

of sorbed water X, temperature T, and water activity aw.[93,94]

ln aw
K1 K 2 X (3)
1/ T 1/ T

13
where T, K1, and K2 are parameters. The equation was applied to the amounts of sorbed

water for both the sorption and desorption processes. The T, K1, and K2 values for the

sorption process were 448 K, 6.37 103 K, and 0.814, respectively, whereas the T, K1,

and K2 values for the desorption processes were 400 K, 9.55 103 K, and 0.821,

respectively. The plots of the X values calculated using the estimated T, K1, and K2

values against the observed X values for both the sorption and desorption processes lay

on the line with a slope of unity, indicating that the equation is applicable to moisture

sorption onto durum semolina. The isosteric heat for the sorption and desorption

processes calculated from Eq. (3) coincided with those calculated from Eq. (2). Therefore,

in addition to Eq. (2), Eq. (3) is also useful for calculating the moisture-content

dependences of the isosteric heats.

The moisture sorption isotherms on starch and gluten, which were isolated from durum

wheat flour, at 30C were categorized as sigmoidal type II according to Brunauer et al.[86]

and could be expressed by the GAB equation. The a, b, and c values were 0.0876

kg-H2O/kg-d.m., 45.6, and 0.715 for starch and 0.0763 kg-H2O/kg-d.m., 37.0, and 0.728

for gluten.

The equilibrium moisture content could be expressed by summing the moisture content of

the constituent components for weak flour.[95] The carbohydrate and protein contents of

durum semolina are about 81.7% and 15.0% (dry basis), respectively. The moisture

sorption isotherm calculated from the isotherms on starch and gluten and their contents

was almost the same as the observed moisture sorption isotherm on durum semolina.[95]

14
Although the moisture sorption isotherm on pasta at 60C lay slightly over that on durum

semolina, the difference was not significant, except at very high water activities. This fact

indicated that the pasta making process had no significant effect on water sorption.

Dilatometric Measurement Of Partial Molar Volume Of Water Sorbed Onto Durum

Semolina

The interaction of water molecules with the durum wheat flour plays an important role in

the drying and rehydration processes. The partial molar volume of water would provide

useful information on this interaction,[96] and the volume was measured by dilatometry.[97]

Pasta made from pregelatinized durum wheat flour has been prepared to offer consumers

the convenience of a short cooking time.[98] Gelatinized flour is more water-accessible,[99]

while dry-heating increases the hydrophobicity of the flour.[100] These facts indicate that

moist- or dry-heating of the flour changes its properties. The partial molar volumes of

water molecules sorbed onto untreated, dry-heated, and pregelatinized durum wheat flour

samples with various moisture contents were measured at 25C using dilatometry. The

moisture sorption isotherms of the flour samples were also measured to better understand

the interaction of water with durum wheat flour. The differential scanning calorimetry

(DSC) curve for the untreated flour exhibited an endothermic peak near 60C, which is

ascribable to starch gelatinization. The curve for the pregelatinized flour had no peak near

60C, indicating that the flour had been gelatinized. The mean pore sizes of the

dry-heated and pregelatinized samples were slightly larger than that of the untreated

15
sample, while there was no significant difference in the specific surface areas of the

untreated, dry-heated, and pregelatinized flour samples.

Figure 2 shows the partial molar volumes, V , of water sorbed onto untreated, dry-heated,

and pregelatinized flour samples with various moisture contents, or mass ratios of water

to flour. The V value of the untreated flour sample was 9 cm3/mol at a moisture content

of 0.03 kg-H2O/kg-d.m., and increased with increasing moisture content. This value

reached a constant value of 1718 cm3/mol at moisture contents of ca. 0.2

kg-H2O/kg-d.m. or higher. It took a longer time to reach equilibrium at lower moisture

contents, e.g., 15, 7, and 2 d at respective moisture contents of 0.05, 0.15, and 0.30

kg-H2O/kg-d.m. Because the V values of the dry-heated and pregelatinized flour

samples showed a similar dependence on the moisture content to that of untreated flour,

dry-heating and pregelatinization had no significant influence on the interaction with

water. These facts suggest that water molecules interacted more strongly with the flour at

lower moisture contents. This interaction would cause slower drying rates at the lower

moisture contents.[16]

In order to estimate the volumetric behavior of the water molecules sorbed onto the flour,

the moisture sorption isotherm and the partial molar volume of the untreated flour are

illustrated together in the inset of Fig. 2. The water molecules would be sorbed as a

monolayer at a moisture content of less than ca. 0.1 kg-H2O/kg-d.m., and such water

molecules have a very low V value owing to the strong interaction with or incorporation

into the flour. With more layers, the water molecules exhibit higher V values, which

16
became constant in the multilayer region at moisture contents of 0.2 kg-H2O/kg-d.m. or

higher. The moisture content was the same as that when the glass transition of durum

semolina occurred at 25C.[101] The sorbed water molecules in the multilayer region

behaved like bulk water molecules owing to their very weak interaction with the flour.

Thermogravimetric Study On Drying

The drying rate during the constant drying-rate period and mass transfer coefficient,

which are usually determined using laboratory-scale experimental apparatus, are

necessary to predict the change in the moisture content during drying. In this section, the

drying rate during the constant drying-rate period and the mass transfer coefficient for

drying pasta under various conditions were estimated by thermogravimetry (TG). This

technique is commonly used for the analyses of thermal reaction processes, including

heat decomposition, gassolid reactions, and quantitative determination of crystallization

water, because it allows accurate measurements of changes in weight using very small

sample amounts (tens of milligrams) of durum semolina dough.

Constant-Drying Conditions

The hydrated durum wheat semolina was packed into a single-sided open cell using a

glass syringe equipped with a vacuum pump (Fig. 3 inset), and the weight loss during

drying was measured using a thermometer. Figure 3 shows an example of the drying

characteristic curves obtained by thermogravimetry.[16] No preheating period was

observed, but the constant drying-rate period could be distinguished from the decreasing

drying-rate period. That is, the drying rate was constant at high moisture contents during

17
the early stage of drying, and the drying rate then decreased with subsequent lowering of

the moisture content. Similar behavior was observed under all conditions from 30 to

90C and from 0% to 80% RH. About 20% of the water evaporated during the constant

drying-rate period. Under a very fast drying rate, pasta is predisposed to form cracks,

which lowers its quality. Using the finite element method to calculate the moisture

distribution within udon, Inazu et al.[38] indicated that the early stage of drying is a crucial

step for crack formation based on the mechanical balance. Therefore, the constant

drying-rate period should be considered to precisely predict the drying curve for the

prevention of crack formation in pasta.

For the drying curves of hydrated semolina with three different thicknesses (0.7, 1.0, and

1.4 mm) at the same temperature and relative humidity, all the plots of moisture content

against drying time divided by the square of the thickness lay on a curve during the

decreasing drying-rate period. This fact indicated that water diffusion mainly governs the

water migration during drying of pasta. Thus, the quotient of time and the square of the

thickness, t/L2, was replaced by time, t, during the decreasing drying-rate period. To

simplify the model, the following assumptions were made: (1) the product temperature is

constant owing to rapid heat transfer in the pasta; (2) the moisture diffusivity within the

pasta is independent of the moisture content; (3) the volumetric concentration of the pasta

is constant at any moisture content; and (4) there is no shrinkage during drying. The

drying rates during the constant and decreasing drying-rate periods are given by Eqs. (4)

and (5), respectively.

w dX
Rw (4)
S dt

18
dX
k(X Xe) (5)
d(t /L2 )

where Rw is the drying rate, w is the dry weight of the sample, S is the drying area, X is

the moisture content at time t, L is the thickness, k is the mass transfer coefficient, and Xe

is the equilibrium moisture content. The analytical solutions for the one-dimensional

rectangular and cylindrical geometries are given by Eqs. (6) and (7), respectively, under

the assumptions that the initial moisture distribution is uniform at the moisture content X0

and the surface are kept at the same moisture content Xe.[102]

X Xe 8 1 (2n 1) 2 2 Det
2 2
exp (6)
X0 Xe n 1 (2n 1) 4 L2

2
X Xe 1 n Det
4 2
exp (7)
X0 Xe n 1 n r2

where De is the effective diffusion coefficient of water in the sample, n is the nth positive

root of J0(n) = 0, J0(x) is the Bessel function of the first kind of order zero, and r is the

radius. Because water diffusion controls the drying rate during the decreasing drying-rate

period, the mass transfer coefficient for a slab, ks, is related to that for a cylinder, kc, by

Eq. (8) based on Eqs. (5), (6), and (7):

4ks kc
2 2
De (8)
1

The drying rate during the constant drying-rate period, Rc, and the ks value for the

decreasing drying-rate period were estimated using Eqs. (4) and (5) from the drying

19
curves obtained by thermogravimetry operated under various conditions, and the

estimated Rc was expressed as a binominal function of the temperature, T, and the relative

humidity, H, by Eq. (9).

1
Rc (6.57 10 2.69 10 1 T 9.48 10 2 H 1.10 10 3 T 2
(9)
3.26 10 4 H 2 7.76 10 4 TH ) 10 5

The ks value, which is derived from De, can also be expressed as a function of T and H

because De depends on both T and H.[103,104]

1
ks ( 4.27 10 6.45 10 2 T 1.32 10 2 H 2.73 10 4 T 2
(10)
4.05 10 5 H 2 1.02 10 4 TH ) 10 10

Good correlations for both the Rc and ks values were obtained between the observed and

calculated values with coefficient of determination, R2, of 0.976 and 0.985, respectively.

The Rc value increased with a decrease in the relative humidity at low temperatures

owing to the greater difference in the absolute humidity between the bulk air phase and

the layer adjacent to the sample surface at the lower humidity. On the other hand, the Rc

value scarcely depended on the relative humidity at high temperatures, suggesting that

film mass transfer of water on the surface might be the rate-controlling step at high

temperatures. Because the diffusion of water within the sample is the rate-controlling step

during the decreasing drying-rate period, the ks value weakly depended on the relative

humidity at every temperature.

The water sorption isotherms of durum semolina and pasta over wide ranges of

temperatures and relative humidities were shown in the previous section, and the
20
isotherms of durum semolina and pasta under specific conditions are expected to overlap

at relative humidities higher than 80%.[96] The GAB equation for describing the water

sorption isotherm at a specific temperature is expressed as a function of H by Eq. (11),

where the equilibrium moisture content of pasta, Xe, and the relative humidity, H, are

used instead of X and aw in Eq. (1). The coefficients of the GAB equation, a, b, and c, are

expressed as a function of T by Eqs. (12), (13), and (14) in order to estimate the Xe value

at any T and H.

abcH
Xe (11)
(1 cH )(1 cH bcH )

a 1.08 10 6 T 3 1.99 10 4 T 2 1.26 10 2 T 3.46 10 1


(12)

b 1.71 10 4 T 3 2.64 10 2 T 2 1.04T 7.06 (13)

c 7.93 10 8 T 3 1.51 10 5 T 2 9.26 10 4 T 1.18 10 2


(14)

Equations (11), (12), (13), and (14) were applicable for estimating the Xe value under any

condition in the temperature and relative humidity ranges of 3090C and 10%90% RH,

respectively.

Drying Under Programmed Temperature-Raising Conditions

A typical drying curve for pasta is concave, i.e., the moisture content rapidly decreases

during the early stages of drying, and gradually decelerates to become very low at later

stages.[16] Because a large part of the entire drying period is occupied by drying in the

low-moisture regime, any increase in the drying rate in this region will reduce the drying

time. Pasta transforms from a rubbery state to a glassy state with a concomitant decrease

in moisture content during drying,[105] similar to that observed for the drying of

21
strawberries,[106] tomatoes,[107] apricots,[108] wheat,[101] and starch.[109,110] The drying

process can usually be described by Fick's law of diffusion, [17,18,111-114] in some cases,

taking into consideration the dependence of the diffusion coefficient on the moisture

content. However, near the glass transition point of durum wheat flour, this law cannot

predict drying behavior of pasta exactly because of the occurrence of non-Fickian

phenomena.[105,115] As a consequence, precise prediction of the drying behavior is difficult

in the low moisture-content region where this glass transition occurs. Rational design of

the pasta drying process requires knowledge of the drying rate over a wide range of

temperatures and moisture contents. To evaluate constant drying rates and mass-transfer

coefficients in the regime where the rates decrease, the drying processes based on a

decrease in the weight of the dough were analyzed in the previous section,[16] as

measured using TG at constant temperatures and humidities.[16] In this section, the drying

rate of durum wheat dough was measured using thermogravimetry at various

temperature-raising rates to estimate the dependence of the activation energy on the

moisture content.[116] DSC measurements were also performed under the same conditions

as the thermogravimetric ones.[116] These measurements enabled a discussion of the

effects of the temperature and the moisture content on the drying rate.

When the change in the sample weight is measured under a flow of dry nitrogen, the

equilibrium amount of water sorbed onto the flour should be zero and the fraction of

water remaining on the flour, Y, is calculated by Eq. (15):

X
Y (15)
X0

22
where X0 and X are the initial moisture content and the moisture content at time t,

respectively. The activation energy, E, at a specific Y value was estimated by an

isoconversion method using the drying data obtained at various temperature-raising rates.

Various methods are available,[117] and the differential-type method proposed by

Friedman[118] was adopted. The change in the Y value with time, dY/dt, corresponding to

the drying rate, can then be expressed by Eq. (16):

dY
k f (Y ) (16)
dt

where f (Y) is a kinetic function concerning the driving force for drying and k is the mass

transfer coefficient. The temperature dependence of the coefficient k obeyed the

Arrhenius equation:

E
k A0 exp (17)
RT

where A0 is the frequency or pre-exponential factor and E is the activation energy.

Substituting Eq. (17) into Eq. (16) gives the following equation:

dY E
ln ln[ A0 f (Y )] (18)
dt RT

The dY/dt values at a specific value of Y were evaluated by numerical differentiation of

the drying curve obtained at different temperature-raising rates.

23
Drying curves under the temperature-raising conditions were concave and similar to

those under the isothermal drying conditions in the previous section. The drying rate was

higher at increased temperature-raising rates, and became low at low moisture contents.

The applicability of Eq. (18) was examined by plotting the value of dY/dt versus 1/T on a

semi-logarithmic scale for a number of Y values. For each Y value, the points lay on a

straight line, which was used to obtain E and A0 f (Y).

The E values at Y values above 0.35, which corresponded to a moisture content of 0.14

kg-H2O/kg-d.m., were fairly constant at 32 kJ/mol, while the E values were larger at Y

values below 0.35, with E = 53.0 kJ/mol at Y = 0.2. These facts indicate that the drying

rates markedly decrease in the later stages of drying.

The DSC curves at various temperature-raising rates exhibited endothermic peaks. The

observation of such peaks at low moisture contents should result from the enthalpy

recovery of amorphous starch rather than from gelatinization. This assumption seemed to

be supported by the fact that the endothermic peak was broader at lower

temperature-raising rates and that the structure of amorphous starch was more relaxed

because of the longer aging time at lower rates. Using a hermetic cell, the glass transition

can typically be observed at temperatures above the endothermic peak during the early

stage of enthalpy relaxation in DSC.[119] The glass transition cannot clearly be observed

in measurements using an open cell, and the moisture content gradually decreases.

However, a slightly endothermic shoulder after the endothermic peak seemed to suggest

the occurrence of a glass transition.

24
Figure 4 shows the DSC curves as a function of the moisture content, based on the TG

and DSC measurements. Endothermic peaks were observed between moisture contents of

0.10 and 0.20 kg-H2O/kg-d.m., with the peak shifting to lower moisture contents for the

drying processes at higher temperature-raising rates. The constant drying rate was higher

at higher temperature-raising rates because of the higher temperature at a given moisture

content. The drying rate markedly decreased at moisture contents of 0.150.20

kg-H2O/kg-d.m. or lower, which roughly corresponds to the moisture content at which

the activation energy started to increase.

From the relationship between the conclusion temperature of the endothermic peak in the

DSC measurements and the moisture content, as well as those between the temperature

where the drying rate started to decrease rapidly or the glass transition temperature of

durum semolina flour and moisture content (Fig. 5), it was strongly suggested that the

glass transition occurred after the endothermic peak.[101] The plots for the temperature of

the inflection points of the drying characteristic curves were located near the glass

transition curve. Therefore, the glass transition of dough from the rubbery to the glassy

state produced a rapid increase in the activation energy and a rapid decrease in the drying

rate.

Hot Air Drying

The temperature and humidity are changed step-by-step with time in the practical process

of manufacturing dry pasta to produce a high-quality product. The reliability of the

above-mentioned model (previous section) and the estimated parameters was

25
demonstrated for tubular and cylindrical pasta (fettuccine and spaghetti, respectively) by

comparing the observed drying curves under programmed-drying conditions in the oven

to those calculated using the model and the parameters.[16] The drying conditions are

shown in Table 1. The fettuccine and spaghetti were dried at maximum temperatures of

80 and 60C, respectively.

Figure 6 illustrate the drying curves for spaghetti.[16] The solid curve was calculated

based on the proposed model (Eqs. (4) and (5)) using the estimated parameters, Rc, ks,

and we. The broken curve was calculated by assuming that the decreasing drying-rate

period starts at the beginning of the drying process without consideration of the constant

drying-rate period. The drying curves during the early stage of drying are shown in the

inset of Fig. 6. The solid curve well represented the experimental results, especially, the

drying behavior during the early stage for the fettuccine (data is not shown) and spaghetti.

Therefore, it was verified that the model and the parameters, which were estimated by

thermogravimetry on a small scale, were useful for predicting the drying curves of pasta

with various geometries under any condition.

Shrinkage And Tensile Stress During Drying

Knowledge of the mechanical properties of a food material, such as the stress-strain

curve[120] and critical stress,[57] is important for optimization of the drying process

because these properties depend on the product's moisture content and affect the

consumers preference. Many researchers have reported the apparent strength, apparent

stress-relaxation coefficient, Young's modulus, strain, and yield stress of pasta.[38,121123]

26
Pasta shrinks in association with the evaporation of water. Locally heterogeneous stress

and shrinkage within pasta during drying cause cracks to be generated, which largely

reduce the pasta quality.[38,124] The precise prediction of the moisture distribution within

the pasta, which causes shrinkage and stress, is necessary in order to find the conditions

for which no cracking occurs during drying. In this section, the shrinkage and tensile

stress of cylindrical and sheet-like pastas with various moisture contents of different

distributions are discussed.

The effect of the moisture distribution in pasta on shrinkage was examined by drying

fresh pasta under three different conditions.[125] In the first and second cases, the pasta

was dried at 50C and 40% RH for 8 h, and at 50C and 80% RH for 6 h, respectively, in

a temperature-humidity controlled chamber. In the third case, pasta prepared under the

conditions of the second case was put into a plastic bag and stored at 50C for 2 d to

obtain a homogeneous moisture distribution. The pasta samples dried in the first, second,

and third cases were designated as 40%, 80%, and equilibrated samples, respectively. The

moisture distribution of the 40% sample was more uneven than that of the 80% sample,

while the equilibrated sample had an even distribution.

The slope of the shrinkage strain against moisture content plot gives the shrinkage

coefficient. The coefficients for the width, height, and depth of the 40% and 80% samples

were 0.13, 0.12, and 1.03 and 0.20, 0.21, and 1.14, respectively.[125] For the equilibrated

sample, the coefficients for the width and height were 0.29 and 0.26, respectively, but the

coefficient for the depth could not be accurately evaluated owing to adhesion of the

27
samples to the plastic bag. The coefficients were the lowest in all the directions for the

40% sample, followed by those for the 80% and equilibrated samples. Therefore, the

moisture distribution within the pasta affected its shrinkage. During the early stage of

drying, the moisture content of the sample surface in air with the lowest humidity

decreased more rapidly to form a rigid layer near the surface, which prevented shrinkage

and produced the lowest shrinkage coefficient. The equilibrated sample gradually and

evenly shrank with a decrease in the moisture, resulting in the highest shrinkage

coefficient.

The negative shrinkage strain in the depth direction during the early stage of drying

indicated the expansion of the pasta during this stage. Because the pressure on the pasta

during extrusion was the highest in the depth direction, shrinkage occurred in the width

and height directions from the beginning of drying, but compression relaxation was

predominant over shrinkage in the depth direction, resulting in expansion during the early

stage of drying.

The shrinkage strains for the height and width directions were linearly related. The

shrinkage ratios, which are the slopes of these lines, were 0.93, 0.96, and 0.94 for the

40%, 80%, and equilibrated samples, respectively. Ratios of less than unity indicate

anisotropic shrinkage, that is, the pasta shrank more in the width direction than in the

height direction. This shrinkage anisotropy might come from the network of polymer

constituents, such as gluten and starch. The polymer constituents were stretched in the

28
width direction during extrusion, and their creep recovery occurred rapidly during drying.

A shrinkage ratio of less than unity could be ascribed to this recovery.

The shrinkage coefficient in the longitudinal direction was 0.23 for the shrinkage strain of

the cylindrical 80% sample. The shrinkage strain in the diametric direction linearly

increased with the decrease in the moisture content. The strain mostly increased at

moisture contents of less than 0.17 kg-H2O/kg-d.m. The shrinkage ratio was 0.73 at high

moisture contents, at which the shrinkage strain linearly increased with the decrease in

the moisture content, and this anisotropy indicated that longitudinal shrinkage was

predominant over diametric shrinkage. For moisture contents of less than 0.17

kg-H2O/kg-d.m., the shrinkage ratio significantly increased, indicating an increase of the

shrinkage strain in the diametric direction.

The glass transition temperature of durum semolina was reported to decrease as its

moisture content increased because water acts as a plasticizer.[102] The glass transition of

durum semolina at 50C occurred at a moisture content of 0.17 kg-H2O/kg-d.m. The glass

transition affected the mechanical properties, such as the strength, stress relaxation, and

tensile stress strain.[121,122] Therefore, the major changes in the shrinkage strain and the

shrinkage ratio at moisture contents of less than 0.17 kg-H2O/kg-d.m. could be ascribed

to the glass transition.

The creep relaxation of the polymeric constituents during the early stage of drying

resulted in high shrinkage in the longitudinal direction. On the other hand, during the late

29
stage of drying, longitudinal shrinkage scarcely progressed, but diametric shrinkage

occurred. Therefore, the shrinkage ratio is postulated to become high at low moisture

contents.

A dumbbell specimen of durum semolina dough equilibrated at a specific moisture

content was stretched at 0.5 mm/s to measure the tensile stress as a function of the strain.

Figure 7 shows the Youngs modulus of samples with various moisture contents. The

plots for the 40%, 80%, and equilibrated samples all lay on a single curve. The modulus

increased as the moisture content decreased, and then became almost constant at moisture

contents of less than 0.17 kg-H2O/kg-d.m., where the glass transition occurred.[102]

The lack of dependence of Young's modulus on the drying conditions indicated that the

modulus was not affected by the moisture distribution within the pasta. Therefore, the

hard surface of the pasta seemed to scarcely contribute to its strength against the tensile

strain.

QUALITY EVALUATION OF PASTA BY IMAGE ANALYSIS TECHNIQUES

Quantification Of Surface Roughness Of Spaghetti

The surface roughness of pasta largely influences pasta quality during and after cooking.

Two types of instruments, contact and noncontact, are used to measure the surface

roughness of pasta. The first type of instruments includes contact roughness meters and

atomic force microscopes, whereas the second type includes white light interferometers

and laser-scanning microscopes. Noncontact instruments are preferable for surface

30
observations of soft pasta. A laser-scanning microscope can measure surface geometry to

a high degree of accuracy, which meets the requirements of the International

Organization for Standardization (ISO) and the Japanese Industrial Standards (JIS).

However, this instrument is expensive and measurements of steeply sloped or shaded

samples are difficult owing to a reduction of reflectance or shielding of light.

Objects with rougher surfaces refract more light. Paying attention to this fact, we

developed a novel and low-cost method to rapidly measure the surface roughness of

spaghetti, including the steeply sloped and shaded regions, with a high degree of accuracy,

using a transfer image produced by a digital camera.[126] The surface roughness is

determined using the gray-level frequency distribution from the transmission image taken

with a digital camera.

Figure 8 (a) shows transmission images of dried spaghetti prepared using various

molding dies (bronze, aluminum, polycarbonate, polypropylene, and Teflon) under

illumination by an LED light board.[126] The image was photographed from the other side

of the light board using a digital camera. The surfaces of the spaghetti prepared using the

various dies showed decreasing roughness in the order of bronze, aluminum,

polypropylene, polycarbonate, and Teflon. The frequency distribution of the gray level

was generated from the green image of the pixelated image (Fig. 8 (b)). Gray-level values

can range from 0 to 255 (black to white), and a larger value indicates a whiter surface.

Spaghetti samples with smoother surfaces exhibited larger gray-level values and shaper

frequency distributions because of weak scattering of light on the surface and equable

31
transmission of light through the spaghetti. Spaghetti with rougher surfaces, which were

prepared using the bronze and aluminum dies, had smaller gray-level values and broader

distributions.

Figure 8 (c) shows the frequency distributions, normalized to the maximum profile valley

depth, which were evaluated by measurements with a three-dimensional laser-scanning

microscope. The heights of the frequency distribution peaks were on the same order as

the calculated average roughness (Ra) of the spaghetti surfaces for all the samples tested.

Because the gray-level frequency distributions produced by the digital camera method

were similar in shape to the distributions of the profile valley depths evaluated by

laser-scanning microscopy, the digital camera method should have nearly the same

accuracy as the laser-scanning microscope method. The gray-level frequency distribution

for the smoother spaghetti surface had a higher peak. The dispersion of the gray and

height levels reflected variations in the surface roughness at specific locations. The gray-

and height-level frequencies were symmetric for spaghetti prepared using the Teflon,

polycarbonate, and polypropylene dies, whereas the spaghetti prepared using the

aluminum and bronze dies had asymmetric distributions and tailing at the lower gray

levels. These results showed that the digital camera method could accurately measure

steeply sloped surface geometries, which are difficult to measure with a laser-scanning

microscope.

32
The gray-level height distribution characteristics of the digital camera method were

compared with those of the laser-scanning microscopy method based on the average

(AVE), skewness (SKEW), and contrast (CNT) of each distribution, as calculated using

Eqs. (19), (20), and (21), respectively.

AVE = I P( I ) (19)

SKEW = (P(I ) AVE)3 / (28 VAR3 ) (20)

CNT = I 2 P( I ) (21)

where I is the gray or height level and P(I) is the normalized frequency at I. The variance

(VAR) is defined by Eq. (22).

VAR = ( I AVE)2 P(I ) (22)

The AVE and SKEW of the gray-level distributions correlated linearly with those of the

distributions of the height level, and the CNT for the distribution of the gray level could

be correlated by a power law with that for the distribution of the height level. The

coefficient of determination (R2) of the AVE, SKEW, and CNT were 0.904, 0.942, and

0.958, respectively. Therefore, the surface roughness of spaghetti could be easily

quantified using a digital camera once the calibration curve for a parameter characterizing

the surface roughness had been prepared between the proposed and existing methods.

The digital camera method has several advantages. First, the method does not require

scanning to obtain a two-dimensional transmission images and can be used on samples

with a long side of approximately 10 cm or less. Second, the method takes approximately

33
one-tenth of the time required for laser-scanning microscopy. Finally, the method enables

measurements of the geometry of steeply sloped and shaded surfaces.

Moisture Distribution Within Pasta

When dry pasta is cooked, it increases in color brightness with increasing moisture

content. Based on this fact, we developed a method for measuring the moisture

distribution in spaghetti during its rehydration process by using an image processing

technique.[85] The method consists of the following five steps:

In the first step, one of two sets of rehydrated spaghetti samples was used to measure the

moisture distribution and the other was used to prepare a calibration curve correlating

color brightness and moisture content. The calibration curve was constructed using

samples that had been wrapped in polyethylene sealing film to prevent moisture loss and

then stored at 70C for 3 d to equalize the moisture distribution in the samples.

In the second step, cross-sectional images of the rehydrated samples, which were cut

crosswise using a sharp stainless-steel blade and covered with a light shield with the same

diameter or thickness as that of the sample, were acquired using a digital camera (Fig. 9

inset (a)). Each image contained 3,888 2,592 pixels (1.6 m/pixel), which corresponded

to a spatial resolution of approximately 3.2 m. The resolution was about 20 times higher

than that obtained using MRI methods.[79,82,83]

34
The third step involved digital image processing to pixelate the original 24-bit RGB color

image into red, green, and blue images, and converting the blue image into an 8-bit

gray-scale format (Fig. 9 inset (b)). In order to visually clarify the gray level of the image,

the original gray-level G0 of each pixel was converted to the level G by gamma

correction using Eq. (23), and G was further converted to the level Gc through contrast

correction using Eq. (24).

0.5
G
G 255 0 (23)
255

255
Gc 2 G (24)
2

The fourth step involved preparation of a calibration curve correlating the corrected gray

level Gc (color brightness) with the moisture content using samples with equalized

moisture distributions.

In the final step, the moisture distribution of the rehydrated spaghetti was obtained by

converting the corrected gray level Gc of each pixel in the sample image (Fig. 9 shows

the gray level distribution of spaghetti rehydrated for 10.2 min) to the moisture content

using the calibration curve.

The precise moisture distribution in the al dente state at a rehydration time of 10.2 min

was measured for the first time by the method because the moisture content is too low to

be measured by commonly used methods.

35
REHYDRATION OF DRIED PASTA

Rehydration Kinetics Of Pasta

Rehydration At Constant Temperature

A quantitative understanding of the change in the moisture content of pasta during

rehydration at any temperature is necessary to determine the mechanical properties and

the optimal rehydration time of pasta.[120,127,128] Peleg's model and the Weibull

distribution function can be used to predict the moisture content of pasta at any

rehydration time.[71] However, the temperature dependence of the rehydration kinetics

does not seem to be properly characterized. In particular, the effect of starch

gelatinization on the characteristics of the rehydration kinetics is not fully understood.

Further, the reported equations cannot be directly applied to pasta of different diameters,

even for identical materials. In this section, the effects of the temperature of the

rehydration water on the characteristics of the rehydration kinetics, such as the

equilibrium moisture content and the initial rate of rehydration, are discussed.[129]

Cylindrical pasta (spaghetti) with a carbohydrate content of 72 wt% and initial diameters

of 1.4, 1.6, and 1.8 mm were rehydrated at different temperatures. A reduction in the

spaghetti mass occurred during rehydration because the components leaked into the

immersion water. The loss of spaghetti mass (initial matter (i.m.) basis), Mt, was greater

with longer rehydration times and at higher temperatures. The loss of mass reached

approximately 0.2 kg/kg-i.m., corresponding to approximately 20% of the spaghetti mass.

The amount of spaghetti mass loss in the spaghetti samples with three different diameters

36
could be expressed as a function of the quotient of time and the square of the diameter,

t/d 2, where d is the initial diameter of the spaghetti, by Eq. (25):

kl t
Mt M e 1 exp (25)
d2

where Me is the equilibrium amount of spaghetti mass loss and kl is the rate constant.

The rehydration curves for the spaghetti at 2090C (part of this data is shown in Fig. 10)

could be empirically expressed using a hyperbolic-type equation (Eq. (26)) with t/d2 as

the independent variable rather than the time t.

a (t /d 2 )
X X0 (26)
b (t /d 2 )

where a and b are constants, which were estimated to best fit the calculated X values to

the experimental values. The solid curves in Fig. 10 were drawn using the estimated a

and b values for each temperature.

The equilibrium moisture content at t = , Xe, is given by

Xe lim X a X0 (27)
t

and can be approximated by a because of Xe << a. The initial rate of rehydration, v0, is

derived to be a/b by differentiating X at t = 0 (Eq. (28)).

dX a
v0 (28)
d(t / d 2 ) t 0
b

37
The temperature dependence of the a value obeyed the vant Hoff equation separately in

the high-temperature, transition, and low-temperature regions (Fig. 11).

dln a H
(29)
d(1 / T ) R

where H is the change in the enthalpy of rehydration. The H values in the low- and

high-temperature regions were evaluated as 1.44 and 25.1 kJ/mol, respectively. The

transition occurred at approximately 50C, which was in the middle of the onset

temperature of 45.3C and the peak of 60.5C for starch gelatinization in the DSC

measurement. The temperature dependence of the initial rate of rehydration obeyed the

Arrhenius equation (Fig. 11).[64]

a E
v0 A0 exp (30)
b RT

where A0 is the frequency factor and E is the activation energy. The E and A0 values were

30.5 kJ/mol and 1.36 10-4 m2kg-H2O/(skg-d.m.), respectively. The a/b value showed

no temperature dependence, indicating that the initial rate of rehydration was not affected

by starch gelatinization.

Method For Estimating The Gelatinization Temperature From The Rehydration Curve

Under Constant Temperature-Raising Condition

Temperature greatly affects both the drying of wet food material and the rehydration of

dried foods, including pasta. Starchy foods have greater rehydration capacities at

38
temperatures higher than the gelatinization temperature.[129] Therefore, determination of

the gelatinization temperature of a starchy food is important for reasonable design of its

manufacturing and cooking processes. The methods for estimating gelatinization

temperature include amylography, polarization microscopy, and DSC.[130] Among these,

DSC is most often used.

Based on the facts that the driving force necessary for rehydration of dried starchy food is

large and that the rehydration rate increases at temperatures higher than that of

gelatinization,[131] a method for estimating the gelatinization temperature of

starch-containing foods, without pulverization of the sample, was developed using the

rehydration curve under temperature-programmed heating conditions.[132] The method

was applied to estimate the gelatinization temperatures of udon, kishimen (noodle made

of wheat flour), juwari-soba (noodle made of buckwheat flour), hachiwari-soba (noodle

made from a mixture of buckwheat and plain wheat flour at a weight ratio of 8:2),

kuzukiri (noodle originally made of ground arrowroot, but today made of potato starch),

and Malony (the main component of which is potato starch). The rehydration curves

obtained for pasta at various temperature-raising rates had inflection points at which the

rehydration rate increased markedly, and this inflection-point temperature corresponded

to the gelatinization temperature. The temperature-raising rate scarcely affected the

inflection-point temperature. The gelatinization temperatures estimated by the proposed

method were compared with those estimated by DSC.

39
Figure 12 shows plots of the amounts of water rehydrated against temperature for all the

noodles tested. The inflection-point temperature of low-temperature (LT) pasta was

53.1C, which was almost the same as that of HT pasta (52.3C). These temperatures are

close to the gelatinization temperature of durum wheat semolina.[133] The inflection-point

temperatures of udon and kishimen were 57.0 and 57.8C, respectively, which are close

to the previously reported values.[134] The inflection-point temperatures of kuzukiri and

Malony, which are rich in potato starch, were 49.1 and 48.4C, respectively. The

inflection-point temperatures of juwari-soba, hachiwari-soba, and common soba were

61.1, 59.6, and 57.4C, respectively. The inflection-point temperature was higher for soba,

which has a higher content of buckwheat flour.

The DSC curves were determined for all of the noodles, except for Malony. The

inflection-point temperatures observed by the proposed method were between the onset

and peak temperatures in the DSC measurements. The gelatinization temperature of a

starchy food depends somewhat on the measurement technique.[130] Hence, the

inflection-point temperature should reflect the gelatinization temperature of each type of

noodle.

Effect Of Salts On Rehydration Kinetics

Because dried pasta is typically cooked in water containing 0.11% (w/v) NaCl, the

effect of salts on the rehydration kinetics of dried spaghetti was measured at 2090C in

1.83 mol/L NaCl, equivalent to 10% (w/v), as well as in 1.83 mol/L LiCl, KCl, NaBr, and

NaI. Although salts other than NaCl and extremely high salt concentrations are unrealistic

40
for cooking, these conditions were evaluated to understand the role of NaCl and other

salts in the rehydration process of pasta.[135]

The rehydration processes in 1.83 mol/L NaCl at every temperature could be expressed

by Eq. (26), and the a and a/b values of Eq. (26) were estimated at each temperature. The

temperature dependences of the a and a/b values were similar to those in water in the

previous section and were analyzed using Eqs. (29) and (30), respectively (Fig. 11). The

H values in the low- and high-temperature regions were 1.08 and 33.1 kJ/mol,

respectively. The H values in the low- and high-temperature regions in the NaCl

solution were nearly equal to and higher than those in water, respectively. As Chiotelli et

al.[136] reported that the gelatinization enthalpy in approximately 2.0 mol/L NaCl was

larger than that in water, the difference between H in the high-temperature region in the

NaCl solution and that in water should reflect the difference in gelatinization enthalpy.

The temperature dependence of the a/b value, which corresponds to the initial

rehydration rate, as shown by Eq. (27), in 1.83 mol/L NaCl was also expressed by Eq.

(29) in the tested temperature range, and the activation energy was evaluated as 25.6

kJ/mol, which was slightly lower than the value in water.

The equilibrium moisture content, a, was estimated for rehydration at 55 and 60C in

various salt solutions. The a value for rehydration in salt solutions of the alkaline metal

ions with chloride was in the order of LiCl < NaCl < KCl, and that in sodium salt

solutions containing different halogen ions was in the order of NaCl < NaBr < NaI. These

orders are opposite to the Hofmeister (lyotropic) series.

41
Based on Ficks law of diffusion, an apparent diffusion coefficient, Da, can be estimated

for cylindrical pasta according to the following equation:[107]

2
X Xe 1 n Da t
4 2
exp (31)
X0 Xe n 1 n (d / 2)2

where n is the Bessel function root (1 = 2.4048) and X0, Xe, and X are the initial

moisture content, the equilibrium moisture content, and the moisture content at time t,

respectively. For long times or small diameters, the Da value could be estimated by

plotting (X Xe)/(X0 Xe) versus time t for rehydration.[71] The Da values were obtained

at 60C in the 1.83 mol/L LiCl, NaCl, and KCl solutions, and the Da value was higher for

salts of the alkaline metal ions with chloride and a larger crystal radius; however, this

trend is not reasonable. Alkaline metal ions appear to migrate in water in the hydrated

state, and the Da value in the salt solution with hydrated ions with larger Stokes radii[137]

was lower. This result indicates that hydration of the ions plays an important role in water

migration in pasta.

Effect Of Dies On Rehydration Kinetics

Although pasta has been traditionally prepared using die made of bronze, die made of

Teflon have recently been increasingly used because of the longer lifetime of the die, the

smoother surface of the pasta, and the good appearance of the dried pasta.[3-5] The die

material affects the surface roughness of the pasta, as well as the properties of the pasta.

Therefore, the effects of the die material on the properties of raw pasta and the

rehydration behaviors of dried pasta were examined.[138]

42
Spaghetti samples with varying surface roughnesses were prepared using bronze,

aluminum, polypropylene, polycarbonate, and Teflon dies. The orifice diameter and

length of each die were 1.8 and 5 mm, respectively. The raw spaghetti was dried to a

moisture content of approximately 0.12 kg-H2O/kg-d.m. in a temperature-humidity

controlled chamber. We measured the surface images of dried spaghetti prepared using

the various dies at 200- and 1000-fold magnifications.[138] The observations at 1000-fold

magnification revealed that the surface of the spaghetti prepared using the Teflon die had

the smoothest surface, while the smoothness of those prepared using the polypropylene,

polycarbonate, aluminum, and bronze dies decreased in this order.

The spaghetti with a smoother surface had a faster extrusion velocity for passing through

the die, a higher bulk density, and a greater breaking strength after drying, which was

measured using a creep meter with a wedge-shaped plunger.[138] Moreover, the die

material did not affect the gelatinization temperature measured by DSC.[138]

The rehydration process of the spaghetti, which had been cut into 8 cm-long sticks, was

measured in 0.5% (w/v) NaCl at 99.7 0.3C. The amount of water rehydrated per unit

surface area, x, was calculated using Eq. (32).

w2 w1
x (32)
S

where w1 and w2 are the sample weight before and after rehydration, respectively, and S is

the surface area of the stick.

43
The rehydration curves for spaghetti prepared using the various dies were measured.[138]

Because the early stage of rehydration could not be well expressed by Eq. (26), only the a

value was estimated using Eq. (26). The a value was almost the same for all of the

spaghetti samples prepared using the dies made from various materials. Since the initial

diameter (surface area) affects the rehydration velocity, rehydration during the early stage

(within 60 s) was characterized based on the x value defined by Eq. (32). The x value was

correlated with t by a quadratic equation, and the momentarily rehydrated amount of

water when the sample was immersed in the boiling water, x0, was estimated by

extrapolating the equation to t = 0. This amount was higher for spaghetti with rougher

surfaces.

Thus, the die materials affected the physical properties of spaghetti, i.e., surface

roughness and density, but not the chemical properties, i.e., starch gelatinization.

Therefore, the momentarily rehydrated amount of water, not the equilibrium amount of

water, depended on the die material.

Effect Of Surface Roughness On Rehydration Kinetics

The water-sorption kinetics was measured for spaghetti prepared using dies made of

bronze, aluminum, polycarbonate, polypropylene, and Teflon.[138] When dried spaghetti

was rehydrated, very rapid water intake was observed. The momentarily rehydrated

amount of water, Xt0, was estimated by extrapolating the amounts of water sorbed per

unit surface area in the initial 60 s, which was a good indication of the initial water

sorption amount.[138] Although the Xt0 value was larger for spaghetti with a rougher

44
surface,[138] an explanation for this observation remains to be elucidated. The relation

between the momentarily rehydrated amounts of water per unit area and the surface area,

which was evaluated from three-dimensional laser microscope measurements, was

quantitatively examined.[139]

Figures 13 (a) and (b) show the surface images of the spaghetti prepared with different

dies, which were measured using a three-dimensional laser microscope. The surface

roughness was expressed by the average roughness, Ra, as determined using Eq. (33).

1 l
Ra g ( x)dx (33)
l 0

where l is the reference length and g(-) is the roughness curve. The Ra values of spaghetti

decreased in the following order: spaghetti prepared using Teflon, polypropylene,

polycarbonate, aluminum, and bronze. Starch gelatinization would not affect the Xt0

value because this value is estimated by extrapolating the amount of water sorbed to 0

min. Furthermore, the die material does not affect the gelatinization temperature of starch

in spaghetti. Therefore, the Xt0 value should be affected only by the surface roughness

or inner structure of spaghetti.

Spaghetti with a rougher surface had the smallest bulk density.[138] However, the bulk

density of spaghetti prepared using the Teflon die was 1.36 0.01 g/mL, which was the

largest among the tested spaghetti samples and was similar to that (1.31 0.02 g/mL) of

spaghetti prepared using the bronze die with the lowest bulk density. Because water

45
diffusivity within spaghetti was approximately 10-10 m2/s,[71] the Xt0 value was not

controlled by the inner structure, but by the surface roughness.

The momentarily rehydrated amounts per unit area were calculated based on two

different surface areas for spaghetti: one area was evaluated using 2rLs, where r and Ls

are the radius and length of the spaghetti sample, respectively, under the assumption that

the spaghetti was a cylinder with a smooth surface, and the other was measured using a

laser microscope. The roughness surface was divided into small triangles. For a triangle

with vertex coordinates of A(xA, yA, zA), B(xB, yB, zB), and C(xC, yC, zC), its area S was

calculated by Eq. (34).

2 2 2
xA xB xC yA yB yC zA zB zC
1
S det yA yB yC det zA zB zC det xA xB xC (34)
2
1 1 1 1 1 1 1 1 1

The surface area was calculated by summing the S values of all the triangles. Xt0

(smooth) and Xt0 (rough) denote the amounts determined based on the former and latter

areas, respectively. Although Xt0 (smooth) was larger for spaghetti with a rougher

surface and larger Ra value, Xt0 (rough) did not depend on the Ra value (Fig. 13 (c)).

Both the initial water-sorption rate and the equilibrium amount of sorbed water determine

the cooking time for the optimal rehydration of spaghetti.[129,135] It has been reported that

chemical modification of the raw material controls the equilibrium of the sorbed water;

for example, by modifying the starch gelatinization properties.[129,135,140] Thus, the initial

water-sorption rate could be controlled by modifying the surface roughness of spaghetti.

46
Mechanism Of Rehydration

Moisture Profiles

Figure 14 shows the moisture distributions of spaghetti at different rehydration times.[85]

The rehydration time of 0 min represents dried spaghetti. Swelling occurred in an

approximately 0.2 mm region near the surface during the first 1 min. During this period,

the moisture content remained at the same level as that of dried spaghetti in the inner

region. Water quickly penetrated only near the surface, as indicated by the change in the

distribution between 0 and 1 min. Water likely entered the spaghetti through small

surface holes and cracks present on the surface of the pasta.[141,142] The flat moisture

distribution near the surface suggested that water imbibition during the early stage was

not attributable to water diffusion, but rather to water filling the holes and cracks.

Gelatinization was reported to be complete within 1 min of assessing the extent of

gelatinization.[131] That is, starch gelatinization is a fast process. However, the moisture

content on the surface gradually increased over time. Because the average moisture

content of spaghetti was approximately 9 kg-H2O/kg-d.m., as shown in a previous section,

the gelatinization of starch granules in the sample did not reach equilibrium, even after 20

min of rehydration, and swelling of the granules was restricted. The starch granules are

deeply embedded in the protein matrix of the dried pasta, and the structure below the

surface appears to be a honeycomb-like structure after 4 min of rehydration.[141,142] The

structure of the protein matrix shrinks and relaxes during drying and rehydration.

Therefore, gradual swelling of the starch granules owing to gelatinization occurred,

together with relaxation of the protein matrix during rehydration. Structural relaxation of

47
the protein matrix plays an important role in moisture content changes during water

migration.

The moisture distributions of the samples rehydrated for 10.2 min or longer showed a flat

moisture distribution near the surface and a parabolic distribution of the moisture content

in the inner region. Water migration in the inner region should be diffusion-controlled,

while the diffusion of water was not the rate-limiting step for rehydration near the surface.

Because the moisture content near the surface of the sample rehydrated for 20 min was

much lower than the equilibrium content, most of the water supplied from the bulk phase

is sorbed by the starch granules to increase the moisture content near the surface. In the

inner region, water diffused according to the gradient in the moisture content to gradually

shift the distribution to the higher level. However, the consumption of water by starch

granules sorbed near the surface restricted the penetration of water into the inner region.

Water migration near the surface may bring about the expansion of the region with the

flat moisture distribution.

Relaxation Of The Gluten Network

As shown above, the moisture content on the surface of pasta gradually increased rather

than quickly reaching equilibrium. The non-Fickian diffusion of a low-molecular-weight

molecule into a matrix of high-molecular-weight molecules is attributed to the slow

reconfiguration of high-molecular-weight molecule segments after accepting the

penetrant,[143] which has been observed in many high-molecular-weight

polymer-penetrant systems.[144] Therefore, the diffusion of water and the relaxation of

48
wheat flour components should play important roles in water migration within pasta.

However, water migration in dried foods has been extensively studied based on water

diffusion, but little attention has been given to the relaxation of high-molecular-weight

molecules. We kinetically examined the rehydration process of pasta with an infinitely

small diameter, where the effects of water diffusion on rehydration were negligible, in

order to elucidate the role of the relaxation of high-molecular-weight molecules (wheat

flour components) in rehydration.[145]

Rehydration curves were measured at 98C for spaghetti with different diameters. The

moisture content of infinitely thin durum pasta, designated as 0 mm durum pasta, was

estimated by plotting the average moisture content at a specific time against pasta

diameter and extrapolating to 0 mm, as shown in the inset of Fig. 15. The moisture

content of 0 mm durum pasta was estimated from the y-intercept of the line, and the

moisture content for 0 mm durum pasta, which was normalized by the content at

equilibrium, was plotted versus the rehydration time in Fig. 15.

Gluten was isolated from durum wheat flour[98] to prepare gluten pasta. The gluten pasta

was placed in a drying chamber to prepare the dried pasta. The rehydration process of the

gluten pasta was measured under the same conditions as for the durum pasta (Fig. 15).

The processes for both gluten pasta and 0 mm durum pasta could be expressed by Eq.

(35).[146]

X X0
1 exp( kr t ) (35)
Xe X0

49
where kr is the rate constant. Although the Xe value of 0 mm durum pasta was much larger

than that of the gluten pasta, the kr value of 0 mm durum pasta (7.53 10-4 1/s) was

nearly the same as that of gluten pasta (7.42 10-4 1/s). The similarity in the k values for

0 mm durum pasta and gluten pasta indicates that the rehydration kinetics at the pasta

surface is governed by the relaxation of the gluten network, in which starch granules are

embedded. The starch granules swell via gelatinization immediately after water intake

from their surroundings because of the high gelatinization rate of starch. However, the

gluten network prevents the swelling of starch granules. Therefore, the moisture content

near the surface gradually increased with gradual relaxation of the gluten network.

Rehydration And Texture

The palatability of food is affected by its flavor, taste, and texture. Texture plays a

particularly important role in the palatability of noodles. The moisture distributions and

stress-strain curves of spaghetti rehydrated to its optimal cooking state, known as al dente,

at 60, 80, and 100C in distilled water or 0.1, 1.0, and 2.0 mol/L NaCl solutions were

examined to identify the major factors governing the texture of rehydrated spaghetti. The

texture of al dente pasta is characterized by a soft external zone and a firm core, with the

firm core providing a mouthfeel of resistance while biting. The factor affecting the

texture of rehydrated pasta was previously reported to be the moisture distribution based

on the relationship between the moisture distribution of pasta, as measured using MRI,

and its mechanical properties, as measured using a texture analyzer.[82] Although the

average moisture content is lower at lower rehydration temperatures[128] and at higher

NaCl concentrations,[135] the effects of the rehydration temperature and NaCl

50
concentration on the moisture distribution and texture of rehydrated pasta are not fully

understood. The moisture distributions and stress-strain curves of spaghetti (initial

diameter of approximately 1.6 mm) rehydrated to the al dente state (average moisture

content of 1.70 0.05 kg-H2O/kg-d.m.) at various temperatures and NaCl concentrations

were measured.[147]

Figure 16 shows the moisture distribution in spaghetti rehydrated at 60, 80, and 100C to

the al dente state. The spaghetti rehydrated at 80 and 100C had lower moisture contents

near the center than in the peripheral region, and the moisture distribution exhibited a

concave shape. The moisture content of the spaghetti rehydrated at 60C exhibited a

nearly flat distribution across the entire cross-section, although the moisture content

decreased slightly around the peripheral region. The loss of spaghetti mass increases

exponentially with time.[129] Rehydration at 60C to reach the al dente state proceeded

more slowly than rehydration at the other temperatures. Therefore, the increased loss of

spaghetti mass during rehydration at 60C resulted in a slightly convex or nearly flat

moisture distribution. The difference in moisture content between the inner and

peripheral regions was greater in spaghetti rehydrated at higher temperatures.

Figure 17 shows the stress-strain curves of spaghetti rehydrated at 60, 80, and 100C. The

spaghetti rehydrated at higher temperatures showed smaller and larger stresses under

lower and greater strains, respectively. The rehydration temperature affected the breaking

stress, but not the breaking strain. Spaghetti rehydrated at higher temperatures exhibited

higher breaking stress. The stress-strain curves generated at lower and greater strains

51
reflect the mechanical properties of the spaghetti surface and of the material near the

spaghettis core, respectively. The characteristics of the curve for the spaghetti rehydrated

at higher temperatures came from higher and lower moisture contents near the surface

and at the center, respectively. The difference in the moisture content brought about

compression of the spaghetti at lower and greater stresses under lower and greater strains,

respectively. The similarity in the breaking strain of all the spaghetti samples with

different moisture distributions indicates that the moisture distribution does not affect the

breaking strain, but rather controls the breaking stress.

The moisture distributions of spaghetti rehydrated at 100C in 0, 0.1, 1.0, and 2.0 mol/L

NaCl solutions to the al dente state were also measured. No significant effect of the salt

concentration on the moisture distribution of al dente spaghetti was observed. As

mentioned above, water migration in spaghetti is determined by starch gelatinization,

water diffusion, and gluten matrix relaxation,[85,145] all of which are affected by the

rehydration temperature. The salt concentration affects the gelatinization temperature, but

does not directly affect the diffusion coefficient of water. The difference in the moisture

distribution of spaghetti rehydrated at different salt concentrations is ascribed to the

difference in the sensitivity of the salt concentration to starch gelatinization.

The slope of the stress-strain curve in the region of small strains was larger at higher salt

concentrations, indicating that spaghetti rehydrated at higher salt concentrations is more

elastic. The salt concentration only minimally affected the breaking stress, while the

breaking strain was lower in spaghetti rehydrated at higher salt concentrations. The low

52
strain may reflect a reduction in a property equivalent to the stickiness of the spaghetti.

No significant difference in the breaking stress was observed among all the spaghetti

samples rehydrated in salt solutions, as they all had nearly the same moisture distribution.

The gluten framework of spaghetti embeds starch granules,[141,142] and salt changes the

molecular conformation of gluten to a fibrous network during dough preparation.[148] The

addition of NaCl altered both the interactions and distances between glutenin and gliadin

in gluten during dough preparation, with the distances being shortened,[149] causing a

reduction in gluten volume.[150] Although these effects of NaCl were observed during

dough preparation, similar effects may be observed during rehydration in the presence of

NaCl. Rehydration at higher salt concentrations may contract the gluten network to lower

the breaking strain.

EFFECTS OF DRYING CONDITIONS ON THE QUALITY OF COOKED PASTA

Characterization Of Programmed-Drying And Constant-Drying Conditions

The effects of temperature and humidity during drying on various pasta properties,

including color, surface structure, breaking strength, rehydration characteristics during

cooking, texture, and sauce retention capacity, were examined.[140,151153] The dried

spaghetti samples were prepared under the conditions where the temperature and

humidity were alternately set as independent variables, although crack generation during

drying was inevitable under these conditions of constant temperature and humidity. The

properties of pasta dried under constant-drying conditions were compared with those of

pasta dried under programmed-drying conditions (denoted as LT, HT, and VHT).

53
Figure 18 shows four examples of the drying curves. The drying behavior at 50C and

50% RH was similar to that at 85C and 80% RH. Raw spaghetti is usually dried to a

dry-matter-based moisture content of 11%13% to prevent microbial deterioration during

storage. Under the most severe conditions of 85C and 50% RH, drying was complete in

ca. 50 min. On the other hand, drying to the required moisture content could not be

achieved under the mildest conditions (50C and 80% RH), even after 15 h. Based on the

moisture sorption isotherm of durum semolina wheat flour, the equilibrium amount of

water sorbed under these conditions was 0.15 kg-H2O/kg-d.m.[91]

The effects of temperature and relative humidity on the thermogravimetrically

determined constant drying rate, Rc, and the rate constant during the decreasing

drying-rate period, ks, for a cylindrical tablet of durum semolina dough[16] were shown in

a previous section. Both Rc and ks were more affected by temperature than relative

humidity, being larger at 85C and 80% RH than at 50C and 50% RH. On the other hand,

the equilibrium amount of sorbed water at 85C and 80% RH and at 50C and 50% RH

was almost the same. The similar drying behavior under these two conditions is attributed

to the almost identical equilibrium amount of sorbed water, despite the somewhat shorter

drying time at 85C and 80% RH owing to a faster drying rate during the early stage.

Product color is an important aesthetic factor for the selection of goods by consumers.

The product color was assessed by the absorption spectra of the methanolic extracts of

dried spaghetti. Every extract exhibited a trimodal spectrum with peaks at 440 and 466

nm in the visible light range. The absorbance at 466 nm showed a temperature and

54
humidity dependence similar to that at 440 nm. Although temperature and humidity did

not explicitly affect spaghetti color, the absorbance of the spaghetti extract prepared at

50C tended to be lower at a higher relative humidity. The absorbance of the spaghetti

extract prepared at 70C showed the highest values at 60%70% RH, whereas the relative

humidity scarcely affected the color of spaghetti prepared at 85C. At a low drying

humidity, the color was hardly affected by temperature. On the other hand, temperature

had a large effect on the color of spaghetti prepared at high humidity. The highest

absorbance was exhibited by the spaghetti extract prepared at 70% RH and 70C. The

absorption spectra of the extracts of spaghetti prepared under three different

programmed-drying conditions, designated as LT, HT, and VHT, where the maximum

temperature and total drying duration were 50C and 20 h, 70C and 11 h, and 85C and

6 h, respectively, were also measured. The humidity was between 70% and 80% RH for

all LT, HT, and VHT conditions. The absorption spectra of the spaghetti prepared under

the conditions of constant temperature and humidity were similar to that of LT spaghetti,

but different from those of HT and VHT spaghetti. Although the color of spaghetti

prepared under programmed-drying conditions was mainly discussed in terms of the

maximum temperature during drying, the maximum temperature was not the sole factor

affecting the color, especially at high temperatures. It is reported that pasta dried at high

temperatures exhibited a deeper brown color,[53,54,58,154] but the drying temperature is also

reported to have no effect on the color of dried pasta.[155] Because a high drying

temperature during the early stage inhibited the progress of the Maillard reaction,[57,58] the

humidity hardly affected the color development during drying at 85C.

55
The AFM surface images of spaghetti dried under programmed-drying conditions were

obtained (Table 2), as well as the images of spaghetti dried under constant-drying

conditions (Fig. 18 (a)(d)). The spaghetti dried at 85C had crater-like surface dents.

More dents were observed for spaghetti dried at lower relative humidities, while spaghetti

dried at 50C had fewer dents. The LT, HT, and VHT spaghetti also exhibited dents on

their surfaces.[152] Therefore, the occurrence of dents was not specific to drying at

constant temperature and humidity. The pore size distribution near the spaghetti surface

was measured using a mercury porosimeter. The pores smaller than 0.2 m were not

affected by compression during the measurement, and the pores larger than 1.0 m

included the effects of voids among the samples and chipping, fissure, and asperity of the

sample surface. The LT, HT, and VHT spaghetti prepared under programmed-drying

conditions showed monomodal pore size distributions in the range of 0.110 m, but the

spaghetti prepared at constant temperature and relative humidity did not. The latter might

be attributed to the very small cracks generated by the shrinking of spaghetti owing to the

rapid evaporation of water during drying, which results in the inner stress and strain

exceeding the critical values. The generation of cracks is especially promoted by rapid

dehydration at an early drying stage. Since the maximum relative humidity in the

practically used programmed-drying processes is between 70% and 80% RH,

microscopic cracks should be hardly generated under mild drying conditions, such as at

50C and 80% RH, and a pore size distribution similar to that of spaghetti dried under

programmed conditions is obtained. The pore size distribution (3.0 103 to 1.8 102

m) for 50% RH was similar to that for 80% RH, irrespective of temperature. Therefore,

the relative humidity affected pore generation more than temperature. However, the

56
temperature affected the pore size distribution in the range of 0.110 m at a high relative

humidity of 80% RH. The pore size distribution of spaghetti prepared at 85C was flatter

than that of spaghetti prepared at 50C. The spaghetti prepared at 85C had crater-like

dents, with diameters of around 0.51.0 m, but the remainder of the surface was very

dense. When spaghetti was dried at 50% RH, the fraction of pores larger than a few

micrometers rapidly increased. On the other hand, such large pores or cracks were not

observed in the AFM images. Therefore, such large pores are generated progressively

under reduced pressure during the pretreatment phase of the mercury porosimetry

measurements. Microscopic cracks were generated during drying at low humidity, which

progressively developed to form large pores.

The breaking force and strain of spaghetti were measured immediately after preparation

and after storage for 3 weeks. The breaking force was larger for spaghetti produced at

high temperature or low humidity immediately after preparation. On the other hand, the

breaking strain was not affected by the drying temperature and humidity. Spaghetti

prepared under programmed-drying conditions exhibited no significant difference in the

breaking strain, while VHT spaghetti had the largest breaking force, followed by HT and

LT spaghetti.[152] The breaking force and strain were almost equal to those of spaghetti

prepared at constant temperature and relative humidity. These facts indicated that the

visible cracks generated during drying at constant temperature and relative humidity

hardly affected the breaking behavior. The larger breaking force of spaghetti prepared at

high temperature is caused by the promotion of protein denaturation and resultant firm

protein network. The temperature and humidity dependences of the spaghetti stored for 3

57
weeks were the same as and contrary to those of the spaghetti immediately after

preparation, respectively. Similar to the spaghetti immediately after preparation, no

significant effects of drying temperature and humidity on the breaking strain were

observed for the spaghetti stored for 3 weeks. The breaking strain was markedly small for

the spaghetti prepared at low humidity. Because the spaghetti dried at 50% RH had many

pores of over a few micrometers, the cracks seemed to develop during storage, resulting

in a fragile spaghetti structure. The spaghetti dried at low humidity exhibited a larger

breaking force immediately after preparation owing to a denser surface structure, which

was formed by the more rapid dehydration at low humidity. On the other hand, after

storage for 3 weeks, the cracks generated during drying at low humidity and high

temperature developed, lowering the breaking force. On the contrary, the spaghetti

prepared at high humidity was further dried during storage to lower the content of water,

which acts as a plasticizer, resulting in an increased breaking force.

The spaghetti dried at 50% RH was rehydrated more than that dried at 80% RH,

irrespective of the drying temperature. When compared at the same drying humidity, the

spaghetti dried at a lower temperature absorbed more water than that dried at a higher

temperature. The rehydration processes could be expressed by Eq. (26)[129,135] for

spaghetti prepared under any conditions. Similar to previous results for LT, HT, and VHT

spaghetti,[140] no significant effect of the drying temperature on initial rehydration rate

was observed. The v0 value, defined by Eq. (28), was smaller for spaghetti dried at high

relative humidity. When cracks were generated during drying, water very quickly

penetrated the cracks. Since small cracks are prone to be generated at low relative

58
humidity, the spaghetti dried under these conditions exhibited the highest initial

rehydration rate. Both the drying temperature and humidity showed no significant trend

in their effects on the equilibrium moisture content. The equilibrium moisture content is

governed by the gelatinization of starch.[129,135] The effect of starch gelatinization on Xe

was decreased at a high rehydration temperature. Although the spaghetti prepared at

different temperatures and humidities had different gelatinization states of starch, no

significant effects of drying temperature and relative humidity on Xe were observed for

rehydration at 100C.

The texture of spaghetti prepared at various temperatures and humidities and rehydrated

at 100C was characterized using a creep meter. The spaghetti dried at high temperature

or high relative humidity exhibited greater hardness after rehydration. For spaghetti dried

at low relative humidity, the cohesiveness of the rehydrated spaghetti prepared at high

temperature was lower, but the drying temperature did not affect the cohesiveness of the

spaghetti dried at high relative humidity. The adhesiveness of the rehydrated spaghetti

was higher for spaghetti dried at lower temperature or lower humidity. However, the

adhesiveness of the rehydrated spaghetti was not simply governed by the maximum

drying temperature. The effects of drying temperature and humidity on the stickiness of

the rehydrated spaghetti were similar to those on adhesiveness. Drying at constant

temperature and humidity elucidated the effects of drying temperature and humidity on

stickiness. Extensive starch swelling and amylose leaching onto the spaghetti surface are

involved in determining both adhesiveness and stickiness.[44,152] Drying at high

temperature restricts extensive starch swelling and amylose leaching onto the spaghetti

59
surface owing to the increase of the starch gelatinization temperature by high-temperature

drying. Because the generation of more cracks at lower humidity facilitates amylose

leaching through the cracks, the drying temperature and humidity affect the adhesiveness

and stickiness of the rehydrated spaghetti.

Sauce Retention Of Cooked Pasta

The sauce retention capacity was measured for rehydrated spaghetti prepared at various

drying temperatures and humidities (Table 2). More sauce seemed to be retained by the

spaghetti dried at higher temperature or higher humidity. Spaghetti with a rougher surface

retains more sauce. The spaghetti dried at lower temperature has a smoother surface.[152]

During rehydration, more amylose leaches onto the surface of the spaghetti dried at lower

temperature, covering the surface to form a thin gelled film. This amylose coverage

makes the surface even smoother. Although the spaghetti dried at lower humidity had a

rougher surface, it exhibited a smaller sauce retention capacity. On the other hand, the

effects of drying temperature and humidity on the sauce retention capacity were opposite

to those on adhesiveness and stickiness, which are greatly related to amylose leaching.

Therefore, amylose leaching predominantly governs the sauce retention capacity. It has

been generally believed that spaghetti extrusion-molded through a bronze die has a

rougher surface and retains more sauce. However, the surface roughness has no

significant effect on the sauce retention capacity, which is instead governed by amylose

leaching during rehydration. Therefore, drying of spaghetti at higher temperature and

higher humidity can help retain more sauce after cooking and provide better palatability.

60
R & D NEEDS AND OPPORTUNITIES

Due to recent diversification in food products, the following properties are required in

dried pasta, in addition to palatability and suitability for home cooking: distinctive texture

(such as hardness, softness, and adhesiveness), tolerability to microwave heating,

long-term storage in the form of frozen and chilled foods, and requirement of reduced

cooking time. Technological development to realize such demands is necessary in the

steps involving the selection of raw material, molding, and drying. The drying step,

which is the final step in dried pasta manufacturing, plays a particularly important role in

determining the properties of the pasta.

Because reductions in cooking time and complexity are desired by consumers,

development of instant and microwaveable noodle products is important. Operational

conditions in the manufacture of instant noodles using steam or superheated one affect

product qualities such as their texture, color, rupture behavior, and starch

gelatinization[156] and have been extensively studied to improve the properties of the

noodles.[157-160] Microwave-assisted drying of food has also been reported.[161,162]

Consumption of easily cooked foods is likely to increase in the future. Therefore, studies

on the effects of drying conditions on the qualities of microwaveable foods are necessary.

Manufacturing dried pasta is a process industry and its operational conditions have

been determined empirically. Elucidation of water-migration mechanisms and changes in

starch and proteins during drying enables us to rationally determine the drying conditions.

Knowledge of the relationship between water migration during pasta cooking and the

61
texture of the cooked product is useful for industrially producing pasta with high

palatability as well as suitability for home cooking.

A simple and low-cost method to measure the moisture profile within pasta would be

useful for visualizing the rehydration processes of other dried foods.

CLOSING REMARKS

The moisture sorption isotherms of durum semolina can be expressed by the

GuggenheimAndersonde Boer (GAB) equation. A slight hysteresis was evident in the

desorption isotherm when it was superimposed over the sorption isotherm. The partial

molar volume of water sorbed onto durum semolina was measured by dilatometry. It

increased with increasing moisture content and reached a constant value of ~17.5

cm3/mol at moisture contents of 0.2 kg-H2O/kg-d.m. or greater. The moisture content was

the same as that at which glass transition of the durum semolina occurred. These results

indicate that the sorbed water molecules in the multilayer region behaved similarly to the

molecules in bulk water because of their very weak interactions with the durum semolina.

The drying characteristic curve of durum semolina dough, which was obtained by

thermogravimetric analysis, showed that ~20% of the water evaporated during the

constant-drying-rate period, which has been ignored in previous studies. This

demonstrated that the constant-drying-rate period should be taken into account in order to

predict drying curves with high accuracy. The drying rate during the constant-drying-rate

period and the mass transfer coefficient estimated by the thermogravimetric analysis were

62
expressed as functions of the temperature and relative humidity, and were useful for

predicting the drying processes of pasta under any drying conditions, including

industrially applied ones. Analysis using a combination of thermogravimetry and

differential scanning calorimetry (DSC) under linearly increasing temperature conditions

revealed that the conclusion temperature of the endothermic peak and the temperature of

the inflection point of the drying characteristics curve were located near the glass

transition curve of the durum semolina. Moreover, the glass transition significantly

affected the mechanical properties of pasta during drying; for example, the Youngs

modulus of the dumbbell pasta specimen decreased with decrease in the moisture content

and became almost constant at moisture contents lower than 0.17 kg-H2O/kg-d.m.,

although the modulus did not depend on the drying conditions.

Empirical equations were proposed to express the rehydration characteristics of pasta,

such as the moisture content and loss of pasta mass, which determine the quality loss of

cooked pasta and affect its mechanical properties. The relationship between the moisture

content and rehydration time can be described empirically by the hyperbolic equation,

regardless of the drying conditions, rehydration temperature, or salt content of the

rehydrated solution. The hyperbolic equation contributed to the definition of the

rehydration kinetics; the initial rate of rehydration is not affected by starch gelatinization.

The apparent diffusion coefficient of water into pasta was not correlated with the crystal

radius of the salts in the rehydrated solution, but was correlated with the Stokes radii of

the hydrated ions. The equilibrium moisture content is limited by the state of starch

gelatinization. Therefore, the content can be indexed using the Hofmeister series of ions

63
in the rehydrated solution. Starch gelatinization is usually determined by DSC. However,

a novel method in which the rehydration curve is constructed under linearly increasing

temperature conditions was proposed to estimate the gelatinization temperature of various

noodles, including pasta. The gelatinization temperatures estimated by this method were

between the onset and peak temperatures obtained by DSC.

Pasta is prepared using dies made of various materials. The surface of the pasta

became increasingly rough when prepared using the Teflon, polypropylene,

polycarbonate, aluminum, and bronze dies, respectively. The die materials affected the

physical properties of extruded pasta, such as the bulk density and rupture strength, but

did not affect the chemical properties, such as the starch gelatinization. The momentarily

rehydrated amount of water, which is a hypothetical quantity, can be used to characterize

the initial water intake. When based on a hypothetically smooth surface, this amount was

larger for pasta with greater average roughness. However, when based on a rough surface,

it did not depend on the average roughness. This showed that regulation of the surface

roughness of pasta is effective in controlling the initial rehydration rate. In this context, a

novel, rapid, and low-cost method, based on the principal that an object with a rougher

surface refracts more light, was developed to measure the surface roughness of pasta

using images acquired with a common digital camera. This method determines the

surface roughness using gray-level frequency distributions from transmission images and

facilitates the measurement of surface geometry to the same accuracy as that of

traditional, high-performance, laser-scanning microscopy techniques.

64
Knowledge regarding transient changes in the moisture profiles of pasta during

rehydration is useful to understand the mechanism of rehydration; therefore, a novel

method using an image processing technique was developed to measure the moisture

profiles. This method is based on the increase in sample brightness with increasing

moisture content. Compared to currently used methods, the method has the advantage

that moisture contents of around 0.1 kg-H2O/kg-d.m. can be easily measured at a spatial

resolution of 3.2 m. The moisture profiles obtained by this method suggested that

penetration of water into small holes and cracks in the pasta surface, water diffusion in

the pasta, and structural relaxation of the protein matrix play important roles in the

rehydration mechanism. Additionally, the rehydration kinetics of hypothetical, infinitely

thin, pasta, where the effects of the diffusion of water on the moisture content were

negligible, suggested that the swelling of starch by rapid gelatinization was restricted by

the structural network of gluten and relaxation of the gluten network during pasta

rehydration. In addition to the moisture profiles, stressstrain curves were examined to

identify the major factors governing the texture of rehydrated pasta. Both the breaking

stress of rehydrated pasta and the difference in the moisture content between its inner and

peripheral regions were greater at higher rehydration temperatures, whereas the

rehydration temperature did not affect the breaking strain. The sodium chloride

concentration of the rehydrated solution did not affect either the moisture profile or

breaking stress, while the breaking strain was decreased by rehydration at higher sodium

chloride concentrations. These results suggest that the moisture profile within the pasta

and its material properties are governed by its breaking stress and strain, respectively.

65
The effects of drying conditions on pasta properties (such as its color, surface

structure, rupture strength, rehydration characteristics, texture, and sauce retention

capacity) were discussed. Notably, the sauce retention capacity of cooked spaghetti was

evaluated using a dextran solution as a simulated sauce. The effects of temperature and

humidity were independently examined under constant drying conditions, which were

compared to those applied industrially where the temperature and relative humidity are

changed stepwise with time. The results obtained in this study would be useful for

accurately determining the drying conditions for producing pasta with desirable

properties.

NOMENCLATURES

a constant of Eq. (1), (11), and (26)

A (xA, yA, zA)vertex coordinate

A0 frequency factor

AVE average

aw water activity

b constant of Eq. (1), (11), and (26)

B (xB, yB, zB)vertex coordinate

c constant of Eq. (1) and (11)

C (xC, yC, zC)vertex coordinate

CNT contrast

d initial diameter

Da apparent diffusion coefficient

66
De effective diffusion coefficient

H change in the enthalpy of rehydration

E activation energy

f(-) kinetic function concerning driving force for drying

g(-) roughness curve

G0 original gray-level

Gc level converted by contrast correction

Gr level converted by gamma correction

H relative humidity

I gray or height level

J0(-) Bessel function of the first kind of order zero

k mass transfer coefficient

K1 parameter in Eq. (3)

K2 parameter in Eq. (3)

kc mass transfer coefficient for a cylinder

kl rate constant (for mass loss)

kr rate constant (for rehydration)

ks mass transfer coefficient for slab

L thickness

l reference length

Ls length

Me equilibrium amount of spaghetti mass loss

Mt spaghetti mass loss

67
P(I) normalized frequency at I

q isosteric heat

R gas constant

r radius

R2 coefficient of determination

Ra average roughness

Rc drying rate during constant drying-rate period

Rw drying rate

S surface area

SKEW skewness

T absolute temperature

t time

T parameter in Eq. (3)

V partial molar volume

v0 initial rate of rehydration

VAR variance

w dry weight

w1 sample weight before rehydration

w2 sample weight after rehydration

X moisture content

x amount of water rehydrated per unit surface area

X0 initial moisture content ( = (w2 w1)/w)

xA x-coodinate of A

68
xB x-coodinate of B

xC x-coodinate of C

Xe equilibrium moisture content

Y fraction of water remaining on the flour

yA y-coodinate of A

yB y-coodinate of B

yC y-coodinate of C

zA z-coodinate of A

zB z-coodinate of B

zC z-coodinate of C

n nth positive root of J0(n) = 0

ACKNOWLEDGMENTS

This study was carried out as part of a project conducted by the Cereal Science

Consortium, the Graduate School of Agriculture, Kyoto University, Gifu University, and

the Nisshin Seifun Group, Inc. This work was partially supported by a Grant-in-Aid of

the Japan Society for the Promotion of Science Fellows (T.O.; Grant Number 14J02443).

REFERENCES

1. Wrigley, C.; Corke, H.; Walker, C.E. Encyclopedia of Grain Science; Elsevier: Oxford,

2004.

69
2. Feillet, P.; Dexter, J.E. Quality requirements of durum wheat for semolina milling and

pasta production. In Pasta and Noodle Technology ed. by Kruger, J.E.; Matsuo, R.R.;

Dick, J.W. American Association of Cereal Chemists: St. Paul, MN, 1996.

3. Donnelly, B.J. Teflon and non-Teflon lined dies: Effect on spaghetti quality. Journal of

Food Science 1982, 47, 10551058, 1069.

4. Dalbon, G.; Grivon, D.; Pagani, M.A. Continuous manufacturing process. In Pasta

and Noodles Technology ed. by Kruger, J.E.; Matsu, R.B.; Dick, J.W. American

Association of Cereal Chemists: St. Paul, MN, 1996.

5. Dawa, P.R. Pasta shape design. In Pasta and Semolina Technology ed. by Kill, R.C.;

Turnbull, K. Blackwell Science Ltd.: Oxford, 2001.

6. Lucisano, M.; Pagani, M.A.; Mariotti, M.; Locatelli, D.P. Influence of die material on

pasta characteristics. Food Research International 2008, 41, 646652.

7. Mercier, S.; Des Marchais, L.P.; Villeneuve, S.; Foisy, M. Effect of die material on

engineering properties of dried pasta. Procedia Food Science 2011, 1, 557562.

8. Pronyk, C.; Cenkowski, S.; Muir, W.E.; Lukow, O.M. Effects of superheated steam

processing on the textural and physical properties of Asian noodles. Drying Technology

2008, 26, 192203.

9. Pronyk, C.; Cenkowski, S.; Muir, W.E.; Lukow, O.M. Optimum processing conditions

of instant Asian noodles in superheated steam. Drying Technology 2008, 26, 204210.

10. Pronyk, C.; Cenkowski, S.; Muir, W.E. Drying kinetics of instant Asian noodles

processed in superheated steam. Drying Technology 2010, 28, 304314.

11. Lagoudaki, M.; Demertzis, P.G.; Kontominas, M.G. Moisture adsorption behaviour of

pasta products. LWT - Food Science and Technology 1993, 26, 512516.

70
12. Acker, L.W. Water activity, enzyme activity. Food Technology 1969, 23, 2740.

13. Labuza, T.P. Application of chemical kinetics to deterioration of foods. Journal of

Chemical Education 1984, 61, 348358.

14. Al-Muhtaseb, A.H.; McMinn, W.A.M.; Magee, T.R.A. Moisture sorption isotherm

characteristics of food products: A review. Food and Bioproducts Processing 2002, 80,

118128.

15. Van den Berg, C.; Bruin, S. Water activity and its estimation in food systems.

Rockland, L.B.; Stewart, F. (Eds.), Water activity: Influence on food quality; Academic

Press: New York, 1981.

16. Ogawa, T.; Kobayashi, T.; Adachi, S. Prediction of pasta drying process based on a

thermogravimetric analysis. Journal of Food Engineering 2012, 111, 129134.

17. Aversa, M.; Curcio, S.; Calabr, V.; Iorio, G. An analysis of the transport phenomena

occurring during food drying process. Journal of Food Engineering 2007, 78, 922932.

18. Curcio, S.; Aversa, M.; Calabr, V.; Iorio, G. Simulation of food drying: FEM

analysis and experimental validation. Journal of Food Engineering 2008, 87, 541553.

19. Ruiz-Lpez, I.I.; Garca-Alvarado, M.A. Analytical solution for food-drying kinetics

considering shrinkage and variable diffusivity. Journal of Food Engineering 2007, 79,

208216.

20. Hemis, M.; Singh, C.B.; Jayas, D.S.; Bettahar A. Simulation of coupled heat and mass

transfer in granular porous media: Application to the drying of wheat. Drying Technology

2011, 29, 12671272.

21. Ertekin, C.; Yaldiz, O. Drying of eggplant and selection of a suitable thin layer drying

model. Journal of Food Engineering 2004, 63, 349359.

71
22. Westerman, P.W.; White, G.M.; Ross, I.J. Relative humidity effect on the high

temperature drying of shelled corn. Transactions of the American Society of Agricultural

Engineers 1973, 16, 11361139.

23. Agrawal, Y.C.; Singh, R.D. Thin layer drying studies for short grain rice. ASAE Paper

No: 3531, American Society of Agricultural Engineers: St. Joseph, MI, 1977.

24. brahim, D.; Fergun, K. Drying and rehydration behaviors of convection drying of

green peas. Drying Technology 2011, 29, 12731282.

25. Xanthopoulos, G.; Lambrinos, G.; Manolopoulou, H. Evaluation of thin-layer models

for mushroom (Agaricus bisporus) drying. Drying Technology 2007, 25, 14711481.

26. Singh, S.; Raina, C.S.; Bawa, A.S.; Saxena, D.C. Effect of pretreatments on drying

and rehydration kinetics and color of sweet potato slices. Drying Technology 2006, 24,

14871494.

27. Overhults, D.G.; White, H.E.; Hamilton, H.E.; Ross, I.J. Drying soybeans with heated

air. Transactions of the American Society of Agricultural Engineers 1973, 16, 112113.

28. zdemir, M.; Devres, Y.O. The thin layer drying characteristics of hazelnuts during

roasting. Journal of Food Engineering 1999, 42, 225233.

29. Yagcioglu, A.; Degirmencioglu, A.; Cagatay, F. Drying characteristics of laurel leaves

under different drying conditions. In Proceedings of the 7th International Congress on

Agriculture and Mechanical Energy, Adana, Turkey, 1999.

30. Henderson, S.M. Progress in developing the thin-layer drying equation. Transactions

of the American Society of Agricultural Engineers 1974, 17, 11671168, 1172.

31. Sharaf-Eldeen, O.; Blaisdell, Y.I.; Spagna, G. A model for ear corn drying.

Transactions of the American Society of Agricultural Engineers 1980, 23, 12611271.

72
32. Wang, C.Y.; Singh, R.P. A single layer drying equation for rough rice. ASAE Paper

No: 78-3001, American Society of Agricultural Engineers: St. Joseph, MI, 1978.

33. Thompson, T.L.; Peart, R.M.; Foster, G.H. Mathematical simulation of corn drying a

new model. Transactions of the American Society of Agricultural Engineers 1968, 11,

582586.

34. Kassem, A.S. Comparative studies on thin layer drying models for wheat. In

Proceedings of the 13th International Congress of Agricultural Engineering, Morocco,

1998.

35. Verma, L.R.; Bucklin, R.A.; Endan, J.B.; Wratten, F.T. Effects of drying air

parameters on rice drying models. Transactions of the American Society of Agricultural

Engineers 1985, 28, 296301.

36. Karathanos, V.T. Determination of water content of dried fruits by drying kinetics.

Journal of Food Engineering 1999, 39, 337344.

37. Midilli, A.; Kucuk, H. Mathematical modeling of thin layer drying of pistachio by

using solar energy. Energy Conversion and Management 2003, 44, 11111122.

38. Inazu, T.; Iwasaki, K.; Furuta, T. Stress and crack prediction during drying of

Japanese noodle (udon). International Journal of Food Science and Technology 2005, 40,

621630.

39. Cubadda, R.E.; Carcea, M.; Marconi, E.; Trivisonno, M.C. Influence of gluten

proteins and drying temperature on the cooking quality of durum wheat pasta. Cereal

Chemistry 2007, 84, 4855.

73
40. D'Egidio, M.G.; Mariani, B.M.; Nardi, S.; Novaro, P.; Cubadda, R. Chemical and

technological variables and their relationships: A predictive equation for pasta cooking

quality. Cereal Chemistry 1990, 67, 275281.

41. Malcolmson, L.J.; Matsuo, R.R.; Balshaw. R. Textural optimization of spaghetti using

response surface methodology: Effects of drying temperature and durum protein level.

Cereal Chemistry 1993, 70, 417423.

42. Matsuo, R.R.; Dexter, J.E.; Kosmolak, F.G.; Leisle, D. Statistical evaluation of tests

for assessing spaghetti-making quality of durum wheat. Cereal Chemistry 1982, 59,

222228.

43. Novaro, P.; D'Egidio, M.G.; Mariani, B.M.; Nardi, S. Combined effect of protein

content and high-temperature drying systems on pasta cooking quality. Cereal Chemistry

1993, 70, 716719.

44. Zweifel, C.; Handschin, S.; Escher, F.; Conde-Petit, B. Influence of high-temperature

drying on structural and textural properties of durum wheat pasta. Cereal Chemistry 2003,

80, 159167.

45. Aktan, B.; Khan, K. Effect of high-temperature drying of pasta on quality parameters

and on solubility, gel electrophoresis, and reversed-phase high-performance liquid

chromatography of protein components. Cereal Chemistry 1992, 69, 288295.

46. Schofield, J.D.; Bottomley, R.C.; Timms, M.F.; Booth, M.R. The effect of heat on

wheat gluten and the involvement of sulphhydryl-disulphide interchange reactions.

Journal of Cereal Science 1983, 1, 241253.

74
47. Weegels, P.L.; Hamer, R.J. Temperature-induced changes of wheat products. In

Interactions: The Keys to Cereal Quality ed. by Hamer, R.J.; Hoseney, R.C. American

Association of Cereal Chemists: St. Paul, MN, pp.95130, 1998.

48. Lintas, C.; D'Appolonia, B.D. Effects of spaghetti processing on semolina

carbohydrates. Cereal Chemistry 1973, 50, 563570.

49. Cubadda, R. Current research and future needs in durum wheat chemistry and

technology. Cereal Foods World 1989, 34, 206209.

50. Resmini, P.; Pagani, M. Ultrastructure studies of pasta. A review [wheat flour, rice

flour, heat starch modification, protein coagulation]. Food Microstructure 1983, 2, 112.

51. Grant, L.A.; Dick, J.W.; Shelton, D.R. Effects of drying temperature, starch damage,

sprouting and additives on spaghetti quality characteristics. Cereal Chemistry 1993, 70,

676684.

52. Sharma, R.; Sissons, M.J.; Rathien, A.J.; Jenner C.F. The null-4A allele at the waxy

locus in durum wheat affects pasta cooking quality. Journal of Cereal Science 2002, 35,

287297.

53. Laignelet, B. Variazione del coloure e di altri fattori qualitativi delle paste durante

lessicamento. Tecnica Molitoria 1977, 28, 123.

54. Abecassis, J.; Chaurand, M.; Matencio, F.; Feillet, P. Einfluss des Wassergehaltes der

Teigwaren bei der Hochtemperaturtrocknung. Getreide, Mehl und Brot 1989, 43, 5862.

55. Acquistucci, R. Influence of Maillard reaction on protein modification and colour

development in pasta. Comparison of different drying conditions. LWT - Food Science

and Technology 2000, 33, 4852.

75
56. DeStefanis, E.; Sgrulletta, E. Effects of high temperature drying on technological

properties of pasta. Journal of Cereal Science 1990, 12, 97104.

57. Dexter, J.E.; Matsuo, R.R.; Morgan, B.C. High temperature drying: Effects on

spaghetti properties. Journal of Food Science 1981, 46, 17411746.

58. Manser, J. High temperature drying of pasta products. Buhler Diagram 1980, 69,

1112.

59. Zheng, X.; Liu, C.; Chen, Z.; Ding, N.; Jin, C. Effect of drying conditions on the

texture and taste characteristics of rough rice. Drying Technology 2011, 29, 12971305.

60. Cummings, D.A.; Litchfield, J.B.; Okos, M.R. Stress and failure prediction for the

drying, tempering, and cooling of extruded durum semolina. Drying Technology 1993, 11,

18091836.

61. Saguy, I.S.; Marabi, A.; Wallach, R. New approach to model rehydration of dry food

particulates utilizing principles of liquid transport in porous media. Trends in Food

Science and Technology 2005, 16, 495506.

62. Garca-Pascual, P.; Sanjun, N.; Melis, R.; Mulet, A. Morchella esculenta (morel)

rehydration process modelling. Journal of Food Engineering 2006, 72, 346353.

63. Djomdi, E.R.; Ndjouenkeu, R. Soaking behaviour and milky extraction performance

of tiger nut (Cyperus esculentus) tubers. Journal of Food Engineering 2007, 78, 546550.

64. Maskan, M. Effect of processing on hydration kinetics of three wheat products of the

same variety. Journal of Food Engineering 2002, 52, 337341.

65. Misra, M.K.; Brooker, D.B. Thin-layer drying and rewetting equations for shelled

yellow corn. Transactions of the American Society of Agricultural and Biological

Engineers 1980, 23, 12541260.

76
66. Peleg, M. An empirical model for the description of moisture sorption curves. Journal

of Food Science 1988, 53, 12161219.

67. Chhinnan, M.S. Evaluation of selected mathematical models for describing thin-layer

drying of in-shell pecans. Transactions of the American Society of Agricultural and

Biological Engineers 1984, 27, 610615.

68. Becker, H.A. On the absorption of liquid water by the wheat kernel. Cereal Chemistry

1960, 37, 309323.

69. Cunha, L.M.; Oliveira, F.A.R.; Oliveira, J.C. Optimal experimental design for

estimating the kinetic parameters of processes described by the Weibull probability

distribution function. Journal of Food Engineering 1998, 37, 175191.

70. Marabi, A.; Livings, S.; Jacobson, M.; Saguy, I.S. Normalized Weibull distribution for

modeling rehydration of food particulates. European Food Research and Technology

2003, 217, 311318.

71. Cunningham, S.E.; McMinn, W.A.M.; Magee, T.R.A.; Richardson, P.S. Modelling

water absorption of pasta during soaking. Journal of Food Engineering 2007, 82,

600607.

72. Bilbao-Sinz, C.; Andrs, A.; Fito, P. Hydration kinetics of dried apple as affected by

drying conditions. Journal of Food Engineering 2005, 68, 369376.

73. Garca-Pascual, P.; Sanjun, N.; Bon, J.; Carreres, J.E.; Mulet, A. Rehydration process

of Boletus edulis mushroom: characteristics and modelling. Journal of the Science of

Food and Agriculture 2005, 85, 13971404.

74. Sanjun, N.; Bon, J.; Clemente, G.; Mulet, A. Changes in the quality of dehydrated

broccoli florets during storage. Journal of Food Engineering 2004, 62, 1521.

77
75. Sanjun, N.; Simal, S.; Bon, J.; Mulet, A. Modelling of broccoli stems rehydration

process. Journal of Food Engineering 1999, 42, 2731.

76. Del Nobile, M.A.; Buonocore, G.G.; Panizza, A.; Gambacorta, G. Modeling the

spaghetti hydration kinetics during cooking and overcooking. Journal of Food Science

2003, 68, 13161323.

77. Lee, K.T.; Farid, M.; Nguang, S.K. The mathematical modelling of the rehydration

characteristics of fruits. Journal of Food Engineering 2006, 72, 1623.

78. De Temmerman, J.; Verboven, P.; Nicola, B.; Ramon, H. Modelling of transient

moisture concentration of semolina pasta during air drying. Journal of Food Engineering

2007, 80, 892903.

79. Horigane, A.K.; Takahashi, H.; Maruyama, S.; Ohtsubo, K.; Yoshida, M. Water

penetration into rice grains during soaking observed by gradient echo magnetic resonance

imaging. Journal of Cereal Science 2006, 44, 307316.

80. Hills, B.P.; Babonneau, F.; Quantin, V.M.; Gaudet, F.; Belton, P.S. Radial NMR

microimaging studies of the rehydration of extruded pasta. Journal of Food Engineering

1996, 27, 7186.

81. Hills, B.P.; Godward, J.; Wright, K.M. Fast radial NMR microimaging studies of

pasta drying. Journal of Food Engineering 1997, 33, 321335.

82. Irie, K.; Horigane, A.K.; Naito, S.; Motoi, H.; Yoshida, M. Moisture distribution and

texture of various types of cooked spaghetti. Cereal Chemistry 2004, 81, 350355.

83. Sekiyama, Y.; Horigane, A.K.; Ono, H.; Irie, K.; Maeda, T.; Yoshida, M. T2

distribution of boiled dry spaghetti measured by MRI and its internal structure observed

by fluorescence microscopy. Food Research International 2012, 48, 374379.

78
84. Altan, A.; Oztop, M.H.; McCarthy, K.L.; McCarthy, M.J. Monitoring changes in feta

cheese during brining by magnetic resonance imaging and NMR relaxometry. Journal of

Food Engineering 2011, 107, 200207.

85. Ogawa, T.; Adachi, S Measurement of moisture profiles in pasta during rehydration

based on image processing. Food and Bioprocess Technology 2014, 7, 14651471.

86. Brunauer, S.; Deming, L.S.; Deming, W.E.; Troller, E. On a theory of the van der

Waals adsorption of gases. Journal of the American Chemical Society 1940, 62,

17231732.

87. Hebrard, A.; Oulahna, D.; Galet, L.; Cuq, B.; Abecassis, J.; Fages, J. Hydration

properties of durum wheat semolina: Influence of particle size and temperature. Powder

Technology 2003, 130, 211218.

88. Heldman, D.R.; Lund, D.B. Handbook of Food Engineering, Marcel Dekker: New

York, 1992.

89. Becker, H.A.; Sallans, H.R. A study of desorption isotherms of wheat at 25C and

50C. Cereal Chemistry 1956, 33, 7991.

90. Heldman, D.R.; Hall, C.W.; Hedrick, T.I. Vapor equilibrium relationships of dry milk.

Journal of Dairy Science 1965, 48, 845852.

91. Chuma, A.; Ogawa, T.; Kobayashi, T.; Adachi, S. Moisture sorption isotherm of

durum wheat flour. Food Science and Technology Research 2012, 18, 617622.

92. Tsamai, E.; Maroulis, Z.B.; Morunos-Kouris, D.; Saravacos, G.D. Heat of sorption of

water in dried fruits. International Journal of Food Science and Technology 1990, 25,

350359.

79
93. Aguerre, R.J.; Surez, C.; Viollaz, P.E. Enthalpy-entropy compensation in sorption

phenomena: Application to the prediction of the effect of temperature on food isotherms.

Journal of Food Science 1986, 51, 15471549.

94. Falabella, M.C.; Aguerre, R.J.; Suarez, C. Modeling non-isothermal sorption

equilibrium data of cereal grains. LWT - Food Science and Technology 1992, 25,

286288.

95. Roman-Gutierrez, A.D.; Guilbert, S.; Cuq, B. Distribution of water between wheat

flour components: A dynamic water vapour adsorption study. Journal of Cereal Science

2002, 36, 347355.

96. Manabe, M.; Adachi, S.; Watanabe, T.; Kawamura, H. Dilatometric measurement of

the partial molar volume of water hydrated to lipase from Rhizopus delemer. Agricultural

and Biological Chemistry 1989, 53, 571572.

97. Hasegawa, A.; Ogawa, T.; Adachi, S. Dilatometric measurement of the partial molar

volume of water sorbed to durum wheat flour. Bioscience, Biotechnology, and

Biochemistry 2013, 77, 15651568.

98. Fuwa, H.; Komaki, T.; Hidukuri, S.; Kainuma, K. Handbook of Starch Science (in

Japanese; Denpun Kagaku no Jiten), 1st edn. Asakura Shoten: Tokyo, 2003.

99. Ogura, T. Denpun Kagaku Handbook, ed. by Kikuni, J.; Nakamura, M.; Suzuki, S.

Asakura Shoten: Tokyo, 1977.

100. Ozawa, M.; Seguchi, M. Relationship between pancake springiness and interaction

of wheat flour components. Food Science and Technology Research 2006, 12, 167172.

80
101. Cuq, B.; Verniere, C. Characterisation of glass transition of durum wheat semolina

using modulated differential scanning calorimetry. Journal of Cereal Science 2001, 33,

213221.

102. Crank, J. The Mathematics of Diffusion. 1st edn., Clarendon Press: Oxford, 1975.

103. Andrieu, J.; Jallut, C.; Stamatopoulos, A.; Zafiropoulos, M. Identification of water

apparent diffusivities for drying of corn based extruded pasta. In Proceedings of 6th

International Drying Symposium, 1988.

104. Waananen, K.M.; Okos, M.R. Effect of porosity on moisture diffusion during drying

of pasta. Journal of Food Engineering 1996, 28, 121137.

105. Xing, H.; Takhar, P.S.; Helms, G.; He, B. NMR imaging of continuous and

intermittent drying of pasta. Journal of Food Engineering 2007, 78, 6168.

106. Moraga, G.; Martinez-Navarrete, N.; Chiralt, A. Water sorption isotherms and glass

transition in strawberries: Influence of pretreatment. Journal of Food Engineering 2004,

62, 315321.

107. Goula, A.M.; Karapantsios, T.D.; Achilias, D.S.; Adamopoulos, K.G. Water sorption

isotherms and glass transition temperature of spray dried tomato pulp. Journal of Food

Engineering 2008, 85, 7383.

108. Mrad, N.D.; Bonazzi, C.; Boudhrioua, N.; Kechaou, N.; Courtois, F. Influence of

sugar composition on water sorption isotherms and on glass transition in apricots. Journal

of Food Engineering 2012, 111, 403411.

109. Mizuno, A.; Mitsuiki, M.; Motoki, M. Effect of crystallinity on the glass transition

temperature of starch. Journal of Agricultural and Food Chemistry 1998, 46, 98103.

81
110. Liu, P.; Yu, L.; Liu, H.; Chen, L.; Li, L. Glass transition temperature of starch

studied by a high-speed DSC. Carbohydrate Polymers 2009, 77, 250253.

111. Akiyama, T.; Hayakawa, K. Heat and moisture transfer and hygrophysical changes

in elastoplastic hollow cylinder-food during drying. Journal of Food Science 2000, 65,

315323.

112. Mihoubia, D.; Bellagi, A. Two-dimensional heat and mass transfer during drying of

deformable media. Applied Mathematical Modelling 2008, 32, 303314.

113. Pavn-Melendez, G.; Hernndez, J.A.; Salgado, M.A.; Garcia, M.A. Dimensionless

analysis of the simultaneous heat and mass transfer in food drying. Journal of Food

Engineering 2002, 51, 347353.

114. Tripathy, P.P.; Kumar, S. A methodology for determination of temperature dependent

mass transfer coefficients from drying kinetics: Application to solar drying. Journal of

Food Engineering 2009, 90, 212218.

115. Singh, P.P.; Maier, D.E.; Cushman, J.H.; Campanella, O.H. Effect of viscoelastic

relaxation on moisture transport in foods. Part II: Sorption and drying of soybeans.

Journal of Mathematical Biology 2004, 49, 2034.

116. Ogawa, T.; Koizumi, S.; Adachi, S. Thermal analysis of drying process of durum

wheat dough under the programmed temperature-rising conditions. Food and Bioproducts

Processing 2014, 92, 913.

117. Ozawa, T. Estimation of activation energy by isoconversion methods.

Thermochimica Acta 1992, 203, 159165.

82
118. Friedman, H.L. Kinetics of thermal degradation of char-forming plastics from

thermogravimetry. Application to a phenolic plastic. Journal of Polymer Science Part C

1964, 6, 183195.

119. Kawai, K.; Kim, Y.J.; Suzuki, T. Estimation of the stability of glassy product by

thermal analysis. Japan Journal of Thermophysical Properties 2003, 17, 114127.

120. Ponsart, G.; Vasseur, J.; Frias, J.M.; Duquenoy, A.; Mot, J.M. Modelling of stress

due to shrinkage during drying of spaghetti. Journal of Food Engineering 2003, 57,

277285.

121. Cuq, B.; Gonalves, F.; Mas, J.F.; Vareille, L.; Abecassis, J. Effects of moisture

content and temperature of spaghetti on their mechanical properties. Journal of Food

Science 2003, 59, 5160.

122. Liu, H.; Qi, J.; Hayakawa, K. Rheological properties including tensile fracture stress

of semolina extrudates influenced by moisture content. Journal of Food Science 2006, 62,

813820.

123. Ponsart, G.; Vasseur, J.; Mot, J.M. Stress building in pasta during drying, and

prediction of cracks apparition. In Proceedings of the 14th International Drying

Symposium, A687693, 2004.

124. Guinea, G.V.; Rojo, F.J.; Elice, M. Brittle failure of dry spaghetti. Engineering

Failure Analysis 2004, 11, 705714.

125. Mizuno, N.; Ogawa, T.; Adachi S. Shrinkage and tensile stress of sheet-like and

cylindrical pastas with various moisture contents. Food Bioscience 2013, 2, 1014.

126. Ogawa, T.; Adachi, S. A simple method to measure the surface roughness of

spaghetti using a digital camera. Food Science and Technology Research, in press.

83
127. Chillo, S.; Iannetti, M.; Civica, V.; Suriano, N.; Mastromatteo, M.; Del Nobile, M.A.

A study of the relationship between the mechanical properties and the sensorial optimal

cooking time of spaghetti. Journal of Food Engineering 2009, 94, 222226.

128. Chillo, S.; Laverse, J.; Falcone, P.M.; Protopapa, A.; Del Nobile, M.A. Influence of

the addition of buckwheat flour and durum wheat bran on spaghetti quality. Journal of

Cereal Science 2008, 47, 144152.

129. Ogawa, T.; Kobayashi, T.; Adachi, S. Water sorption kinetics of spaghetti at different

temperatures. Food and Bioproducts Processing 2011, 89, 135141.

130. Saol, S.; Turhan, M.; Sayar, S. A potential method for determining in situ

gelatinization temperature of starch using initial water transfer rate in whole cereals.

Journal of Food Engineering 2006, 76, 427432.

131. Watanabe, H. The factor which governs water migration in starchy foods. Japan

Journal of Food Engineering 2004, 5, 143151.

132. Hasegawa, A.; Ogawa, T.; Adachi, S. Estimation of the gelatinization temperature of

noodles from water sorption curves under temperature-programmed heating conditions.

Bioscience, Biotechnology, and Biochemistry 2012, 76, 21562158.

133. Gler, S.; Kksel, H.; Ng, P.K.W. Effects of industrial pasta drying temperatures on

starch properties and pasta quality. Food Research International 2002, 35, 421427.

134. Zaidul, I.S.M.; Yamauchi, H.; Matsuura-Endo, C.; Takigawa, S.; Noda, T. Thermal

analysis of mixtures of wheat flour and potato starches. Food Hydrocolloids 2008, 22,

499504.

135. Ogawa, T.; Adachi, S. Effect of salts on the water sorption kinetics of dried pasta.

Bioscience, Biotechnology, and Biochemistry 2013, 77, 249252.

84
136. Chiotelli, E.; Pilosio, G.; Meste, M.L. Effect of sodium chloride on the

gelatinization of starch: A multi measurements study. Biopolymer 2002, 63, 4158.

137. Uedaira, H. What Is Water? (in Japanese) 1st edn., Kodansha: Tokyo, 1977.

138. Yoshino, M.; Ogawa, T.; Adachi, S. Properties and water sorption characteristics of

spaghetti prepared using various dies. Journal of Food Science 2013, 78, E520525.

139. Ogawa, T.; Adachi, S. Effect of surface roughness on rehydration kinetics of

spaghetti. Japan Journal of Food Engineering 2014, 15, 101104.

140. Aimoto, U.; Ogawa, T.; Adachi, S. Water sorption kinetics of spaghetti prepared

under different drying conditions. Food Science and Technology Research 2013, 19,

1722.

141. Cunina, C.; Handschina, S.; Waltherb, P.; Eschera, F. Structural changes of starch

during cooking of durum wheat pasta. LWT - Food Science and Technology 1995, 28,

323328.

142. Dexter, J.E.; Dronzek, B.L.; Matsuo, R.R. Scanning electron microscopy of cooked

spaghetti. Cereal Chemistry 1978, 55, 2330.

143. Toi, K.; Odani, H.; Nakagawa, T. High-molecular-weight molecule and water (in

Japanese; Koubunsi to Mizu), 1st edn. Kyoritsu Pab.: Tokyo, 1995.

144. Crank, J.; Park, G.S. Diffusion in high polymers: some anomalies and their

significance. Transactions of the Faraday Society 1954, 47, 10721084.

145. Ogawa, T.; Hasegawa, A.; Adachi, S. Effects of relaxation of gluten network on

rehydration kinetics of pasta. Bioscience, Biotechnology, and Biochemistry 2014, 78,

19301934.

85
146. Long, R.A.; Richman, D. Concentration gradients for diffusion of vapors in glassy

polymers and their relation to time dependent diffusion phenomena. Journal of the

American Chemical Society 1960, 82, 513519.

147. Ogawa, T.; Adachi, S. Moisture distribution and texture of spaghetti rehydrated

under different conditions. Bioscience, Biotechnology, and Biochemistry 2016, 80,

769773.

148. Tuhumury, H.C.D.; Small, D.M.; Day, L. The effect of sodium chloride on gluten

network formation and rheology. Journal of Cereal Science 2014, 60, 229237.

149. Ukai, T.; Matsumura, Y.; Urade, R. Disaggregation and reaggregation of gluten

proteins by sodium chloride. Journal of Agricultural and Food Chemistry 2008, 56,

11221130.

150. Larsson, H. Effect of pH and sodium chloride on wheat flour dough properties:

Ultracentrifugation and rheological measurements. Cereal Chemistry 2002, 79, 544545.

151. Ogawa, T.; Chuma, A.; Aimoto, U.; Adachi, S. Effects of drying temperature and

relative humidity on spaghetti characteristics. Drying Technology, in press.

152. Ogawa, T.; Chuma, A.; Aimoto, U.; Adachi, S. Characterization of spaghetti

prepared under different drying conditions. Journal of Food Science 2015, 80,

E19591964.

153. Ogawa, T.; Adachi, S. Effects of drying conditions on moisture distribution in

rehydrated spaghetti. Bioscience, Biotechnology, and Biochemistry 2014, 78, 14121414.

154. Feillet, P.; Jeanjean, M.F.; Kobrehel, K.; Laignelet, B. Le brunissement des ptes

alimentaires. Bulletin de lEcole Nationale Suprieure de Meunerie et des Industries

Cralires 1974, 262, 190194.

86
155. Abecassis, J.; Alary, R.; Feillet, P. Influence des tempratures de schage sur laspect

et la qualit culinaire des ptes alimentaires. Industries des Crales 1984, 31, 1318.

156. Pronyk, C.; Cenkowski, S.; Muir, W.E.; Lukow, O.M. Effects of superheated steam

processing on the textural and physical properties of asian noodles. Drying Technology

2008, 26, 192203.

157. Pronyk, C.; Cenkowski, S.; Muir, W.E.; Lukow, O.M. Optimum processing

conditions of instant asian noodles in superheated steam. Drying Technology 2008, 26,

204210.

158. Pronyk, C.; Cenkowski, S.; Muir, W.E. Drying kinetics of instant asian noodles

processed in superheated steam. Drying Technology 2010, 28, 304314.

159. Gulia, N.; Dhaka, V.; Khatkar, B.S. Instant noodles: processing, quality, and

nutritional aspects. Drying Technology 2014, 54, 13861399.

160. Alfy, A.; Kiran, B.V.; Jeevitha, G.C.; Umesh, H.H. Recent developments in

superheated steam processing of foodsa review. Drying Technology 2016, 56,

21912208

161. Duan, X.; Zhang, M.; Mujumdar, A.S.; Wang, R. Trends in microwave-assisted

freeze drying of foods. Drying Technology 2010, 28, 444453.

162. Zhou, B.; Zhang, M.; Fang, Z.; Liu. Y. A combination of freeze drying and

microwave vacuum drying of duck egg white protein powders. Drying Technology 2014,

32, 18401847.

87
Table 1. Conditions for drying under high-temperature (HT) and low-temperature (LT)

conditions.

High-temperature (HT) Low-temperature (LT)

Step 1 2 3 1 2 3

Time [h] 0.5 3.5 1 1 5 1

Temperature [C] 50 80 30 40 60 30

Humidity [% RH] 60 75 60 60 75 60

88
Table 2. AFM images and sauce retention capacities of programmed-dried spaghetti.

Drying LT HT VHT

condition

AFM

image

Sauce 9.58 0.93 mg/cm2 11.93 1.58 mg/cm2 11.21 0.55 mg/cm2

retention

capacity

89
FIG. 1. Desorption isotherms of water on durum wheat flour at () 30C, () 40C, ()
50C, () 60C, () 70C, and () 80C. The curves are calculated to best fit the
observed moisture contents to the Guggenheim-Anderson-de Boer (GAB) equation.

90
FIG. 2. Partial molar volume at 25C of water sorbed onto () untreated, () dry-heated,

and () pregelatinized durum wheat flour samples with various moisture contents. Inset:

relationship between the () water sorption isotherm and ( ) partial molar volume of

water at 25C for untreated durum wheat flour.

91
FIG. 3. Drying characteristic curve obtained by thermogravimetry at 90C and 20.6% RH.

Inset: apparatus used to press hydrated semolina into a single-sided open cell.

92
FIG. 4. Relationship between moisture content and endotherm during drying, as
determined by differential scanning calorimetry (DSC) and thermogravimetry (TG) at (1)
0.2, (2) 0.4, (3) 0.6, (4) 0.8, and (5) 1.0 C/min. Arrows indicate the locations of
endothermic peaks.

93
FIG. 5. Relationships between moisture content and () temperature of the inflection

points of drying characteristic curves, () conclusion temperature of the endothermic

peaks in DSC, and () glass transition temperature. The glass transition temperatures

were adopted from the literature.[106]

94
FIG. 6. Comparison of () experimental drying curve with those calculated () with

and ( ) without considering the constant drying-rate period. Spaghetti (cylindrical

pasta) was dried under LT program conditions. Inset: expanded figure for the early stage

of drying.

95
FIG. 7. Youngs modulus of dumbbell specimens of durum semolina dough with different

moisture contents. Specimens were dried () at 50C and 40% relative humidity and ()

at 50C and 80% relative humidity. The symbol () indicates pasta prepared at 50C and

80% relative humidity and stored at 50C for 2 d to obtain a homogeneous moisture

distribution.

96
FIG. 8. Transfer images of spaghetti (a), examples of (b) gray-level frequency

distributions derived from green images, and (c) height-level frequency distributions

derived from laser scanning microscopy for spaghetti extruded through dies made of (1)

Teflon, (2) polypropylene, (3) polycarbonate, (4) aluminum, and (5) bronze. Images were

trimmed from their original size to enhance the visualization of surface roughness. The

transverse value of the gray-level frequency distribution (b) is inverted.

97
FIG. 9. Gray-level profiles of a cross-sectional image of spaghetti rehydrated for 10.2

min. Inset: cross-sectional images of spaghetti rehydrated for 10.2 min: (a) original image

and (b) digitally processed image. The broken lines show segments of the moisture

profiles as an example.

98
FIG. 10. Changes in moisture content at () 50, () 60, () 70, (, , ) 80, and ()

90C for spaghetti with initial diameters of () 1.4, (, , , , ) 1.6, and () 1.8

mm.

99
FIG. 11. Temperature dependence of (, ) a/b and (, ) a values. Unfilled and filled

symbols represent sorption in 1.83 mol/L NaCl and water, respectively.

100
FIG. 12. Water sorption curves for () LT spaghetti, () udon, () kishimen, ()

juwari-soba, () hachiwari-soba, () common soba, () Malony, and () kuzukiri at a

temperature-raising rate of 1.38 C/min.

101
FIG. 13. Surface images of spaghetti prepared with Teflon (a) and bronze dies (b) and
dependence of the momentarily rehydrated amounts of water per unit area for () Xt0
(smooth) and () Xt0 (rough) on the calculated average roughness Ra.

102
FIG. 14. Moisture profiles of spaghetti rehydrated for 0 (a), 1 (b), 10.2 (c), and 20 min

(d).

103
FIG. 15. Changes in the normalized moisture contents for ( ) 0 mm durum pasta

and ( ) gluten pasta. The curves represent the moisture contents calculated

using Eq. (11). Inset: estimation of the moisture content of infinitely thin durum pasta

with initial diameters of () 1.30, () 1.35, () 1.63, and () 1.71 mm by

extrapolation at the rehydration times of (a) 5, (b) 20, (c) 40, and (d) 80 min.

104
FIG. 16. Moisture distributions of spaghetti rehydrated at (a) 100, (b) 80, and (c) 60C in

distilled water. The distance from the center was normalized by the radius of the spaghetti.

105
FIG. 17. Representative examples of stress-strain curves for spaghetti rehydrated at (a)

100, (b) 80, and (c) 60C in distilled water.

106
FIG. 18. Relationships between moisture content and drying duration at () 50C and

50% RH, ( ) 50C and 80% RH, ( ) 85C and 50% RH, and ( ) 85C and 80%

RH. The curves were drawn by smoothly connecting the points representing the measured

values. Inset: AFM images of the surface of spaghetti dried at (a) 85C and 50% RH, (b)

50C and 50% RH, (c) 85C and 80% RH, and (d) 50C and 80% RH.

107

Vous aimerez peut-être aussi