Sunteți pe pagina 1din 153

A STUDY OF DRYING BEHAVIOUR OF POTATO SLICES

BY
JOSEPH KANGOGO KIPYAKWAI

UNIVERSITY OF NATROW
library
F, O. Bo* 301*7
HAIR**1

,co.***»^Yl

IB®C ' *£**.


L * * WBB‘

A thesis
lefeis submitted in partial fulfilment of the
requirements for the Degree of Master of Science in
Agricultural Engineering in the University of Nairobi.

Faculty of Engineering

1997
DECLARATION

I declare that this is my original work and has not been


presented for a degree in any other University.

Date- it.) gig's


Joseph Kangogo Kipyakwai

This thesis has been submitted with our approval as


University supervisors:

Date

♦ i
ABSTRACT
A STUDY OF DRYING BEHAVIOUR OF POTATO SLICES

By
Joseph Kangogo Kipyakwai

Potato (Solarium tuberosum) is an underexploited tuber with


promising economic value and the potential to help ease the
world food problem. It produces more calories per unit area
of land than other tropical root crops and tubers. For
practical and efficient utilization, it must be processed into

ration dense, stable forms. A study was made to process the

Desiree variety into dried slices which could be stored at

tropical ambient conditions for at least one year. The quality


of the air dried slices were evaluated. Results indicate this
approach to be feasible.

Forced-convection drying behaviour of potato slices was


experimentally determined as a function of the drying air

temperature, airflow rate and thickness of the slices under

laboratory conditions.
Drying of potato slices is an unsteady-state process under
the control of diffusion rate of water through the slice
surface. Therefore, temperature, thickness of the potato
slices and the drying air flow rate control the drying rate,
in tnis work, the initial thickness of potato slices was kept

constant at 2 mm, 4 mm and 6 mm and the effects of drying air

temperature and airflow rate were investigated. Through the


measurements of moisture content, colour and rehydration,
actual drying behaviour of potato slices under different dryer

conditions were determined. Colour and rehydraticn of the


dried slices were used as criteria for the evaluation and

determination of the optimum conditions.


Statistical analysis of experimental data showed the
drying air temperature had the greatest effect on drying rate

and quality, followed by slice thickness; air mass flow rate

had little effect on the thin-layer drying rate. Specifically,


drying rate was inversely proportional to the square of slice
thickness at constant drying air temperature. Three-term drying
equation models developed using the statistical method of
successive residuals were considered sufficient for practical

applications. A thin layer drying equation that can predict

drying behaviour of potato slices was also developed.

The quality of the air dried slices were evaluated and a


nonlinear model was developed for vitamin C degradation based
on the applicability of the Arrhenius equation.


IV
ACKNOWLEDGEMENT

I am greatly indebted to my two supervisors, Dr. E.N.

Mwaura and Prof. J.K. Imungi, for their guidance and help
throughout the course of this study. I would also like to
thank Prof. L.O. Gumbe and Dr. P.G. Kaumbutho for their useful
suggestions during progress report presentation seminars.

This research work would not have been successful were it


not for the permission given by the department of Food

Technology and Nutrition to use their research facilities. The

assistance given by Prof. S.K. Mbugua and the Technical staff


of that department is greatly appreciated.
Let me also register my sincere appreciation to Prof. D.K.
arap Some and the Netherlands Government for their respective
roles in the provision of the M.Sc. scholarship. I also

acknowledge, with thanks, assistance given by Mr. Frederick

Wanguhu, a senior technician of the Processing Laboratory in

the Agricultural Engineering department. There are other


people who have contributed generously towards the success of

this work, only space does not allow specific mention of their
names. Thank you.

♦v
TABLE OF CONTENTS.

PAGE
LIST OF TABLES viii
LIST OF FIGURES ix
ABBREVIATIONS AND SYMBOLS x
CHAPTER

1. INTRODUCTION

h
2. LITERATURE REVIEW

<j ^
2.1 The Potato as a World Crop
2.2 The origin of the Potato Plant and its
Development in Kenya

in to o
2.3 Potatoes in the Kenyan Food system
2.4 Processes to Extend Potato Storage Life 1
2.4.1 Potato Dehydration 10
2.4.2 Potato Drying Technology 15
2.4.3 Theory and Principles of Drying 19
2.4.4 Rate Periods of Drying 24
2.4.5 Drying Under Simulated Practical
Conditions 28
2.5 Changes in Potato Tissue during Drying 29
2.6 Theoretical Equation for Quality Loss 31
3. METHODOLOGY 34
3.1 Raw Material 34
3.2 Preparation of Potatoes for Processing 34
3.3 Drying of the Slices 36
3.4 Analytical Methods 38
3.4.1 Determination of Moisture Content 38
3.4.2 Determination of Ascorbic Acid 38
3.4.3 Measurement of Browning
Discolouration (heat damage) 39
3.4.4 Determination of Rehydration
Property 39
3.4.5 Development of Drying Equations 39
4. RESULTS AND DISCUSSION 42
4.1 Moisture Reduction 42
4.1.1 Effect of Drying Air Temperature
on Course of Drying of Potato Slice 48
4.1.2 Effect of Thickness on Course of
Drying of Potato Slice 50
4.1.3 Drying Rate as a Function of
Sample Thickness 51
4.1.4 Drying Equations for Potato Slices 55
4.1.5 Regression Analysis for the
Drying equations 63
4.2 Ascorbic Acid Retention 73
4.3 Visible Browning of the dried Product 85
4.4 Case-hardening Behaviour 89
4.5 Appropriate Drying Conditions * 94
*
* «Vi
4.6 Optimum Drying Conditions 96
4.6.1 Models 97
4.6.2 Maximizing Ascorbic Acid Retention 103
4.6.3 Minimizing Drying Time With a
Constraint on Ascorbic Acid Retention 105
CONCLUSIONS 107
SUGGESTIONS FOR FURTHER RESEARCH 111
BIBLIOGRAPHY 112
APPENDICES 120
Appendix A: Statistical Analysis of Moisture
Reduction Data 121
Appendix B: Psychrometric Representation
of Ambient Conditions for Air used
in the Drying experiments 124
Appendix C: Experimentally determined
Equilibrium Moisture Content values 125
Appendix D: Effect of Air Temperature and
Potato Slice Thickness on Course of
Drying 127
Appendix E: Arrhenius plots illustrating
influence of temperature on rate
constant, k 135
Appendix F: Raw Data Used for production
of figures shown in the main text 137
Appendix G: Maps for Potato Production
Areas in Kenya 140

*ii
LIST OF TABLES

2.1 World production of major starch crops


(cereals and tubers) for 1996 in metric
t o n s (000)) 4
2.2 The ten food crops with the highest
production value per hectare per day
in developing countries 5
4.1 The experimental data on time of half response
for all drying conditions studied 42
4.2 The effect of drying conditions on the
time of half-response (minutes) 43
4.3 Time (minutes) of half-response (MR = 0.5)
for the drying at three temperature levels A,
for three slice thicknesses B, and two air
flow rates, C (totals of three replications) 44
4.4 Analysis of variance (ANOVA) of data in
table 4.2 45
4.4a Examination of the BC interaction for the
data of Table 4.3 47
4.4b Examination of the AB interaction for the
data of Table 4.3 48
4.5 Experimental data for development of drying
equations 6

lO O'*
4.6 Regression Equations and parameters for Fig.4.10 6
4.7 Regression Equations and parameters for Fig.4.11
4.8 Experimentally determined equilibrium moisture
content, Me 72
4.9 Experimentally determined first-order rate
constants for reduced ascorbic acid SO
4.10 Experimentally determined Arrhenius equation
constants for reduced ascorbic acid
* degradation 84
4.11 Visible browning values for fully dried
slices 86
4.12 Rehydration ratio values for fully dried
slices 91
4.13 Residence time for case-hardening
occurrence S3
4.14 Appropriate constant drying conditions
for potato slices 95
4.15 Ascorbic acid kinetic model parameter values for
Equation (4.37) and (4.38) (Mishkin et al., 1984b) 98
4.16 Moisture transfer model, parameter values for
Equation (4.43) and (4.44) (Mishkin et al., 1984a) 101

v*iii
LIST OF FIGURES

3.1 Flow sheet for dehydration of raw


potato slices 35
4.1 Effect of air temperature on course
of drying of potato slice 49
4.2 Effect of potato slice thickness on
drying curves 50
4.3 Time from MR=1 to MR=0.1 as a function
of the square of sample thickness 52
4.4 Development of drying equation for
potato slices at 50 °C air temperature
and air flow rate of 0.0054 kg/m2s 57
4.5 Development of drying equation for
potato slices at 60°C air temperature
and air flow rate of 0.0054 kg/m2s 58
4.6 Development of drying equation for
potato slices at 70°C air temperature
and air flow rate of 0.0054 kg/m2s 60
4.7 Regression analysis output for Fig.4.4 63
4.8 Regression analysis output for Fig.4.5 64
4.9 Regression analysis output for Fig.4.6 64
4 .10 K-valuesfor various thicknesses of
potato slices dried at different air
temperatures 68
4.11 K-values for potato slices dried at
various air temperatures as affected
by slice thickness 69
4.12 Concentration-time curve for ascorbic
acid in potato slices dried at 50°C 76
4 .13 Concentration-time curve for ascorbic
acid in potato slices dried at 60°C 77
4.14 Concentration-time curve for ascorbic
acid in potato slices dried at 70°C 78
4.15 Arrhenius plot illustrating influence
of temperature on rate constant, k for
2 mm thick slices 81
4 .16 Arrhenius plot illustrating influence
of temperature on rate constant, k for
4 mm thick slices 82
4 .17 Arrhenius plot illustrating influence
of temperature on rate constant, k for
6 mm thick slices r
O>O
^
4.18 Trend of browning level during drying
of slices 87
4.19 Drying condition illustrating case-
hardening 92
4.20 Drying condition illustrating non­
occurrence of case-hardening 94
ABBREVIATIONS AND SYMBOLS

ADC - Agricultural Development Corporation

ANOVA - Analysis of Variance

AVG - Average (MR in Figure 4.5, 4.6 and 4.7)

A420 nm - Visible Browning, expressed as optical

density of colour extract (Absorbance of


extract at a wavelength of 420 nm)/10g slices
CIP - Centro International de la Papa (International
Potato Centre)

FW - Fresh weight

RAA - Reduced Ascorbic Acid

TCA - Trichloroacetic acid

A - Area of sample, (cm2)


- Water surface area, m 2.
- Indication of paticle shape, (Equation 2.20)
- Initial area of sample, (cm2), (Equation 4.51)

a-g - Parameters in Equation.4.53 and 4.56

C - Concentration of ascorbic acid (normalized with

respect to initial concentration)

C(t) - C as a function of time


c - Shape factor (Equation 2.20)

D - Diffusivity of material, m 2/s


- D in centerplane node
n - Reference D

- Sample diameter, (cm)

- Potato slice thickness (Eqn.4.28 and 4.29), (mm)

db - Dry basis

E.ex - Activation energy, J/mole

ED - Activation energy for diffusion,

(cal/mole)

f - Thermal conductance of air film, (W/m2k)

i
* ♦ x
fv - Water-vapour pressure transfer coefficient at
the water-air interface, (kg/sm2)
- Latent heat of vaporization, (cal/g)

J - Objective function

h - Heat transfer coefficient, (cal/min°C)


h fg - latent heat of vaporization of water at t
(J/kg)
k - Drying parameter
- Reaction rate constant (Equation 2.21), 1/hr

ko - Constant, (min-1)

L - Sample half-thickness, (mm)


- Potato slice thickness, (cm) (Equation 4.52)

M - Moisture concentration

- Moisture content (db) at any time, t

Me - Equilibrium moisture content (db)


M - Average M

- M in centerplane node

M0 - Initial M

ms - Weight of solids in sample

MR - Moisture ratio, dimensionless

N - Number of volume elements


- Parameters in Equation 4.39 and 4.40
- Parameters in Equation 4.51

- Parameters in Equation 4.46 and 4.48


- Parameters in Equation 4.43 and 4.44

Pa - Water-vapour pressure in the air, kg/mz

Ps - Water-vapour pressure at ts, kg/m2

ppm - Parts per million


R - Universal gas constant, 8.314 J/mole K

R2 - Regression coefficient of determination

T - Absolute temperature, (K)


* ♦
xi
Dry bulb temperature, (°C)
Sample temperature, (°C)
Reference temperature, (K)
Drying air temperature, °C

Drying air relative humidity, %

Potato slice thickness, mm


Time
Air temperature, °C
t for nominal drying process (min)

Normalized time, Equation 4.56

Drying time, (min)

Normalized time, Equation 4.53

Water surface temperature (wet-bulb


temperature), °C
Distance in the sample, (cm)

Initial slice density (g/cm3)


1. INTRODUCTION

The potato (Solanum spp.) is grown in 79% of the world's


countries (FAO, 1986). It is second only to maize in terms of
the number of producer countries and fourth after wheat, maize

and rice in global tonnage. Its importance in European


countries, the USSR, North America, Australia and the Andean

Countries of Latin America is well known. Less widely

recognised, however, is the increasing importance of the potato


in developing countries (Woolfe, 1987).
In Kenya, potatoes are grown twice a year in the cool
high-altitude areas of Central, Eastern and Rift valley
provinces, at altitudes of 1500 - 3000 m above sea level. In
these areas, the potato has greater yield potential than other

food crops and is more suitable as a subsistence crop than

maize which can be grown only once a year (Acland, 1971) . The
potato is increasingly assuming an important dietary role among
Kenyans particularly those in the rural areas, but also in the
large urban centres. The per capita consumption in potato
growing areas is 70 to 100 kg, while that in large urban

centres is approximately 50 kg (Kabira, 1983).

Production and use of potatoes in Kenya is strongly

influenced by prices which reflect the inputs and marketing


costs (Woolfe, 1987). The production and utilisation of fresh
potatoes as a low cost food is held back because of problems

encountered in storage, transportation, and marketing of the


bulky, high moisture, and highly perishable product and by the

wide-ranging seasonal fluctuations in price. An important

limitation is the lack of suitable preservation methods


(Woolfe, 1987). Cold storage warehouses are in limited use due
to high operational and maintenance costs. A possible solution

to this problem would be to process the potatoes in order to


2

reduce bulk and moisture content thereby enabling storage in


less costly stores.

Currently, processed potatoes are available as crisp and


imported dehydrated mashed potatoes. One modest approach to
processing in Kenya would, however, be dehydration. This area
therefore, needs greater input in terms of research. Studies
would be required to establish appropriate drying conditions

commensurate with production of high quality, shelf-stable

products. Such studies could lead to design of low cost energy

saving driers which still maintain high product quality.


Reduction of moisture content of a crop to a value between
0 and 12% is referred to as either drying or dehydration,
depending on the author (Hall, 1980) . Moisture level of

approximately 10% is considered optimum for storage stability


of dehydrated potato products under tropical ambient conditions
(Kabira, 1990) .

The characteristic drying mechanism of potato slices in


a conventional hot-air-drier was studied by Ede and Hales
(1948). Results showed that drying time is considerably

affected by particle size and shape. Solar-drying of potato


products has also been reported for various drying systems

(Gupta, 1978; Bhardwaj, 1981; Shakya and Flink, 1986; Kabira,


1990).

It has been postulated that in potato dehydration a


diffusion type mechanism is operational (Saravacos and Charm,
1962). Gorling (1958), however, argues that a combination of

capillary forces and vapour diffusion transport water to the

surface of potato products during dehydration (Brennan et al.,


1976).

Major quality problems that arise during drying include


loss of nutritive value, enzymic and non-enzymatic

3

discolouration ("browning") and case-hardening. Ascorbic acid


is the nutrient mostly affected by any form of processing. In

view of this, its retention is often used as an index for

nutrient and quality degradation (Maeda and Salunkhe, 1981).


Degradation of nutrients during drying is a function of
several parameters such as time, temperature of sample and
moisture content. These parameters can presumably be

manipulated in such a way as to obtain maximum nutrient


retention. Studies on prediction of nutrient retention in

storage and in thermal processing of potato products have been

undertaken. However, very little work has been carried out on


the prediction of nutrient losses upon drying.
The overall objective of this research was therefore, to
provide scientific information that could be used to design
temperature-time processes for potato slice dehydration with
minimal loss of nutrients. The specific objectives were :

1. To determine the effect of drying air temperature on

quality of the products.


2. To determine the effect of airflow rate on the quality
of the products.

3. To determine the effect of thickness of slices on the

quality of products.

4. To determine optimum conditions for dehydration of

potato slices.
5. To develop drying equations for potato slices.


2. LITERATURE REVIEW
2.1 The Potato as a World Crop

Potatoes are widely grown on a world scale and ranks


fourth in food production, following wheat, maize and rice

(Table 2.1). It tops the list for root crops, followed by


cassava, sweet potatoes and yams respectively. The greatest

potato production per country is in the USSR (73,000

kilotonnes), USA (18,331 kt), India (12,642 kt), and the German
Democratic Republic (11,500 kt) , with the Federal Republic of
Germany, France and the UK fluctuating around 7,000-8,000 kt
(Hawkes, 1990) .

Table 2.1 World production of major starch crops (cereals


tubers) for 1996 in metric tons (000)

Wheat 584,874
Maize 576,821
Rice 562,259
Potato 294,834
Cassava 162,942
Barley 155,261
Sweet potato 134,244
Sorghum 69,000
Yams 33,110
Oats 30,967
Millet 29,563
Rye 23,156

(FAO Production Yearbook 50, 1996)

In terms of production value per unit land area per day

in developing countries, the potato does better than rice and

wheat for edible energy and only lags slightly behind wheat for

edible protein, as shown in Table 2.2.


5

Table 2.2 The ten food crops with the highest production value
per hectare per day in developing countries

Crop Growth Dry Edible Edible


duration matter energy protein
(days) (kg/ha/ (000 (kg/ha/
day) kcal/ha day)
/day)
Cabbages 110 12 29 1.6
Tomatoes 125 8 25 1.3
Potatoes 130 18 54 1.5
Yams 180 14 47 1.0
Sweet 180 22 70 1.0
potatoes
Rice, paddy 145 18 49 0.9
Groundnuts 115 8 36 1.7
in shell
Wheat 115 14 40 1.6
Lentils 105 6 23 1.6
Cassava 272 13 27 0.1

(Hawkes, 1990)

Thus, potatoes are clearly a crop of considerable world

importance, fourth in world production and of high nutrient

value.

2.2 The Origin of the Potato Plant and its Development in


Kenya

The short-day, late-maturing S.Andigena types were

introduced into Europe, possibly from South America, in the


second half of the sixteenth century (Burton, 1989). It gained
recognition as an inexpensive and nutritive food during the

18th century (Burton, 1989). When potatoes became a common


crop in Europe the Andigena types turned by selection into

Tuberosum types which are tolerant of long days (Ballestrem and

Holler, 1977) . This explains why potatoes do well under long-

daY and short-day conditions.



6

The potato was introduced into the then British

Protectorate of Kenya at the end of the nineteenth century.


White settlers started growing it in the "white highlands"
which the government
/' had reserved for European settlement.
Because of its origin and to differentiate from sweet potatoes
(Ipomea batatas), potato in Kenya is generally referred to as
English potato, or Irish potato (Durr and Lorenzl, 1980) .

During the advent of World War I, Kenya's potato industry


was viewed as an important source of food for the British
armies in Africa and Asia. After the war, African farmers in
Kenya began to grow potatoes for domestic use as well,
primarily in Rift valley. During the period between World War
I and II, government facilities were established and expanded

to produce new varieties and to conduct agronomic studies

(Crigsman, 1989) . With this support, the potato established


itself among the main crops in the country, despite the
competition from commercial crops. Some degree of
specialization evolved in the potato industry, whereby most of
the export crop was provided by the African farmers and

European farmers produced for seed and for domestic ware-


market.

During the second World War (1939-1947), potatoes were


again among those commodities which were required for supplying
British Armies. Plants were set up at Kerugoya and Karatina
to dehydrate potatoes and other vegetables for army supply.

In 1946, the two plants processed about 5,000 tons of potatoes


within a 6 months period (Durr and Lorenzl, 1980) .

After World War II, the government supported research


primarily on disease control while importing new seeds. By
that time, potato production had moved from Rift Valley to
Kiambu, Murang'a and Nyeri Districts (Crissman, 1989) .

7

Shortly after independence (1963), many expatriates who


had been involved in the national potato programme left the
country, and it took time to replace them with Kenya

professional staff. Responding to a shortage of seed in the


late 1960's, several varieties were imported from West Germany

(Durr and Lorenzl, 1980). The government reestablished the

scientific basis for potato improvement by entering into

cooperative programs with the German Technical Assistance


agency and the British Overseas Development Authority.
Starting in 1968, work was begun on establishing a National

Potato Research Station in Tigoni and three allied substations


(Crissman, 1989).

In 1977, it was estimated that about 30,000 to 40,000 ha


of potatoes were grown annually (Ballestrem and Holler, 1977).
In 1986, the estimate was 75,000 hectares, with an average
yield of 10 metric tons per hectare (Crissman, 1989) . It
should be noted, however, that estimates of potato production
vary, depending on the information source (Durr and Lorenzl,
1980). Potatoes in Kenya are unfortunately still regarded as

a subsistence crop. But with increasing experience, farmers

and consumers are learning the high food value of the crop and
its potential.

The Kenyan highlands, with altitudes ranging between 1600


and 2700 m above sea-level, are the ideal potato growing areas.

Kibirichia, Molo, South- and North Kinangop, Nyahururu, Nyeri,

Embu, Limuru, Kiambu and Taita are only a few of those where

farmers grow potatoes (appendix G ) . The potential of potato


cultivation has not been sufficiently exploited. Kibirichia
(Meru) and Kinangop are the only areas of concentrated potato
growing in Kenya (Ballestrem and Holler, 1977).


8

2.3 Potatoes in the Kenyan Food System


Potatoes is now an essential staple food in some areas in
Kenya. In a recent comparative study (Ballestrem and Holler,
1977) , potatoes were shown to mature four times faster, to
yield seven times more kilograms per hectare and to provide

more kilocalories and almost twice the amount of crude protein

than maize, the local staple. Horton (1987) points out that
the combination of high yields, short vegetative cycle and high
price makes the value of potato production per hectare higher
than most other food crops grown in developing countries.
In relation to the increasing production and the volume
of the domestic market, potato export has attained a stage of
insignificance. Most of the crop is consumed within the

production areas, the market surplus is channelled into


domestic food marketing systems and constitutes an important
item in urban food supply.

A characteristic element of the potato market are supply


originated price fluctuations. This is aggravated in the
producing areas. Storage is considered a major tool in

balancing out seasonal imbalances in supply. At present,

storage facilities on farms and along the marketing channels


are inadequate in capacity. Storage losses are generally high.
The Food Policy Sessional Paper No. 4 of 1981 points out that

the supply of food actually available for consumption can be


increased substantially through reduction in storage and

handling losses. Losses during storage of potatoes are


estimated to be as high as 19% (Kabira, 1982).
Under Kenyan conditions, storage in simple low-cost stores
on farms is more profitable than with technically more
sophisticated storage facilities. Farmers, however, have not
developed storage practices, which
i are especially suitable for

9

potatoes. Potatoes for consumption are stored for 2 to 3

months, which is not enough to bridge the period till the next

harvest. Hence, potatoes are generally consumed fresh and


processing is limited to the production of potato crisps.

It needs to be recognized that potato lends itself to a


$
variety of processed food products. Dehydrated, mashed,
granulated, chipped, freeze-dried, canned ones are all popular
brands of potato. Potato flour is the oldest commercial
product used with advantage by the industry throughout the
world. Manufacture of potato flour is a simple process based
on dehydration of cooked potato. These are just a few lines
along which potato industry can be built to provide high-
quality food and augment the existing food resources in the
country.

, Increased processing of potato will lead to the following


advantages:-

(a) Ensure off-season supplies by providing a means to


carry over surpluses from one season to the next,

(b) Ease distribution problems experienced during


handling, transportation and storage,
(c) Contribute to food diversification, thus reducing
over-reliance on consumption of maize,

(d) Reduced pric fluctuations, thus ensuring consumers


a constant supply of potatoes at reasonable prices
throughout the year,

(e) Enable the export of surpluses to earn foreign


exchange and

(f) Provide employment opportunities in agriculture, agro­


industry and in research institutions.
10

2.4 Processes to Extend Potato Storage Life

Kenya's population is estimated at 24 million. The annual

growth rate, at about 4%, is one of the highest in the world.


Agriculture is the leading sector and the expansion of food
crops is therefore expected to finance development in other
sectors of the economy, but Arable land is limited (Crissman,

1989). Such situations have provided the greatest stimulus to

the development of methods for preserving the potato crop by

dehydration in various forms. A besieged country has no option

but to turn to this crop and learn to adjust its daily diet
(Gray and Hughes, 1978) .
In industrialized countries, the substitution of fresh
potatoes with processed products as food is attributed partly
to enhanced efficiency in processing to produce relatively low-
pricgd dried, canned and frozen forms, and to factors such as
increased demand for convenience foods, fast foods and snacks
(Woolfe, 1987).
In Kenya, a combination of such factors is lacking. The
high cost of potatoes, lack of market infrastructure for

handling frozen foods, and absence of a large restaurant trade,


make it unlikely that the demand for convenience foods such as

chips and precooked frozen french fries will expand

dramatically in the foreseeable future. Moreover, where fresh

potatoes are available for much of the year, it seems unlikely


that the demand for processed products will make inroads into
the demand for fresh potatoes (Horton, 1987) .

2 »4.i Potato Dehydration

Dehydration as a method of food preservation is of unknown


antiquity. The preservation of potatoes by drying was one of
the methods developed and used in the Peruvian Andes at the

11

time of the Incas, 2000 - 3000 years ago (Van Arsdel, 1973a).
The traditional rural process forms the basis of the

modern potato-processing industries which are beginning to

emerge in Peru. The household method of preparation involves


9 '
selecting and cooking the potatoes; hand peeling, cutting or
crushing; and then drying in the sun on trays or sheets of
metal, sometimes on the roof of the house. The moisture

content is reduced by this means to a safe moisture content

(under 15%) below which the risk of mould growth during storage

is considerably reduced (Burton, 1989). The dried product is

an unattractive dark brown colour with a poor texture of


irregular size, and is contaminated with dust and small stones.
Yields and quality are generally low because of lack of
knowledge of the basic principles of production control. It
is estimated that the yield of dried potato is below 15 per

cent of the original raw material weight, after 10 to 15 days


drying of the fresh material (Montero, 1991) .
Some elements of the traditional process needed to be
modified to improve quality and enhance efficiency. The
industrial method of preparation as described by Montero (1991)
is as follows. Potatoes are selected by variety and size.

This improves the quality and yield, and reduces cooking time.
The potatoes are peeled and washed simultaneously in an

abrasive peeler, after which they are put in a container with


(preferably) hot water to prevent the enzymatic reactions which
cause discoloration. The pre-cooked and sliced potato enters
the solar dryer with about a 78 per cent moisture content,

which falls to 9 per cent in the final product. A variation

the process which can control the colour before and during

drying, and which also stops the growth of micro-organisms is


treatment with a 0.025 per cent Sodium Metabisulphite solution.

12

The average yield is about 20 per cent. The colour and

rehydration characteristics were also acceptable to the


technical standards institute of Peru (Montero, 1991).
Potato strips and dice, produced similarly, were made in
9

Europe and the USA in wartime, based on unsophisticated methods


%
$
which lent themselves to use 'in numerous small factories. The
essentials of the process are: peeling and trimming of the

potatoes; cutting them into pieces; inactivating the enzymes

which would cause discoloration in prepared material before and


during drying (prior to their being inactivated by heat in the
drying process); and then drying at a temperature which does
not cause browning of the pieces, down to a moisture content
of 6-8% (Burton, 1989).

Browning results from the reactions between sugars and


amino^ acids present in the potato. Their rate is much

influenced by temperature and by the concentrations of the


reactants- which depend not only on the amounts originally
present in the raw material but also on their increasing
concentration as the removal of water proceeds (Hendel et al.,
1955) .

A typical wartime practice (1940s) as described by Burton


(1989) was to employ a temperature of 95°C at the start of the
drying process, falling to 70°C when the potatoes were nearly
dry, in simple tunnel driers in which the pieces of potato were
placed in shallow trays. The piece size used were either about
5 x 8 mm in cross-section and of a length determined by tuber
size, or cubes of about 5-7 mm side. Enzyme inactivation,

immediately after cutting up and washing the pieces, was by


blanching' them for 2-3 minutes in water held at or just below
100° c # or by steam. If trimmed pieces had to be kept for some
time before blanching, Sodium Sulphite or Metabisulphite was

13

applied in amounts sufficient to give Sulphite in the dried


product equivalent to 300-500 p.p.m. S02, or whatever upper
limit required by local legislation. Drying time was up to
about eight hours.

The presence of sulphite during drying protects against


$
the development of a brown colour, a charred taste and impaired
rehydration of the dried product. This enables higher
temperatures than would otherwise be possible to be employed,
thus reducing the time necessary and improving the quality of
the product. Sulphite in the dried product also retards the
development of a brown colour if storage under warm conditions

is necessary (Burton, 1989).

In peacetime, the principles of the method of manufacture


remain the same, but improvements in the process' and in the
quality of the product have been effected by changing the size
of the pieces of potato, the type of drier and temperature of
drying.

The larger the piece of potato the slower will be the


drying process (although the less will be losses during

preparation). At any cut surface of a tuber the contents of


the broken cells will be lost. In the case of the normal
peeling of a potato this loss is small in relation to the
amount of peel discarded. Where it really becomes of
importance is in treatments in which the tuber is considerably
subdivided, as in slicing for the preparation of crisp or dried
slices. In this case, where the dry weight of the raw material

is an important determinant of yield in commercial production,


the loss of starch from broken cells can be a serious matter.
Reeve (1971) calculated that, with slices 2 mm thick, losses
of dry matter from broken cells could be expected to amount to


14

8-12% of that originally present in the slice, and, with slices

1 mm thick, 16-24%.

Kueneman (1975) used a starting temperature of 135°C


falling to »80°C as drying progressed. This brought the smaller

sizes of potato piece down to a moisture content of 25-35% in

about one hour. Thereafter two to three hours at about 65°C


brought it down further to 10-15%, and final drying to about

6% was done in a finishing drier at about 40-60°C (Burton,


1989) .

The most commonly used method for dehydration is drying


by means of warm air. Often the aim of dehydration is to
reduce the moisture content while retaining the original
quality of the product as far as possible.

Solar-drying of potato products has been reported in the


literature for various drying systems. Gupta (1978) solar-
dried 5 mm thick potato slices using a 2 stage system for
heating the inlet air. At a 30 kg/m2 loading, the moisture
content after 10 hrs of drying had fallen from 386 to 23 g
water/100 g solids. Bhardwaj (1981) compared solar and natural

sun drying in India. For potato sticks, a solar drying time


of 8 hrs gave a moisture content of 5%, while natural sun

drying required 10 hrs to achieve a moisture content of 7%


(Shakya and Flink, 1986).
The effects of commercial dehydration on the nutritional
value of potatoes has been investigated repeatedly (Augustin
®t al., 1979). Most of the studies were laboratory or pilot

Plant experiments and analyses were limited to comparative

nutrient evaluations of the raw versus the dehydrated product.


Most standard methods of ascorbic acid analysis in raw
Potatoes are considered unreliable (Moledina and Flink, 1982;
shakya et al., 1986), when applied to potatoes dried at 60°C
t
* ♦
15

or higher, mainly due to the amino acid-sugar interactions

which yield products that analytically simulate ascorbic acid.

Up to 60% loss of ascorbic acid was realized during dehydration


of potato slices (Augustin et al., 1979).
Kabira
(1990) ‘ investigated the possibility of drying
o
potato slices for flour production by utilizing natural air
flow and solar energy. The results indicated that no

technological, product quality or storage stability advantages

were gained by blanching or cooking potatoes prior to

dehydration. An acceptable dried product could be obtained by


drying raw slices, essentially eliminating the usual practice
of either blanching or cooking. The successful production of
dehydrated raw potato slices with acceptable physicochemical
and sensory properties appeared to be associated with the slow
remoyal of water at the relatively low temperatures (below 40°

C) and the prevention of surface browning by use of sulphite


to inactivate the enzyme polyphenol oxidase. However, the sane
results may not hold at higher drying air temperatures, hence,
there is need for further research on the dehydration of raw
potato slices.

2.4.2 Potato Drying Technology

Dehydrated potatoes in piece form have been successfully


produced in many different types of cabinet, tunnel and
conveyor driers. Until recent years, the tunnel drier was the
predominant type.2

2*4.2.1 Tunnel Driers

Tunnel driers are the most flexible systems in commercial


Use. They are of widely differing designs, but generally of
either counterflow or centre exhaust
l arrangement. The simplest
*
16

form consists of a rectangular tunnel which accommodates trucks

containing the trays on which the product to be dried is


uniformly spread. In the type commonly used, drying takes
place in tunnels about 9 or 12 metres long and 1.8 m high by
1.8 m wide (Von Loesecke, 1943).

The tunnel system is essentially a continuous one in which

the trucks gradually enter one at a time and are similarly

removed at a constant rate when equilibrium conditions are


established. Air flow in the tunnel may be in the same
direction as the trucks (parallel-flow); opposite to travel of
the trucks (counter-current); or parallel flow at the entrance
and counter-current at the exit, or dry end, with air
exhaustion at the centre (centre-exhaust type).

Parallel-flow tunnels are considered the least efficient

for dehydration. It has the advantage that the hottest air


contacts the wettest product, therefore hotter air can be used.
This means that high rates of evaporation are achieved
initially and there is little danger of overheating the
product, since the surface temperature of the food is below the

dry bulb temperature. The product is also protected in the

latter stages of drying since it comes into contact with


progressively cooler air and this is accompanied by a decrease
in the drying rate. On the other hand, the air at the outlet
end becomes cool and moisture laden and this has the ultimate
disadvantage that very low moisture contents cannot be achieved
irrespective of length of tunnel. Where this type is in use,

it is necessary to finish drying in bins with an updraught of


warm (60-67°C), dry air.

In the counter-current flow tunnels, the hot and dry air


contacts the dry product leaving the drier. Thus the "wet"
Product is initially exposed to very slow drying conditions.
i
* *
17

Care must be taken not to overload the drier as the moist


charge may stand in the warm, moist air too long without being
dried to any extent. This would allow time for product

spoilage. On the other hand, the dry product should not be


left in the drier too long since it is in contact with the

hottest air and could become overheated. In general, the

counter flow tunnel uses less heat and produces a drier product
than a parallel flow tunnel. However, shrinkage of the product
occurs when using countercurrent rather than parallel flow.
The result is that the former gives a material of relatively
high density and slow reconstitution properties whereas the
latter gives a product with low density (Holdsworth, 1971) .

In some cases the two types of tunnels are combined into


one unit (centre-exhaust tunnel). The product is first placed
in a, parallel flow tunnel to take advantage of the high
initial rate of drying in this type of tunnel. It is then
placed in a counter-current flow tunnel to get a sufficiently
dry end product. The advantages compared with single stage

units are more uniform drying, increased output and good

overall quality. Finishing bins may also be applied in the

case of counter-current or centre-exhaust tunnels if the aim


is to increase their capacity. The air velocities in tunnel
driers range from 0.8 to 6 m/s (Von Loesecke, 1943).

Typically, a dehydration plant operating a counter current


drying tunnel would have a capacity of 0.5 ton/hr, involve 12

trucks, air temperature 50°C dry-bulb and air velocity 3-3.6


n/s. Total drying time required would be of the order of 8 hr
(Holdsworth, 1971) .


18

2 .4 .2.2 Cabinet Driers

These driers consist of a drying chamber holding trays,

in large driers, the trays are placed on a track for ease of

handling; in small units, trays may be placed on permanent


supports in the drier. Circulation of air is usually across
ig
the trays. In some cabinet driers, air is forced through the
trays, in which case drying is more rapid. Air having a flow
rate of 3 to 6 m/s or more is carried from the heaters through

a main duct from which it can be diverted to the stack of

trays. A cabinet drier of about 90 m 2 of tray area would

handle 450 to 680 kg of prepared material every 6 to 7 hrs (Von


Loesecke, 1943) .
Cabinet driers are used where production is on small
scale. The other appropriate application is in laboratory and
pilot-scale experimentation on the theory and practice of

drying. For example, a properly equipped cabinet drier can be

operated to give a close simulation of the performance of a


tunnel or a continuous conveyor drier (Van Arsdel, 1973a). The
least satisfactory element in cabinet drier performance is
likely to be the distribution of air flow and temperature
across the face of the tray compartment.

2.4.2.3 Conveyor Driers

Diced and sliced potatoes are currently dried on these


driers. it is similar to the tunnel system but the material
is conveyed through the hot air system on a continuous moving
belt. Conveyor driers have the advantage that the high cost

handling products both before and after drying using trays

considerably reduced because loading and unloading are


Carried out automatically.


19

The process conditions are primarily controlled by

sectionalizing the drying system. The conveyor drier is

designed in stages so that it is possible to control


independently flow rates, humidities and temperatures in order
to give optimum output and quality. The entire system lends

itself well to complete instrumentation and automation.


The wet material is loaded uniformly (0.15-0.2 m deep)

onto the belt (either a woven metal mesh or interlocking

perforated plates) and dried initially by air blowing through


the bed and finally by air passing down through the bed. This
ensures that material is not lost due to changes in density
during drying (Holdsworth, 1971). The belts vary in length
from 18-23 m and in width from 2.4 to 3 m (Van Arsdel, 1973b).
Conveyor driers are essentially for high capacity water removal
and^it is necessary to remove the material at a moisture level

of about 15% to avoid the use of excessively long belts and


waste of evaporative capacity. A typical drying operation
would involve 35 tons/hr raw material infeed onto a 23 m long
2.4 m wide conveyor belt loaded at a rate of 4.5 kg/hr and
allowing 2.5 hours for travel through the system (Holdsworth,
1971).

2.4.3 Theory and Principles of Drying


The following were proposed by Gorling (1958) as possible
mechanisms for removal of moisture from agricultural materials
such as potato, wood or macaroni (Van Arsdel, 1973a) ;

1) • Liquid movement due to surface forces (capillary


flow);

2) • Liquid movement due to moisture concentration


differences (liquid diffusion);


20

3) . Liquid movement due to diffusion of moisture on the


pore surfaces (Surface diffusion);
4) . Vapour movement due to moisture concentration
differences (vapour diffusion);

5) . Vapour movement
due to temperature differences
$
(thermal diffusion);

6 ) . Water and vapour movement due to total pressure

differences (hydrodynamic flow).


The following tensor system of partial differential
equations have been developed based on the above physical
mechanisms (Brooker et al., 1982):

= V*klxM + V*^ia0 + (2.1)

- V*k2lM + V2* ^ ♦ V*k33P (2.2)

= V*k3lM * V2*,^ + V2*,,? (2.3)

^ll» k22, an<^ ^33 are the phenomenological coefficients. The


other k-values represent the coupling coefficients. - The
coupling results from the combined effects of the moisture,

temperature, and total pressure gradients on the moisture,


energy and total mass transfer.

In artificial drying of agricultural crops, the


circumstances allow the pressure terms to be dropped. Hence,
the system of equations modifies to:-


21

(2.4)

(2.5)
i

The coupling effects of temperature and moisture has a

very limited application (Brooker et al., 1982). Therefore,


the phenomenological equations become:-

(2 .6 )

(2.7)

In practical analyses, the temperature gradients are

ignored. Thus, the ultimate simplification of the equations


becomes:-

(2 . 8 )

Where it is generally accepted that moisture flow within


the drying material takes place by diffusion (liquid and/or

vapour), the transfer coefficient k n is called the diffusion


coefficient, D, m 2/s. Saravacos and Charm (1962) reported

experimental results on the air drying of potato dice or slice


which appear to be consistent with the molecular diffusion
mechanism (Van Arsdel, 1973a).


22

For a constant value of D, Equation 2.8 can be written in


the following generalized form (Brooker et al., 1982):-

' dM _ p, & M + c dM-,


(2.9)
dt dr2 r dr

where c is the shape factor. It is zero for planar symmetry,

M is the moisture content, dry basis, (db) at time t hrs, r is


a characteristic dimension of the particle and D is the
diffusion coefficient, m 2/s. The following initial and
boundary conditions are applicable in solving the above
equation:-

M(r,0) = M(0) (2 .10)

M(r0, t) - M(eg.) (2.11)

The complete analytical solution of this equation can be

obtained from Crank (1975). For an infinite plane, the

following solution apply (Brooker et al., 1982).

n—
MR T exp [ -Jlnillllix!] (2 .12)
H 2n o 2n+l

Here, the average moisture content and time are expressed as

the dimensionless quantities, moisture ratio, MR and X,


respectively:


23

MR - (2.13)

and
a -
X = — (Dt) 3 (2.14)
V

where A represents the surface area and V the volume of the


body. For the plane, the ratio A/V = half thickness.
Conventional drying systems for agricultural products,
commonly employ convection drying and can generally be analyzed
as either "thin layer" or "deep bed" systems. Thin layer
systems characteristically limit the depth of the product in

the direction of air flow such that as the air passes over or
through the product its properties are not measurably affected;
thus," the product is assumed to be dried in a constant
environment. Deep bed systems, on the other hand, utilize
product depths in the air flow direction such that the
properties of the air are continuously changing as it passes

over or through the product. The environment surrounding the


product then becomes a function of time.
It is obvious, therefore, that the thin layer system is
the easier of the two systems to analyze. In fact, the deep
bed system is so complex that most of the algorithms, lor
analysing such systems assume that the bed can be divided into
a number of "thin" layers and that for small increments of time

the properties of the air passing through each individual layer

can be considered to be constant. Thus, the development of


^ i n layer drying models is basic to the analysis of convection
^tying regardless of the type of system used.

Chiang and Petersen (1985) report experimental data for


thin layer air drying of ♦French fried Russet Burbank
24

potatoes (Petersen, 1986). Raw potato pieces cut in the shape


of French fried potatoes were dried from their initial moisture
content to near equilibrium moisture content. Weight versus
time data were collected on one minute intervals by interfacing
a load cell with a' computer. Petersen (1986) describes a
technique which will numerically determine whether a constant

rate period does exist, the drying rate during the constant

rate period, the critical moisture content, the equilibrium


moisture content, and parameters for any thin layer model of
the falling rate period which the investigator wishes to test,
including various sub-periods during the falling rate period.
The system uses the nonlinear curve-fitting capabilities of the
Statistical Analysis System (SAS) to perform the analysis. The
technique was applied to potatoes dried in a hot air stream

having a dry bulb temperature of 93°C, 10% relative humidity,


and air velocity of 1.06 m/sec. The data predicted by the
equation fit using the SAS program represented the actual data
well. The residual sum of the squares for the 134 observations
was 104, indicating a good fit (Petersen, 1986).

2.4.4 Rata Periods of Drying

Drying processes can be divided into two periods


(a) The constant drying-rate period.
(b) The falling drying-rate period.

In the constant-rate period the potato slices contain so much


water that liquid surfaces exist and will dry in a manner

comparable to an open-surface body of water. This occurs


immediately the slices are brought into contact with the hot
air stream. The evaporation of moisture from the surfaces is
Very rapid. The rate of surface evaporation must therefore be
Proportional to free moisture on the surface, which decreases

25

as drying progresses. This mode of drying will begin to change

as soon as the slice surface dries up, and further drying will

depend upon the rate of moisture or vapour migration from


i
within the slice to its surface. The new mode is referred to
i
aS the falling drying rate period.

The constant-rate drying depends on :-


(a) area exposed,

(b) difference in humidity between air stream and wet


surface,

(c) the coefficient of mass transfer and


(d) velocity of drying air.

If drying is by passing of air through the mass, the


following heat and mass balance exists as expressed in equation
2.15 (Henderson and Perry, 1976):

^ * fyMP.-pJ = fA- taL ta) (2.15)

where

^ = drying rate, kg/s o£ water.

f " thermal conductance of air film W/m2k.


A ■ water surface area, m 2.

latent heat of vaporization of water at tB (J/kg).


= air temperature,°C.

" water surface temperature (wet-bulb temperature),


°C.

fv " water-vapour pressure transfer coefficient at the


water-air interface, kg/sm2.


26

pB = water-vapour pressure at tB, kg/m2.

pa = water-vapour pressure in the air, kg/m2 .

The rate of drying is equal to the vapour pressure driving

force divided by the resistance to drying. The driving force


of moisture movement is given as (p8 - pa) and the resistance
to drying is l/fv A (Hall, 1980). Drying in the falling-rate
period involves two processes:-

(a) movement of moisture within the material to the


surface, and

(b) Removal of the moisture from the surface.

For the falling rate period, when liquid diffusion is


controlling, the movement is analogous to that of heat
conduction. The solution of the equations that hold in this
case yields (Holdsworth, 1971):

M-M0 e -25b
+ ........... ) (2.16) .
25

for the moisture removal from or drying of a slab of infinite


extent. Where

M = Moisture content, dry basis, after a period of


time, 0 , hrs.

M0 = average moisture content at the start of the

diffusional period, 0 = 0 .
Me = average moisture content when the sample is-in

equilibrium with the surroundings, determined from


the relative humidity of the drying air.

(2.17)
b=Dv0n2/4L2

Dv = Diffusivity of the liquid, m 2/s.

k ■ Half thickness of the sample.


e<3uation also applies to an initially uniform moisture

27

distribution. If 0 is large then

M~Ma (2 .1 8 )
M 0- M a

and the drying rate is given by


>

dM (2.19)
- n 2D v (Me-M) /4L 2
c/9

Thus the drying rate for materials requiring large drying times
is inversely proportional to the square of the thickness of
sample when diffusion is the mechanism controlling the drying
process.
When a thin layer (a layer one particle deep) is dried by
passing air with a constant temperature and humidity through
the layer, a curve of (M - Me)/(Mo - Me) vs 0 can be developed
from the loss in weight of the material as it is dried. The
equation of the straight line portion of the curve is

M0 M a = A°e 'M (2 .20)

A 0 is an indication of the shape of the particles because for

a slab, A 0 - 8 / tt 2 =0.811, Brick, A 0 = (8 /rr2)3 = 0.533, sphere,

*0 “ (6 /rr2) ■ 0.608. A 0 is the ordinate intercept of the


straight line portion of the curve.

Igbeka et al. (1976) found that in drying cassava and


potato, no constant rate period was present (Igbeka, 1982) .
The drying proceeded in the falling rate period. A method of

measuring variable moisture diffusivity experimentally, was


aiso investigated. In a later work, Igbeka (1976), a
sthematical representation of variable diffusivity of moisture

28

during the drying of cassava and potato was presented (Igbeka,

1982) .
2 .4 . 5 Drying Under Simulated Practical Conditions

Experimental study of drying should employ a carefully


designed drier equipped for precise control of the temperature,
humidity, and velocity of the air stream in which the body is
supported. These external factors related to the drying

mechanism, are governed by relatively simple and well-known

laws, but not so the internal transfer of moisture.

Ede and Hales (1948) used a cabinet-type experimental


drier of fairly conventional design in their study of the
drying of potato and carrot strips (Van Arsdel, 1973a). The
method involved utilizing 24 measured pieces simultaneously
on a wire-bottom tray, with air flowing down through the almost
unobstructed tray. The results were comparable to those

obtained by single piece thread-suspension method.

A few investigators have reported studies of the air


drying of isolated pieces of wet food materials, planned and
carried out in order to separate the effects of all the
important variables. One of the most significant work was done
by Jason (1958), on the study of the drying mechanisms of lean
fish flesh (Van Arsdel, 1973a). The drying apparatus

automatically recorded drying conditions and sample weight, and

was operated continuously over long periods of time. Frequent


measurements enabled analysis of the drying rates to be made
in great detail.

The work done by Saravacos and Charm (1962) was on drying


®xperiments of single layers of blanched potato di ce. Gross
drying rates (i.e. rate uncorrected for shrinkage of drying

rea) a lways showed a short phase of constant rate and a fairly


arP break at a critical moisture content of 3.5 (db) .
* ♦
29

A somewhat irregular falling rate phase then followed.

2.5 Changes in Potato Tissue During Drying.


(a) Shrinkage effects and case-hardening.
I
Plastic deformation takes place during drying. The volume
shrinkage may be accompanied by various kinds of damage such
as cracking of the tissue. Drying without warping or internal

cracking calls for careful adherence to an empirically

determined safe schedule of drying conditions (Van Arsdel,


1973a) .

If the temperature of the air is high and the relative


humidity is low, there is danger that moisture will be removed
from the surface more rapidly than water can diffuse from the

moist interior of the food particle, and a hardening or casing


will^be formed. This impervious layer will retard the free
diffusion of moisture. This condition is referred to as "case
hardening". It is prevented by controlling the relative
humidity and temperature of the circulating air.
(b) Browning or "Heat Damage".

The most obvious and troublesome of the irreversible

changes that may accompany the drying of a food substance is

the colour change called "browning" or simply "heat damago".


It is commonly associated with overheating. In food
dehydration, it is generally a most serious guality defect.
If the degree of browning is not great, the change in colour
raay be the only noticeable effect, but when the change proceeds
further the flavour, the rehydration capacity, and ascorbic
®cid content may also be adversely affected (Van Arsdel,
l9?3a) .


30

Values representing the brown colour developed during

dehydration of potato slices varied from 0.09 to 0.11 (Kabira,


1990) . This range was much lower than the maximum acceptable
value of 0.30 obtained by Hendel et al. (1950) for dehydrated
potatoes. The rehydration value (g water pickup/g dry matter)
for dried raw slices was 4.24 (Kabira 1990). However, Kabira
(1990) reported the levels of residual S02 in the dried raw
slices as 680 ppm, slightly above the upper limit of 500-550

ppm recommended for conventionally processed potato products

(Roberts and Mcweeny, 1972; Hadzidev and Steele, 1979).


(o) Nutritional Changes.
Although dehydrated potato products have considerable
commercial value and offer convenience in food preparation,
nutritional considerations may limit their widespread
acceptability among marginal populations of developing

countries (Woolfe, 1987). The effects of commercial


dehydration on the nutritional value of potatoes has been
investigated repeatedly (Augustin et al. 1979). Woolfe (1987)
reviewed earlier work on the composition and nutritional
quality of both the conventionally dehydrated potato products
and the traditionally processed potatoes in Peru and Bolivia

(Woolfe, 1987) . The traditional products resulting from

processing of potatoes were reduced in nutrients, especially


nitrogenous constituents and ascorbic acid in relation to the
raw material (Woolfe, 1987). Comprehensive reports on the

nutritional changes during individual unit operations such as


drying are, however, lacking.

Augustin et al. (1979) found that overall retention values


ranged from 4% for thiamine to 85% for protein. Vitamin C

retention were lowest where the potato slices were exposed to


^*9h temperatures for prolonged periods of time. Sulphitinq
i
* ♦
31

had a sparing effect on vitamin C but a deteriorative effect

on thiamine, resulting in losses of up to 96% of the vitamin.

The other vitamin most severely affected by the high

temperature, long time processing was folic acid, of which the

retention were as low as 40%. Retention of niacin and vitamin


B6 were well over 50%. The main losses of all these vitamins
were, however, by leaching during blanching (Augustin et al.,
1979).
2.6 Theoretical Equation for Quality Loss.

The term 'quality' refers to a combination of many


properties. All these properties change during thermal
processing, some relatively fast and others slowly; so there
is a gradual change which usually means a deterioration of
quality. The changes are caused by the activity of micro-
orgeinisms, by endogenous enzymes and by chemical reactions;
physical or physico-chemical changes can also take place.

Since simultaneously occurring conversions usually proceed at

different rates, a change which takes place at a higher rate


will be noticed sooner than others.
Quality can be defined in this context as the sum total
of all those attributes which can combine to make the potato
slices acceptable, desirable and nutritionally valuable as a
human food. The most common and generally valid assumption is

that temperature dependence of the deterioration rate of food

quality will follow the Arrhenius equation:-

K = K0exp(-Ea/RT) (2.21)

Where; K - Rate constant in the nutrient destruction

reaction. 1/hr


32

K0 = Constant (independent of temperature).


Ea = activation energy. Joules/mole.

R = Gas-law constant. Joules/mole K


T = absolute temperature. K
9

The activation energy is generally derived from the slope of

the plot of In k versus 1/T and depends on compositional


factors such as water activity, moisture content, solid
concentration, Ph and others. The Arrhenius equation has been
found to apply to agricultural materials (Henderson and Perry,

1976). It finds application in the prediction/ interpretation


of nutrients, colour, texture and flavour changes.

In this work, ascorbic acid retention was used as an index


of relative nutrient loss during the drying process. Vitamin
C is present in the potato in both the reduced (ascorbic acid)
and^oxidized state (dehydroascorbic acid) but the content of
the latter is usually considerably lower (0-14%). The
concentration of vitamin C in the potato tuber ranges from 3-40
mg/lOOg fresh weight (FW) and is not influenced by tuber dry
matter content or by size and it appears to be evenly

distributed throughout the tuber (Gray and Hughes, 1978) . The


visible colour was evaluated as a representative sensory
quality aspect.

Results of experimental work presented in this thesis will


point out the influence of the product's linear dimension and
drying conditions on the drying process. Correlation of

temperature, humidity, and velocity of air over the material


to be dried has been discussed in detail by Van Arsdel et al.
t973a. in addition, technical information is available on
drying of potatoes (Ede and Hales 1948; Gorling 1958; Saravacos

3°d charm 1962). However, all the results reported have been
n blanched or scalded potatoes, not on the drying of r<iw

33

potatoes, where the gelatinization of starch granules and the

behaviour of starch seems to be important. The Crank's

diffusion equation and the Arrhenius equation, however, tire


considered relevant for the current study.
% i
i


3. METHODOLOGY
This chapter provides the details of data acquisition,

their possible limitations and the analysis. The data were

mainly from drying experiments carried out to establish the

drying behaviour of the potato slices and to determine the


drying parameters for modelling.

The various instruments and equipment that were used for


data collection and analysis are also presented. For drying
experiments to be meaningful the method and precision in dala
collection has to be reproducible.

3.1 Raw Material

The Potatoes used for the study were of the Desiree

variety and were obtained from the Agricultural Development


Corporation (ADC), Molo. The potatoes had been grown during
the May-September season and harvested in October and November,
1991. The potatoes were cured at 18°C and 75% RH (relative
humidity) of the air for 2 to 3 weeks in a common store

belonging to the International Potato Centre (CIP), at the


Kabete campus of the University of Nairobi before they were
processed. A thermohygrograph was used to give charted
readings of temperatures and relative humidity during the
curing stage.

3*2 Preparation of Potatoes for Processing

The potatoes were sorted to remove those unfit for


processing due to "light greening", rot, or mechanical injury.
They were then thoroughly washed in running tap water to remove
adhering soil and peeled in an abrasive (Cavorandum peeler),
after which they were trimmed. The peeled potatoes were cut

into 2 4 mm and 6 mm slices in a Lipps kitchen machine

34
35

(type: C - IRB, model: No. 360101.3, made by URDORF-ZURICH).


The slices were thoroughly washed with tap water, then dipped
In a 0.5% solution of Sodium Metabisulphite (NaHS03) for 5
minutes. These steps are illustrated in figure 3.1 below.

Figure 3.1 Flow sheet for dehydration of raw potato slices

Drying of potato pieces is influenced by drying air

temperature, thickness of the slices, air mass flow rate and

humidity. in this work, the effect of air temperature,

thickness of slices and air flow rates on the drying behaviour

Was investigated. The following were measured and recorded

progressively as a function of the residence time:-

(a) Ascorbic acid - to check on the relative loss of

nutrients.

(b) Browning (heat damage) of the dried slices.

UNIVERSITY OF NAIROBI LIBRARY


36

The ascorbic acid content, visible browning and

rehydration property of the dried product were used as criteria

for evaluation and determination of the appropriate drying

conditions. *

3.3 Drying of the Slices

The slices were loaded onto the tray of a Memmert cabinet

hot-air oven made by 854 Schwabach, W-Germany (model: Schutzart

DIN 40050-IP20, type: UL 3017). The rated capacity of the oven

is 1.3 kw. The air is electrically heated before passing

through the slotted tray holding the slices. The equipment has

provisions for the variation of drying air temperatures within

the range 30-220°C and airflow rates within the range 0-0.0108

kg/m?s.

The slices were dried on a 0.117 m2 rectangular tray (300

mm by 390 mm) giving the following loading rates for a single

layer of the different slice thicknesses:-

6 mm : 5.43 - 5.85 kg/m2

4 mm : 4.40 - 4.49 kg/m2

2 mm : 2.39 - 2.48 kg/m2

Small amounts of potato slices were dried in each case, so that

the relative humidities are constant at different values for

a range of drying air temperatures between 50°C and 70°C, both

temperatures inclusive.

The air temperature entering and leaving the thin layer


^ g
Measured by means of two sets ofcopper-constantant

^tfflocouple connected to a 10-point recorder with the

u*ated cold junctions, placed in a flask of crushed ice.


37

The equipment, accurate to 0.01 of a degree, gives readings in

millivolts directly convertible to degrees celsius using a

given chart. The wet bulb readings were obtained by covering

the thermocouple junctions with a wick dipped in distilled

water in a glass bottle. Three different thermostatically

controlled dry-bulb temperatures (50 ± 1 , 60 ± 1, and 70 ± 1

°C) were studied. The mean ambient temperatures (wet bulb and

dry bulb) during the experiments were measured enabling the

relative humidity of the hot air to be calculated.

Effectively, this was a 3 by 3 by 2 factorial experiment.

The drying air temperature and slice thickness had three levels

each whereas the air velocity had two levels. Hence, this

yielded eighteen treatment combinations. Each treatment

combination was replicated three times giving a total of fifty-

four test runs which were assigned to experimental units (0.117

nr slices) randomly. The drying air temperature levels were

508C, 60°C and 70°C, the slice thickness levels were 2 mm, 1

mm and 6 mm. The airflow rates were 0.0054 and 0.0108 Kg/m2s.

The relative humidity of the drying air was not included as s

factor because it is a variable that is not normally controlled

in dryer design. The samples were withdrawn at different

stages from the oven to give a series of temperature-time

treatments. After completion of experiment, 2 samples for

moisture content determination were taken from pre-selected

Positions on the surface of the tray.

The distribution of air over the cross section of the

^nlet of the drying space determines the uniformity of drying

the plane perpendicular to the direction of air flow.,



38

Airflow up through the tray was used in this series of

experiments. The tray drying arrangement successfully exposes

a large area of drying surface per unit of total space

occupied. In a well-conducted operation, careful attention is

given to maintaining a uniform spread of the cut material on

the trays and uniform and predetermined total weight on the

tray.

3.4 Analytical Methods

Initially and after every 30 minutes during drying the

potato slices were analyzed for moisture content, ascorbic acid

content, visible browning and rehydration property. All these

determinations were done in duplicates.


s

3.4.1 Determination of Moisture Content

Ten grammes of raw slices, of partially or completely

dried potato slices were dried in an air oven set at 70 ± 2*C

to constant weight. The loss in weight was calculated as

percent moisture content of the sample. The weighing balance

used was accurate to 0.01 of a gram.

3.4.2 Determination of Ascorbic Acid (vitamin C)

Reduced ascorbic acid (RAA) was determined by the method

°f Barakat et al., (1955). For the fresh potato samples, 50g

°f slices were blended with 150 ml of 10% of trichloroacetic

a°id solution (CCI3.COOH), (TCA). For the dried slices, lOg

were blended with 50 ml TCA solution till homogeneous. The

8lurry Was filtered through a Whatman No. 91 filter paper. To


* ♦
39

5 ml of the filtrate was added 5 ml Potassium Iodide (Kl!)

solution and 2 ml starch. Ascorbic acid was determined in the

5 ml filtrate by titrating with a standard solution of M-

Bromosuccinimide (C4H4BrN02) , (0.01% solution) to a faint

violet/blue end point. All the chemicals used were obtained

from Kobian (Kenya) Ltd.

3.4.3 Measurement of Browning Discoloration (Heat Damage)

The extent of visible browning of the slices was evaluated

by the method of Hendel et al., (1950) . lOg of the product were

extracted with 50 per cent (v/v) ethanol, and the absorbance

of the solution read at 420 nm on a Beckman model 25

spectrophotometer.
0

3.4.4 Determination of Rehydration property

Duplicate samples of lOg of partially or completely dried

slices were weighed into 600 ml beakers and covered with 300

ml distilled water. It was then covered with watch glass and

boiled for 25 minutes using a hot-plate electric heater, made

by Gerhardt Bonn (type: EV 16, power rating: 2.7 k w ) . This is

the method of Sullivan et al.,(1980). After draining, the mass

°f water absorbed by one gram of dry solids was calculated..

This was expressed as a ratio and referred to as rehydration

ratio.

3 •<. 5 Development of Drying Equations

Various types of thin layer models have been proposed for

deacribing the drying of agricultural products in the falling


rate on­
stage. They can be classified as: (a) semi-empirical, (b)

40

empirical, and (c) theoretical. Each class of model possesses

certain advantages and disadvantages.

The empirical class of models is usually the simplest t:o

use. Their'major disadvantage, however, is that there is no

basis in theory for extending their use beyond the range of the

drying data. Thus the use of such a model is restricted to

conditions covered by the existing data for the product being

considered. This severely limits the usefulness of these

models for products like potato slices which possess no fixed

geometry and for which little or no drying data exist.

Semi-empirical models are generally developed from theory

but include empirical parameters which reguire data for

evaluation. These models (because of their theoretical base)

can 6e more easily extended beyond the range of the data from

which they were evaluated, provided the extension is not great

and the empirical parameters are well behaved; i.e., either

constant or reasonably simple functions of relatively few

independent variables. Unfortunately, the most widely used

semi-empirical model (the exponential model) contains a drying

"constant" which is not only a function of a large number of

variables but of the interactions between these variables as


well.

Theoretical drying models are generally limited in their

usefulness owing to their complexity. Quite often they employ

Variables which are difficult to measure in relationships that

are difficult to solve. The most widely used theoretical

m°dels are based on the diffusion equations. The major

^foitation involved in applying»* these models to the drying of


* ♦
41

potato slices is that temperature gradients within the material

are normally neglected in order to simplify solution of the

equations. Since temperature gradients do exist within porous

materials during drying, use of a diffusion model would require

the consideration of the coupled effects of simultaneous

temperature and concentration gradients. Coupled diffusion

equations of this type are quite complex and do not lend

themselves to easy solution. Therefore, such models were

considered to be inappropriate for the purposes of this study.


4. RESULTS AND DISCUSSION

4.1 Moisture Reduction

The experimental data presented in table 4.1 below was


extracted from graphical plots of moisture ratio (MR)
versus residence•time (minutes) for each drying condition
studied. Rl, R2 and R3 in the table represent the values
for first, second and third replications respectively.

Table 4.1: The experimental data on time of half response


for all drying conditions studied

Drying Slice Air Time of half response


Air Thick flow (minutes)
Temp. ness rate:
(°C) (mm) Rl R2 R3 Mean
kg/m2s
50 2 0.0054 54.5 61.8 54.6 57.0
50 2 0.0108 66.2 67.5 70.0 67.9
50 4 0.0054 133.3 110.0 137.8 127.0
50 4 0.0108 110.0 100.0 131.1 113.7
50 6 0.0054 180.0 182.8 185.7 182.8
50 6 0.0108 180.2 183.0 194.3 185.8
60 2 0.0054 37.3 37.7 38.2 37.7
60 2 0.0108 37.7 45.9 39.1 40.9
60 4 0.0054 91.7 94.2 95.0 9 3 .6
60 4 0.0108 83.3 86.7 81.7 83.9
60 6 0.0054 151.4 142.9 148.6 147.6
60 6 0.0108 150.0 130.9 152.7 144.5
70 2 0.0054 44.4 40.0 46.9 43.8
70 2 0.0108 46.3 46.5 43.8 45.5
70 4 0.0054 73.8 70.0 74.0 72.6
70 4 0.0108 76.3 70.0 80.0 75.4
70 6 0.0054 124.4 110.0 130.0 121.5
70 6 0.0108 116.0 104.0 120.0 113.3

These data were employed to determine the general


effect of the three factors; the drying air temperature,

slice thickness, and air flow rate on the thin-layer drying


behaviour. The dependent variable in this analysis was the
time of half response. The mean values of this variable,

♦42
43

w ith their standard deviation are shown in table 4.2. The


of half response is defined as the time required to
remove the first-half of free moisture or 1/2(M0-Me); and
consequently it corresponds to the time required by a thin-

layer particle drying process to reach a moisture ratio

(MR) of °-5 *

Table 4.2: The effect of drying conditions on the time of


half response (minutes)

Drying Air Slice Air flow Time of half-


Temperature Thickness rate response
(°C) (mm) (kg/m2s) (minutes)*

(Mean + s d ) .
50 2 0.0054 57.0 ± 3.4
50 2 0.0108 67.9 ± 1.6
50 4 0.0054 127.0 ± 12.2
*50 4 0.0108 113.7 ± 13.0
50 6 0.0054 182.8 ± 2.3
50 6 0.0108 185.8 ± 6.1
60 2 0.0054 37.7 ± 0.4
60 2 0.0108 40.9 ± 3.6
60 4 0.0054 93.6 ± 1.4
60 4 0.0108 83.9 ± 2.1
60 6 0.0054 147.6 ± 3.5
60 6 0.0108 144.5 ± 9.7
70 2 0.0054 43.8 ± 2.9
70 2 0.0108 45.5 ± 1.2
70 4 0.0054 72.6 ± 1.8
70 4 0.0108 75.4 ± 4.1
70 6 0.0054 121.5 ± 8.4
70 6 0.0108 113.3 ± 6.8 -

* Each value is a mean of three replications for a thin

layer of slices on 0.117 m2 tray area.

The results displayed in table 4.2 are further

Presented in table 4.3 in a imanner that lends itself well



44

to statistical analysis with a view to establishing the

general effect of the independent variables of the study

(appendix A ) .

Table 4.3: Time (minutes)* of half response (MR = 0.5)


for the drying at three temperature levels A, for three
slice thicknesses B, and two airflow rates, C.

Tetnp«reture A ir flow Slice thickness Total ■ b,


■ A rate « C • B ♦ b2 ♦ b3
2 mn* * b, A f r m=b2 6 fTfn3 b ^

50 ■ a , 0.0054*c, 170.9 381.1 548.5 1100.5


0 .0 10 8 *c2 203.7 341.1 557.5 1102.3
Total 3 74.6 722.2 110 6.0 2202.8

60 ■ a 2 0.0054 113.2 280.9 442.9 837.0


0.0108 122.7 25 1.7 433.6 808.0
Total 235.9 532.6 876.5 1645.0

70 • a3 0.0054 131.3 217.8 364.4 713.5


0.0108 136.6 226.3 340.0 702.9
Total 267.9 444.1 704.4 1416.4

Total ■ 0.0054 415.4 879.8 1355.8 2651


• \ 4 *2 ♦ 0.0108 463.0 819.1 1331.1 2613.2
Total 878.4 1698.9 2686.9 5264.2
®3 * G

* Each value is a total of three replications for a thin


layer of slices on a 0.117 m 2 tray area.

Table 4.3 is necessary for the computation of the sums


of squares of the main effects and interactions.

Statistical analysis of the data presented in table 4.3,


based on the analysis of variance (ANOVA), is summarised
in table 4.4.

The results of this factorial experiment lend


themselves to a relatively simple explanation because of
the variety and nature of treatment comparisons. The ANOVA
of great value since this study is exploratory and
iittle is known concerning the optimum levels of the

45

factors or even which ones are important for mathematical

modelling of the drying process.


Table 4.4: ANOVA of data in table 4.3

Source ’ df ss MS Fb

Replications r - 1 - 2 541.70 270.85

A * Temperature a - 1 - 2 18181.92 9090.96 221.19**

B » Slice thickness b - 1 - 2 91111.79 45555.90 1108.42**

C ■ Airflow rate c - 1 « 1 26.46 26.46 0 . 6 4 n*

AB (a - 1 )( b - 1 ) - 4 836.85 959.21 23. 3 4 * *

AC (a-1)(c-1) - 2 26.69 13.34 0.32"’

BC (b-1)(c-1) - 2 338.00 169.00 4.11*

ABC (a-1)(b-1)(c-1) - 4 355.83 88.96 2 . 1 6 n‘

Error (r - 1 ) ( a b c - 1 ) - 3 4 1397.59 41.10

Total abcr-1 ** 53 115816.83

b ** = significant at 1% level, * significant at 5% level,


6

ns = not significant.

For the data on the time of half-response, analysis

of variance shows that two main effects (A and B) and two

interactions (AB and BC) are significant. The significant

BC interaction implies that the differences between

responses to C vary with the level of B, where responses

are measured over all levels of A. Alternatively,' the

differences among responses to levels of B vary for the two

levels of C, where responses are again measured as totals

or means over all levels of A (Steel and Torrie, 1980) .

Specifically, the differences in time of half response:

(minutes) when averaged over all drying air temperatures,

between the two airflow rates are not the same for the

three slice thicknesses; or*the differences among times of


46

half response for a thin layer of potato slices of the

three different slice thicknesses are not the same for the

lower as for the higher airflow rates. Therefore, whereas

the main effect C is nonsignificant, its contribution to

the drying process is observed in its interaction with the

main effect B.

Considering that a single mathematical model

describing acceptably the behaviour of a system for all

cases is more valuable than several models describing

accurately certain conditions only, the ANOVA was used to

establish the experimental factors to be incorporated in

a general equation describing the drying behaviour of

potato slices, developed later in this chapter. If the

'factors were largely independent, the table of treatment

means and ANOVA summarize the data well. Since some factors

are not independent, further tests are necessary for a

detailed study.

Since the AB and BC interactions are significant, it

seems logical to begin by examining the BC interaction

followed by the AB interaction. Factor C is at two levels

only, therefore the test begins by considering the simple

effects of C at the various levels of B. Apparently, the

simple effects of C are not homogeneous for the three slice

thicknesses. The two-way table contained in Table 4.3, and

used in computing the BC interaction is hereby examined.

This is presented as Table 4.4a where the responses to air

flow rate are compared for the various slice thickness. The

difference 60.7 stands out immediately.



47

Table 4.4a: Examination of the BC interaction for the data


of Table 4.3

Airflow rate
c2

0
h*
fo
ci
Slice
thickness: >
bl 415.4 463.0 47.6

<Ti
-0
i
b2 879.8 819.1

O

b. 1355.8 1331.1 -24.7
b i + b, + b-, 2651 2613.2

C within b, SS* (463.0-415.4)a _125 8B,


1 2ra 2(3)3

C within b, SS= H t ic , ) - ( V , ) ] ’ = (819.1-879.S ) ^ 204 6 9 .


2 2ra 2(3)3

#C within b3 SS= -■— Cl-— 1355.8) 2 =33 gn9


3 2ra 2(3)3

Table 4.4a shows that the difference in time of half­


response for potato slices at level c : and c2 of air flow
rates, averaged over the three drying air temperatures, is
not significant for 2 mm slices or for 6 mm slice thickness
but is for 4 mm slice thickness.
Because of the problems of interpreting significant
interactions in the analysis of variance, it was found
prudent to examine certain simple effects at the expense
of a loss of replication; each of the simple effects of C

is measured at only a single level of B.

The significant AB interaction indicates that the


difference in time of half-response between the drying air-
temperatures, when measured over the two levels of air flow
fate, differs for the three slice thickness. A further
examination of the simple effects are shown in Table 4.4b

48

Table 4.4b: Examination of the AB interaction for the delta


of Table 4.3

Slice
thickness
Temperature: bi b.i b3 " b l
al 374.6 1106.0 731.4
a2 235.9 876.5 640.6

a3 267.9 704.4 436.5


3 , 4- 3, + a-, 878.4 2686.9

_ [(a3b3) - ( a ^ ) ]2 _ ( H06-374,6 )2 _
B within a3 SS =44578.8'
2 rc 2 (3)2

5£._ [ ( a ^ - t a , ^ ) ] 2 (876.5-235.9) 2 _
B wi thin a, =34197.4'
2 rc 2 (3)2

t (a3b3) - ( a ^ ) ]2 (704 .4-267.9 )2


B within a3 SS= = 15877 .7**
2 rc 2(3)2

The above information presented by individual sums of


squares for simple effects shows that the differences in
time of half-response are not homogeneous. These F-tests
and those shown in Table 4.4a are essentially least

significant difference comparisons but made after the


significant interaction showed evidence of real differences

among simple effects. They are result-guided as a


consequence of a test of significance. Severe criticism of
these tests would seem unjustified.
4.1.1 Effect of Drying Air Temperature on Course of Drying
of Potato Slice

The effect of air temperature on the drying time of

potato slices was investigated and figure 4.1 illustrate

the results. The dry bulb temperature was varied in the


range 50-70°C for all the runs. As is to be expected, the
increase in temperature reduces the drying time, but the


49

choice of optimum working temperature must be done


considering the effect of temperature on the physico­
chemical characteristics of the dried product.
The temperature, humidity and velocity of the air
being used for drying are the physical properties of the
drying environment which have an effect on the rate of
drying and on the economics of the process. The combined
•r

effect of humidity and temperature of air is determined by

the psychrometric relationship (appendix B) and is obtained

— 50 degrees celcius 60 degrees celcius <


70 degrees celcius

A n H o w ral e ( k g l r r A ) 0 0 0 5 4
Sl i c e T h i c k n e s s |m m | 4

Figure 4.1 Effect of Air Temperature on Course of Drying


of Potato Slice

by measuring the wet bulb temperature (adiabatic saturation


temperature for water). The wet bulb temperature is a
unique function of dry bulb temperature and humidity. The
drying rate has been shown to be proportional to the wet
bulb depression for a number of vegetables (Ede and Hales,
* ♦
50

1948) . However, when the relative humidity of the air is


less than about 40% the rate of drying is independent of
wet bulb depression (Rumsey, 1985).
Figure 4.1 shows that the dry-bulb temperature of the
f
air has considerable effect on the course of drying at low

moisture levels but n o t 'at high. At the low moisture

levels the cooling effect of evaporation is small and

consequently the heat is utilized in the internal


redistribution of moisture. The limiting temperatures are

determined by the biochemical changes which will cause the


development of off-flavours and discoloration. The
combined effect of temperature, humidity and velocity of
the air have a predominant effect on the rate of drying.
4.1.2 Effect of Thickness on Course of Drying of Potato
Slice
» Figure 4.2 shows the effect of thickness on the
residual moisture content of potato slice.

Time, mm

— 2 mm — 4 mm — 6 min

Ah temperaluie |®C|: 50
Aiillow rale (kg/m*s): 0 0054

figure 4.2: Effect of potato £lice thickness on drying curves


51

These typical drying curves show that the drying time: is

considerably affected by the slice thickness.

Piece size has a tremendous effect on the rate of


drying. The thicker the piece, the slower the drying rate.
This is a limitation in expanding the use of dehydrated

foods and especially in producing products for


remanufacturing uses, such as in stews, salads, and other

recipes that use a relatively large piece. For conveyor

driers, an increase in thickness of piece from 3 mm to 13

mm will reduce drier capacity by 90% (Van Arsdel, 1973b).


4.1.3 Drying Rate as a Function of Sample Thickness
Eguation (2.16) applies only in the cases in which the
surface resistance to mass transfer is negligible as
compared to the resistance to internal diffusion of liquid,
end the diffusivity is constant.

The experimental results fulfil the condition of


uniform moisture distribution in the slices at the start
for two of the three dry-bulb temperatures studied since
the whole drying process of the material occurs during the

falling rate period. At 50°C a short constant rate period


seems to exist, but it is reasonable to assume that for

practical purposes this period does not play an important


role and that mathematical models can be developed assuming
that drying for this system occurs in the falling rate
period.

The approach presented in equation (2.16) for large


drying times requires that the drying rate of the material
vary inversely as the square of the thickness. The results
fulfil this condition as shown in Fig.4.3, where the time
required to reduce the average fraction of evaporable water
from level 1.0 to level »*0.1 is found to be directly
* ♦
52

proportional to the square of the thickness.

It is considered that for an overall analysis, of


drying behaviour a series of assumptions can be made to
facilitate the developing of mathematical models . In this
work, there was no correction for shrinkage. Measurements
of the changes of slice thickness as a function of time was

difficult to carry out because of the small magnitude of


these changes.
TIME (min) from MR=1 to MR= 0.1

50°C, 0.0054 kg/m2s 50°C, 0.0108 kg/m2s Jfr_60°C, 0.0054 kg/m2s


60°C, 0.0108 kg/m2s 70°C, 0.0054 kg/m2s 70°C, 0.0108 kg/m2s

figure 4.3: Time from MR=1 to MR=0.1 as a function of the


square of sample thickness

Experimental measurements of the drying rate are


Usually preferred over using equations for prediction.
* ♦
53

However, the equations are quite helpful to predict the

effect of changing the drying process variables when


limited experimental data are available. The rate of
drying, R, is defined as follows;

La dM (4.1)
R = -
A dO

where M = free moisture content (kg of free water/kg of dry


solid)

L b = kg of dry solid used, and


A = exposed surface area for drying in m 2.

When R is a linear function of M in the falling rate


period,

R = aM , (4.2)
0

where a is a constant. Equating the two expressions for R,

£, dM (4.3)
R aM
A dQ

Rearranging,

dM aA
M (4.4)
dt

When liquid diffusion of moisture controls the rate of


drying in the falling-rate period, Fick's second law for
unsteady-state diffusion can be used.

(4.5)
* v dx2

where Dv is the liquid diffusion coefficient in m 2/hr and


x is distance in the solid in m.

During diffusion-type^drying, the resistance to mass


54

transfer of water vapour from the surface is usually very

small, and the diffusion in the potato slice controls the


rate of drying. Then the moisture content at the surface
is at the equilibrium value M0 . This means that the free
moisture content at the surface is essentially zero.

Assuming that the initial moisture distribution is


uniform at 0 = 0, Fick's second law for unsteady state

diffusion may be integrated to give equation (2.16).

- M _ 8 r -Dv6(n/2X1)1+ _1 -90v6(n/2x,)1+ _l_ -25Dvfl(D/2x1)J+ ,.,


M0-M, Mx D 2 9 25 ■( > >

where M = average free moisture content at time, 0 hrs;


Mj = initial free moisture content at 0 = 0,
Me = equilibrium free moisture content,

Xj = L = 1/2 the thickness of the slab when drying


0
occurs from the top and the bottom parallel
faces, and
Xj = L = total thickness of slab if drying only from
the top face.
This equation assumes that Dv is constant, but Dv is rareLy
constant; it varies with moisture content, temperature, and
humidity. For long drying times, only the first term in the
equation is significant, and the equation becomes

Ji ■ JL in/axji (4.7)
nj

Solving for the time of drying,

e . i s ! In 6«l (4.8)
n !z>„ n=M

In this equation if the diffusion mechanism starts cit



55

M = Mc (the critical free moisture content) then M L = M c.

Differentiating this equation with respect to time and


rearranging,

dM m -n2£>v M (4,9)
dQ 4;qJ
i

Multiplying both sides by -L0/A,

_ dM _ (4.10)
A d6 4x x2A

Hence, these last two equations state that when


internal diffusion controls for long times, the rate of
drying is directly proportional to the free moisture M and
the liquid diffusivity and that the rate of drying is
inversely proportional to the thickness squared. Or, stated
0
as the time of drying between fixed limits, the time varies
directly as the square of the thickness. This analysis is
proven in this study.
4.1.4 Drying Equations for Potato Slices

Drying curves normally plot exponentially as moisture


content drops with time. The linearisation of the

dimensional parameters of MR and X in the Crank diffusion


equation is needed in order to obtain a best fit curve.
The method of successive residuals is a standard
statistical technique that considers a decay curve as a
series of small lines and so the general equation can be

presented in the form of a Fourier series. When it is


aPplied to drying curves, the number of terms of the

eguation obtained depends on the drying behaviour of the


Product.

The change in moisture ratio (MR) is generally


represented by an exponential decay equation as:-
56

MR - ^2 AsExp (-kjLt) (4.11)

where
t'
t = time, hrs
= dimensionless constants, and
= a drying rate parameter.

Mohsenin (1980) cites the application of the method


to establish the stress and force relaxations in various

agricultural products.

In general, the relaxation equation is expressed as


shown in equation 4.12 below:

°c = °e + E 0ne(_t:/Tn) (4.12)

Where
ae = Stress at infinite time in N/m2
n = 1 to oo

Note the similarity between equations (4.11) and (4.12).


For any specific problem it may be desired to
eliminate the long-time or the short-time variations and
consider the value of time-constant which describes the
behaviour over a range of time important to the problem.

In the case of the potato slices, the long-time constant


corresponding to the falling rate period is sufficient for
practical applications of the equations thus obtained.

Figures 4.4-4.6 show the method of successive


residuals for the analysis of the moisture reduction data

at drying air temperatures of 50°C, 60°C, and 70°C.

The straight line segment of the original curve of log


MR against time (the first exponential) is extended to the
ordinate axis and its equation is obtained by slope and
* ♦
57

intercept method (line 1). The intercept gives the

coefficient of the first exponential term defining this


segment of the curve. Where more than one term is
required
»
to describe the drying curve, the difference
between this straight line and the original curve is

plotted on the same semi-log paper to obtain the first

residual curve. This procedure can be repeated several

times until the true drying curve is represented by a


sufficient number of terms.

0 30 60 90 120 150 180 210 240


Time(mm)
AVGMR Line 1

Potato Slice Thickness|m m ) ?

figure 4.4 Development of drying equation for potato slices


at 50°c air temperature and air flow rate of 0.0054 kg/m2s.

For the calculation of time constant Tc from the slope

°f the straight line segment, equation (4.13) can be used.

T = ---- — ---— ---(4.13)


c In MR, - In MR2

50

Time (minutes)
AVG.MR Line I

Polalo Slice Thickne$s(mm| 6

Figure 4.5 Development of drying equation for potato slices


at 60°C air temperature and air flow rate of 0.0054 kg/m2s.

Equation (4.13) is based on the fact that time constant Tc


in an exponential term is the inverse of the slope of the

straight line segment representing that particular term.


When the scale factor in the semi-log paper is taken into

consideration, and using Eq.(4.13), the longest time


constant is found to be

________ (240 - 0)________


240.36 minutes.
2 . 3 (log 0.02 - log 0.007)

Since the ordinate intercept is at MR = 0.02 the first


e*ponential term for the straight line segment of the
0riginal curve is
* ♦
59

MRX a 0.02 e't/240-3 (4 .14 )

Similarly, the equation for the second exponential term is

found by calculating T2 and reading the intercept for the


t
straight line segment of the second curve, which is
obtained by difference. ;

____________ ( 120 - 0 )____________


T3 = = 20.02 minutes.
In (0.800) - In (0.002)

Hence,

MR2 = 0.8 e ‘t/2° (4.15)

Repeating this procedure once more yields

T u ________ (30 - 0)________


=23.4 minutes.
3 In (0.018) - In (0.005)

Hfence

(4.16)
MR3 = 0.018 e't/23,4

The complete drying curve can be described by a three-term


equation

(4.17)
MR = 0.02e"t/240,3 + 0.8e't/20 + 0.018e't/234

For Figure 4.5, the longest time constant, T x is given by

_____ (360 - 160)


Ti = 166 minutes.
In (0.1) - In (0.030)

hence,

MRX * 0.26 (4.18)

For the second exponential term, T2 is given by


______ (180 - 60)______
= 173 minutes.
In (0.08) - In (0.04)
60

Time (minutes)
~B_ AVGMR Lme I

Potato Slice Thickness|rnm| ?

Figure 4.6 Development of drying equation for potato slices


at 70°C air temperature and air flow rate of 0.0054 kg/m2s.

hence,

MR2 = 0.12 e't/rn (4 -19)

Repeating this procedure once more yields,

< a _______ (49 - 0) 120.8 minutes.


3 In (0.03) - In (0.02)

Therefore,

(4.20)
=0.03 e" 5-8

Hence, the drying curve is described by the following


three-term equation «

61

MR = 0.26 e't/16S +0.12 e"t/173 + 0.03 e't/120-a

Similarly, for figure 4.6, the longest time constant, is


given by

________(12 0 ,t 0)_______ =99.7 minutes.


In (0.1) - In (0.03)

hence,

MR1 = 0 . 1 e ‘t/99-7

For the second exponential term, time constant T2 is given


by

T ______ (50 - 30)______


12 In (0.06) - In (0.04)
=49.3 minutes.

hence,

MR2 = 0.13 e't/49,3

Repeating this procedure once more yields,

(30 - 0)
114.3 minutes.
In (0.013) - In (0.01)

Thus,

(4.24)
MR3 = 0.013 e't/114-3

Therefore, the drying curve eguation becomes

MR = o.l e *t/99-7 + 0.13 e't/49-3 + 0.013 e -t/U4-3 (4 ’2 5 J

In these equations t represents the residence time in


minutes.

The drying curves show that the drying behaviour of

the potato slices can be described by a three-term equation


i
* ♦
62

corresponding to three stages of drying. Gorling (1958)


observed that when he made allowance for sample shrinkage
in the one-sided drying of a potato slice, the dehydration
occurred
r
in three stages which included no constant-rate
drying (Akeredolu, 1987).
.v
Equilibrium moisture content (Me) in MR is an

important parameter in developing mathematical expressions

for the drying process. In evaluating this parameter


experimentally, one makes the distinction between static
and dynamic equilibrium moisture contents. The former is
obtained by exposing the material to a constant volume of
air, at known temperature and humidity for a considerable

duration. The latter is achieved by passing air of known


temperature and humidity through a sample of the material
#until its weight remains constant. It therefore appears
that static equilibrium moisture content is more suited for
use in storage problems whereas the dynamic equilibrium
moisture content should be applied to drying experiments.
It is the drying air temperature that brings about the

variation in relative humidity. Equilibrium moisture


content is predominantly a function of relative humidity,
with temperature having a secondary effect. At drying air
temperatures of 50°C, 60°C and 70°C for the potato variety
(Desiree) studied, the dynamic equilibrium moisture content
values obtained are shown later (Table 4.8 ). During the

tests, drying of triplicate samples was carried out in the

drier for several days until there was no change in

moisture content for three consecutive daily readings


(appendix C ) .

Mohsenin (1980) cites the application of the method


°f successive residuals to i establish the stress relaxation
* ♦
63

in compressed alfalfa, force relaxation in pea beans, and

stress relaxation of potato tuber. Nindo (1991) used the


method to describe the drying behaviour of maize grains.

4.1.5 Regression analysis for the drying equations

Figures 4.7-4.9 show the regression analysis output


for the drying equations developed in section 4.1.4.
Curve Fitted Values, MR

Figure 4.7 Regression analysis for figure 4.4


64

^Figure 4.8 R e g r e s s io n a n a l y s i s o u t p u t f o r f i g u r e 4 .5
Curve Fitted Values, MR

Experimental Values, MR (dimensionless)

Figure 4.9 R e g r e s s io n a n a l y s i s o u t p u t f o r f i g u r e 4 .6
65

The equations developed above are regression equations


for specific drying conditions. Therefore, there is need
to develop an equation that can predict drying behaviour

of potato slices at any drying temperature and relative


i

humidity, given any potato slice thickness. Table 4.5


<C
below presents the experimental data for the development
of such an equation. A mathematical expression of moisture
ratio (MR) as a function of the three independent variables
in the table (X3, X2, and X3) is required.

Table 4.5 Experimental data for development of drying


equations

Experiment Drying Drying Potato Air flow


number air air slice rate
• temperat relative thickness (kg/m2s)
ure humidity (mm), X3
(°C), X, (%), X,
1 50 13 2 0.0054
2 50 16 2 0.0108
3 50 20 4 0.0054
4 50 22 4 0.0108
5 50 22 6 0.0054
6 50 25 6 0.0108
7 60 16 2 0.0054
8 60 15 2 0.0108
9 60 16 4 0.0054
10 60 15 4 0.0108
11 60 13 6 0.0054
12 60 14 6 0.0108
13 70 11 2 0.0054'
14 70 9 2 0.0108
15 70 8 4 0.0054
16 70 8 4 0.0108
17 70 8 6 0.0054
18 70 8 6 0.0108
66

There have been many attempts, with varying degrees


of success, to develop mathematical models that describe

the drying characteristics of biological materials. The

most commonly used model is an empirical equation analogous


to Newton's law of cooling (eqn. 2.20). When A0 = 1, this
p
relationship is often referred to as the thin-layer drying
equation and is

M ~M > _ e -Kt , (4.26)


Mo~M.

where
M - Moisture content (db) at any time, t
M0 - Uniform initial moisture content, db
Me - Equilibrium moisture content, db
0
K - Thin-layer drying constant, hr-1
t - Time, hr

The two constants that must be evaluated before


practical use can be made of the equation are K and Me .
Henderson and Pabis (1961) have proposed the use of the

dynamic equilibrium moisture content for the latter in


preference to the static equilibrium moisture content.

Drying air temperature is usually considered to be the only

factor that significantly affects the drying constant K.


Relative humidity and air velocity in the usual range of
drying are not considered significant.
Henderson and Pabis (1961) have suggested the use of
an Arrhenius type equation to describe the variation of K
with temperature;


67

K = a exp (- (4.27)
T

where ,
v
a,b - Material constants,
s

T - Absolute temperature of drying air, K.


Equations [2.20] and [4.27] were applied to the drying of
potato slices to determine basic drying characteristics.

The effects of temperature and slice thickness on the

value of K are shown graphically in Fig. 4.10 and 4.11.

Each point shown on the graphs is the average of three


replications. Fig. 4.10 is a semi-logarithmic plot of K
versus the reciprocal of absolute drying air temperature.
A linear regression equation of the Arrhenius type was
^fitted to the values of K. Table 4.6 gives these equations
and the corresponding parameters of interest. Statistical

analysis showed that the regression coefficients were


significantly different from zero with the indicated r2
values.

The variation of thin layer drying constants with


slice thickness is shown in Fig. 4.11. Exponential
equations were fitted to the data by linear regression.

These equations are given in Table 4.7. Statistical


analysis showed that the regression coefficients were
significantly different from zero with the indicated r2
values.
o
u
CT>

0.001'--------------------1--------------------L_
291 300 309

Reciprocal of an temp., Ik I E -5

2 mm - 4 - 4 mm 6 mm

figure 4 . 1 0 K-values for various thicknesses of potato


slices dried at different air temperatures.
69

Tabla 4.6. Regression Equations and Parameters for


Fig.4.10.

Size (mm) Regression equation* r2


2 In K = 1.208 - 1605/T 0.666
4 In K = 2.546 - 2274/T 0.787
6 In K = 3.187 - 2662/T 0.820

*K - Thin layer drying constant, min-1


T - Temperature, K.

figure 4.11 K-values for potato slices dried at various air


temperatures as affected by slice thickness.


70

Table 4.7 Regression Equations and Parameters for Fig. 4.1.1

Temperature, (°C) Regression equation r2


50 In K = 3.207 - 0.311d* 0.969
60 In K = 2.913 - 0.323d 0.962
70 . In K = 3.008 - 0.262d 0.980

*d - Slice thickness, 2L (mm)

The information in Table 4.6 was used to establish a


mathematical relationship of K in terms of the slice
thickness, d and the drying air absolute temperature, r.
By data fitting using Grapher computer software package,
the following relationship was found to give the best fit.

# (1 .81525L n (tf ) -0.0287 331) - 2:45 L /i (d) *.


? 38'3871 /. j n \
Kx = e T (4.28)

Similarly, for Table 4.7, data fitting using piecewise-

polynomial approximation (Conte and de Boor, 1980) with


Grapher plotting package gave the best fit equation (4.29)
shown.

e <221.9 0 5 -1 .3 0 5 3 2 7 * 0 .0 0 1 9 4 5 7 *) - ( 0 . 2« 06 4 T - 3 9 . 3 3 5 7 - 0 . 0 0 0 3 6 S T 1 ) ei
*2

(4.29)
71

Since equation (4.29) is derived from regression


equations with a good fit (r2 of 96 to 98%) , k2 is

considered a sufficient estimate of the thin-layer drying


constant, K. K x in equation (4.28) is also an estimate of

K based on regression equation data with r2 of 67 to 82%.


>
It gives a more accurate estimate than K2 and is therefore
preferred to K2 .

The experimentally determined dynamic equilibrium


moisture content, Me , are shown in Table 4.8. The

following linear regression equation was fitted to the


data;

Me = 0.055 - 0.00075TC ,, or

M g = 0.25986 - 0.00075T (4 -30)


i

Equation (4.30) was obtained for a relative humidity range


of 8% to 26%.

where

Tc, - Temperature of drying air, °C


T - Absolute temperature, K

Me ~ Dynamic equilibrium moisture content, db.


The r2 value was found to be 0.727, indicating that

approximately 73% of the effect of drying air temperature


°n Me can be explained by regression.
72

Tabla 4.8: Experimentally determined dynamic equilibrium


moisture content, Me

Drying conditions Equilibrium


moisture content
t (db)
Air Temperature Slice
(°C) thickness
(mm)
50 6 0.017 ± 0.004
50 4 0.024 + 0.003
50 2 0.011 ± 0.004
60 6 0.014 ± 0.001
60 4 0.006 + 0.001
60 2 0.011 + 0.002
70 6 0.004 + 0.001
70 4 0.003 + 0.002
70 2 0.000 ± 0.0001

*
73

4.2 Ascorbic acid retention

i'
The loss of nutritional quality during drying and

storage of foods has become an increasingly important

problem. The loss of some nutrients, including ascorbic


acid, may actually become the limiting factor in
determining the shelf life of some products. There is cin
obvious need for methods which can be used to predict the
degree of nutrient loss occurring in specific food
products. One such method is through the application of
kinetic analysis to nutrient retention data.

The majority of data concerning the degradation cf


ascorbic acid in food is the result of endpoint analyses;
that is, the measurement of ascorbic acid concentration
prior to and at the conclusion of a given process.
Nutrient retention data gathered in this manner are

generally not subject to kinetic analysis. Existing

kinetic studies of ascorbic acid degradation in food are

limited in one or more other respects. First, these


studies have been performed at relatively low temperature,

characteristic of product storage rather than processing.


Secondly, predictions of nutrient stability based on

kinetic analysis have been tested under conditions of


constant temperature storage only, while temperatures most
often vary during the processing and storage of foods.
In theory, the kinetics of ascorbic acid degradation
for a given food system, in conjunction with the
temperature history of the system, should provide
Efficient information to make accurate predictions of the
t

r

74

amount of ascorbic acid degradation occurring during a


given nonisothermal process. The present investigation
was undertaken with the aim of developing a procedure in
which actual dehydration data are used for establishing the

mathematical kinetic model which describes the rate of the


nutrient destruction as a function of the drying process
parameters.

Vitamin C is calculated from the following expression:


Vitamin C = V . C . (176/178) mg.

where: V = volume of N-Bromosuccinimide (mg/ml)

C = concentration of N-Bromosuccinimide (mg/ml)


The vitamin C contents were expressed as mg/lOOg on dry
matter basis.

The approach taken to develop equations for quality


]»oss as indexed by vitamin C was the classical isothermal
kinetics approach. The retention of the given nutrient was
measured as a function of time at constant temperature,

keeping the other independent variables constant. The


reaction constant is determined from the equation

-S!£ = KCn (4.31)


dt

in which n, the order of the reaction is 1 for this system.

Vitamin C was selected as the test nutrient because of its


temperature and moisture-sensitivity. Since the;
degradation of ascorbic acid was studied as a function ol

temperature, the applicability of the Arrhenius equation


was tested.


75

k = Jc0e'v *r (4.32)

where: 0748421 ^ - ^

C = concentration of the nutrient at any time,


t = time, hr

n = order of the reaction

k = reaction constant, 1/hr


Ea = activation energy, J/mole
R = molar gas constant, J/mole K
T = absolute temperature, K
k0 = frequency of molecular collision.
The constancy of k0 has been experimentally confirmed in
the case of many homogeneous reactions taking place both
in the dissolved and gaseous state, over limited
temperature ranges and providing no phase changes occur.

Retention of ascorbic acid is usually considered an


index of the mildness of a heat process. It is the

intention here, to analyse the experimental data and

attempt to formulate an equation which will express vitamin


C loss as a function of two important parameters: drying
air temperature and drying time.
The application of a first-order reaction equation
is more evident if equation (4.31) is solved and expressed
in the following form:

In— = -Jet (4.33)


Cn
o
76

Equation (4.33) can be rewritten in the following manner:

log C = — ~kt ♦ log CQ (4.34)


2.303 a 0

Applying the initial conditions, i.e at t=0, C=1 then the


Jr
constant of integration is zero. Hence, by plotting the
data on semi-logarithmic coordinates, the rate constant (K)
can be evaluated directly from the slope of the resulting

curve for each drying air temperature (K = 2.303 x Slope).


Figures 4.12-4.14 below show typical curves for the

—— An I low rale(00054) An How rale(00108)

Slice t hi c kness( mm) 2

Figure 4.12 Concentration-time curve for ascorbic acid in


Potato slices dried at 50°C.

G*perimental study to determine the loss of vitamin C. In


theory, the air flow rate does not directly affect the rate
degradation of ascorbic acid during thin layer drying.

77

The air flow rate may affect the rate at which the potato
slices are heated and therefore the loss of ascorbic acid
»•
may be affected by the flow rate. The figures clearly

depict the same trends of ascorbic acid degradation.

Time, I, min

— Airflow rale|0.0054) — Aul l ow iale(O.OIO0|

Slice Ihickness(mm) 6

Figure 4.13 Concentration-time curve for ascorbic acid in


potato slices dried at 60°C.

From the graphical representation, it is apparent that


the first order reaction equation represents the vitamin
c degradation best. Hence, it is concluded that the
* ♦
78

mechanism of vitamin C loss is most likely diffusion to the

surface of the slice. Subsequently, in addition to heat

destruction, oxidative destruction of some of the original

ascorbic acid content inevitably occurs (Woolfe, 1987).


Vitamin C exists in the slices in both the reduced and
a
oxidized forms. Here, the term 'vitamin' is a physiological
one rather than a chemical one, expressing a certain
physiological activity, which is related to the chemical
substances which are responsible for this activity.

—— Airflow rale(0 0 05^| Air M ow iaie(0 010B)

Slice thickness(mm| 4

fig.4.14: Concentration-time curve for ascorbic acid in


potato slices dried at 70°C.
r

79

The most common compounds with vitamin C activity (vitamers

of the vitamin C group) are ascorbic acid, dehydroascorbic

acid and ascorbyl palmitate. In the fresh potato slice, the

reduced form, 1-ascorbic acid is quantitatively the most

important (Woolfe, 1987). The enediol group on C-2 and C-3

of this compound can be oxidized to a diketo group in a

reversible reaction. The resulting dehydroascorbic acid has

full vitamin C activity. However, dehydroascorbic acid is

more labile than the reduced form and is readily, and

irreversibly, oxidized to 2,3-diketogulonic acid, with

consequent loss of vitamin C activity. A kinetic model

which describes the loss of ascorbic acid during air drying

based on the actual experimental data points is developed

later in this section.

Although there is virtually a limitless choice of


mathematical models that can be used for fitting equations
to experimental data, the considerations in choosing the
model followed these main guidelines:

1. Simplicity of the model

2. Lowest possible deviation between the model

predictions and experimental data


3. No discontinuities within the range of
variables

4. The deviation of the predicted value from the

experimental ones should be uniform over the

whole range and without a consistent trend.

Table 4.9 presents a summary of the first-order rate


constants for the degradation of ascorbic acid in potato
slices under all the drying conditions in this study.


80

Table 4.9 Experimentally determined first-order rate


constants for reduced Ascorbic acid

Temperature Slice Airflow Rate


(K) thickness rate(Kg/m2s) constant,K,
(mm) 1/min.
323 2 v 0.0054 0.0050
323 2 0.0108 0.0079
323 4 0.0054 0.0038
323 * 4 0.0108 0.0038
323 6 0.0054 0.0034
323 6 0.0108 0.0032
333 2 0.0054 0.0082
333 2 0.0108 0.0139
333 4 0.0054 0.0058
333 4 0.0108 0.0065
333 6 0.0054 0.0045
333 6 0.0108 0.0059
343 2 0.0054 0.0092
343 2 0.0108 0.0130
343 4 0.0054 0.0088
343 4 0.0108 0.0110
343 6 0.0054 0.0063
*34 3 6 0.0108 0.0075

The disappearance of ascorbic acid in each case can


be described by a first order reaction model. This
conclusion is similar to the results reported by several

workers including Waletzko and Labuza (1976) and Wagy and


Smoot (1977). These authors described the loss of ascorbic

acid using a first order kinetic model (Laing et al.f


1978) .

One of the more important factors in the study of


reaction rates during processing is the influence of
temperature on the reaction. One of the widely used
methods of describing this was presented by Arrhenius in

1899 (equation (4.34)). The same equation can also be

expressed as;
81

In K ~1± ♦ In K0 (4.35)
RT

where K0 is a constant of integration, also referred to as


a frequency factor.

Plotting the reaction rate constant (K) versus the

inverse of absolute temperature on semi-logarithmic

coordinates leads to the evaluation of the activation

energy (Ea). Figures 4.15 to 4.17 illustrate the influence


of temperature on rate constant, K for reduced ascorbic
acid in potato slices of 2 mm, 4 mm and 6 mm thickness
respectively. The values of K calculated by means of
equation (4.34) and as tabulated in table 4.9 were plotted
as

— 2mm, 0 0054 l:g/m*s

figure 4.15 Arrhenius plot illustrating influence of


temperature on rate constant, -K for 2 mm thick slices
* ♦
82

a function of the inverse of the absolute temperature and

seem to indicate that K varies exponentially with


temperature, approximating to an Arrhenius type equation.

From the Arrhenius equation the value of Ea corresponding

to the degradation process ranged from 24-48 kJ/mole, with


an average of 34+8 kJ/mole. However, data on chemical
changes during dehydration present an important problem.
There is a substantial need for critical reviews of
analytical procedures in connection with changes in vitamin
C content during dehydration processes.

— 4mm, 0 005 4 Ig/m^i

Figure 4.16 Arrhenius plot illustrating influence of


temperature on rate constant, K for 4 mm thick slices


83

From the graphs, it is clear that at each drying air

temperature, there is little difference between vitamin C


i

loss for different slice thicknesses. In addition, using

Arrhenius equation, activation energies for different ijslice


thicknesses were calculated and the values given in Table
4.10. These values are of the same order of magnitude and

cosequently, a single value of activation energy based on


the rate constants for all the slice thickness was
calculated and found to be 33.9 + 8.4 kJ/mole.

6mm, 0 006'1 lglm*$

Figure 4.17 Arrhenius plot illustrating influence of


temperature on rate constant, K for 6 mm thick slice

The rate of relative loss of vitamin C is same all


slice thicknesses tested. However, in absolute terms the
thiner slices loose comparatively more vitamin C than the
thicker slices because of the increased exposure of cells
* ♦
84

to enhance the chemical reactions.

Table 4.10 Experimentally determined Arrhenius equation


constants for reduced ascorbic acid degradation

Slice Air flow Ea K0 (1/min)


thickness rate (kJ/mole) X lo3
(mm) (kg/m2s)
2 0.0054 -26.110 1.047
2 0.0108 -23.936 0.578
4 0.0054 -38.076 59.48
4 0.0108 -47.868 188.8
6 0.0054 -28.721 2.079
6 0.0108 -38.494 66.64

The nutritive value of most products is more affected by

pretreatment of raw material and storage of dried products


than by the drying operation itself. It is therefore
sufficient to express the rate constant in the first-order
equation for reduced ascorbic acid degradation as a

function of drying air temperature by the Arrhenius
equation as follows:

In K = I n Kq 11 (4.36)
RT

where R is the molar gas constant in kj/mole k. and T is


the absolute drying air temperature (K).
The rate constant, K thus obtained is then substituted

into equation (4.34) to give the final equation for loss


of vitamin C as a function of drying air temperature and
drying time.
Although in judging quality of dehydrated foods,
priority is given to wholesomeness and nutritive value,

85

subjective quality factors such as visible colour and


reconstitution (as indicated by the rehydration ratio)
should be considered as well. Customer appeal dictates; the
»

potato slices to be of characteristic shape and retain

their original colour after drying.

4.3 Visible Browning of the dried product

Hendel et al. (1955) used rates of browning,


determined at constant moisture contents, to predict heat
damage during drying of white potato dice. They found a
reasonable agreement between estimated and observed
browning values. The data presented showed that the rate
of browning is most rapid at a moisture content of about
,15-20 per cent dry-basis.
Browning kinetics and drying behaviour can be combined
to determine times at which a level of browning,
arbitrarily chosen to be the limit of acceptability, is
achieved. Usually, the formation of brown complexes is

measured directly from the increase in absorption of the


extracted solution at a set wavelength. This one is
commonly set somewhere between 400 and 490 nm, which is

considered to give a visual assessment of the amount of


browning (Resnik and Chirife, 1979). In this work, a
visible browning value, A420 nm (optical density) of 0.30

was adopted from Hendel et al. (1950) as the limit of

acceptability of browning of the fully dried slices


(moisture content of 10 per cent wet basis). Table 4.11
shows the optical density of slice extract from lOg of
dried product attained under the various drying conditions
studied. The colour values relate to eye appeal and
* ♦
86

flavour. Generally, as the product browns, they become


less desirable (Sullivan et al., 1980).

Table 4.11 Visible browning values (A420 nm) of dried


slices

Drying Airflow Slice Drying Visible


Air Rate Thickness Time Browning
Tempera (kg/m2s) (mm) (min.)* (A420nm)*
ture
(°C)
50 0.0054 2 145 ± 2 0.205 + 0.015
50 0.0108 2 130 ± 7 0.124 + 0.005
50 0.0054 4 275 ± 5 0.222 + 0.008
50 0.0108 4 252 + 5 0.051 + 0.001
50 0.0054 6 387 + 3 0.033 + 0.003
50 0.0108 6 384 ± 3 0.041 ± 0.006
60 0.0054 2 116 + 2 0.119 + 0.008
60 0.0108 2 112 ± 2 0.105 + 0.022
60 0.0054 4 237 + 3 0.083 + 0.008
60 0.0108 4 205 + 4 0.060 + 0.006
60 0.0054 6 359 ± 2 0.031 + 0.003
60 0.0108 6 361 ± 1 0.064 + 0.003
*
70 0.0054 2 147 ± 2 0.253 ± 0.004
70 0.0108 2 126 + 2 0.073 + 0.006
70 0.0054 4 175 + 3 0.104 + 0.009
70 0.0108 4 173 ± 2 0.054 + 0.001
70 0.0054 6 285 ± 4 0.042 + 0.005
70 0.0108 6 284 ± 3 0.050 ± 0.002

* = mean and standard deviation for three replications.


The table shows that the limit of acceptability of

browning (A420 nm °f 0.30) of the fully dried slices was


not exceeded in any of the drying conditions studied.

Figure 4.18 shows a typical trend of level of browning in

the course of drying the potato slices. Using the same


sulfhiting level, Kabira (1990) found that the levels of
residual S02 in the dried raw slices were slightly above
the upper limit of 500-550 ppm recommended for processed
potato products in terms of visible browning (Roberts and
McWeeny, 1972; Hadzidev and Steel, 1979).


87

m
<&

AVG A420 nm

Drying air lempeialure ( C )• 60


Air How rale (kg/m*s) 0.0054
Slice thickness (mm): 2

Figure 4.18 Trend of browning level during drying of potato


slices

However, further reductions in S02 levels are expected

during storage and preparation for consumption

(Sahasrabudhe et al., 1976; Taylor and Bush, 1986;

Wedzicha, 1987).

Figure 4.18 appears to show browning as decreasing

after attaining a maximum value at 80 minutes residence

time. That could be concluded if the optical density, A420

nm was given per gram dry matter. However, as shown in the

figure, it is given per lOg partially or fully dried

product. Hence, the trend perhaps may be explained on the

basis of the various roles which water can play in


88

partially dehydrated food systems. The reactions between

the food components involved in nonenzymatic browning are

often accelerated during air dehydration, because the

dissolved reactants are concentrated as the water is


V

removed, and the temperature is raised due to heat transfer

to the food. At high moisture contents water may have; an

inhibitory effect due to the fact that most of the heat

supplied is used to evaporate the moisture rather than

raising the temperature of the material. On the other

hand, low moisture contents may decrease the reaction rate

due to an increasing diffusion resistance which lowers the

mobility of the sugars and the organic acids which catalyze

the reaction.

The treatment with a 0.5 per cent Sodium

Metabisulphite solution was effective in controlling the

colour before and during drying. Browning during the

drying may be due to the action of polyphenol oxidase

enzyme and/or non-enzymatic maillard reactions between

carbohydrates and amino acids at elevated temperatures.

Because of the importance of obtaining samples which are

free from browning on processing, considerable efforts have

been made to specify an acceptable upper limit of reducing

sugar. This is usually about 0.10% on a FW basis but

levels up to 0.25% may sometimes be accepted (Harris,

1978) .
89

4.4 Case-hardening Behaviour

It has been observed that during the drying of some

food materials, a hard impermeable skin often forms at the

surface. This usually results in a reduction in drying

rate, and the phenomenon is usually known as Case-

hardening. The exact mechanism of case-hardening is far

from fully understood but is probably influenced by a

number of factors, including migration of soluble solids

to the surface and high surface temperature towards the end

of drying resulting in complex physical and chemical

changes in the surface layer.

In addition to colour, rehydration of the product was

a physical attribute evaluated. These two properties were

chosen because they most clearly reflect the quality of the

dried product. Fast rehydration or water pickup is a

desired characteristic of a dehydrated material. The

greater the dehydration, the more porous the product, and

the more readily the water is picked up, as reflected by

rehydration ratio values.

A plot of rehydration ratio against time for all the

eighteen drying conditions in this series of experiments

was used to quantify the onset of case hardening.

Rehydration ratio in this context was obtained as follows;

a log sample was placed in boiling water for 25 minutes,

then drained and weighed. The amount of water gained per

gram of dry solids was determined by subtracting the



90

original sample weight (log) from the weight of the

rehydrated potato slices and dividing this weight by that

of the solids in the lOg sample. Thus, rehydration ratio

is defined as the mass of water (grams) gained per gram of

dry matter of sample. Generally, the rehydration ratio

increases with a reduction of moisture content of a dried

product. However, in some cases, the rehydration ratio

increased and then started decreasing before the attainment

of the 10 per cent moisture content (wet basis) . Such

drying conditions were considered inappropriate as they

indicated the occurrence of case-hardening of the potato

sfices. Table 4.12 shows the values for rehydration ratio

for the fully dried slices.

The most pronounced changes that take place in food

materials on dehydration are associated with shrinkage.

The change in rehydration ratio value is no exception, hs

water is removed from cells of potato tissue, the cells

shrink and cell walls may rupture. Adjacent fibres may

fuse together and be cemented by residues of soluble

constituents. The extent of these changes depends on the

rate of drying. An initial rapid drying is desirable

because a rigid outer dry layer is formed while the wet

interior has its original volume. During the latter stages

of drying, the interior shrinks back against the rigid

shell, leaving a porous structure of cracks and voids that

facilitates rehydration. ,If drying is slow so that


* ♦
91

shrinkage is uniform throughout the volume, the result is

a hard, dense product that rehydrates with difficulty.

Potato pieces dried rapidly in the high-moisture range

continue to dry relatively rapidly in the low moisture

range because they become somewhat porous, and this

structure favours more rapid moisture diffusion.

Table 4.12 Rehydration ratio values for fully dried


slices

Drying Airflow Slice Drying Rehydration


Air Rate Thickn Time Ratio (g H20/g
Tempera (kg/m2s) ess (min.)* dry matter),
ture (mm) *
(°C)
50 0.0054 2 145 ± 2 2.08 ± 0.11
50 0.0108 2 130 ± 7 3.21 ± 0.15
,50 0.0054 4 275 ± 5 1.99 ± 0.05
50 0.0108 4 252 ± 5 1.93 ± 0.03
50 0.0054 6 387 ± 3 1.32 ± 0.04
50 0.0108 6 384 ± 3 1.49 ± 0.04
60 0.0054 2 116 ± 2 3.95 ± 0.04
60 0.0108 2 112 ± 2 3.90 ± 0.03
60 0.0054 4 237 ± 3 2.33 ± 0.02
60 0.0108 4 205 ± 4 1.93 + 0.07
60 0.0054 6 359 ± 2 1.44 ± 0.01
60 0.0108 6 361 ± 1 1.32 ± 0.03
70 0.0054 2 147 ± 2 3.54 ± 0.08
70 0.0108 2 126 ± 2 3.93 ± 0.12
70 0.0054 4 175 + 3 2.12 ± 0.07
70 0.0108 4 173 + 2 2.41 ± 0.05
70 0.0054 6 285 ± 4 1.90 ± 0.05
70 0.0108 6 284 ± 3 1.53 ± 0.02

* = mean and standard deviation for three replications.

The rehydration ratio increases with a reduction of


moisture content of a dried product due to the gradual
increase in porosity as drying progresses. Thus, ideally
the rehydration ratio value for' fully dried potato slices
should be the maximum attained in a graphical plot of
t
♦ ♦
92

rehydration ratio versus residence time. Consequently, a


structural defect was considered to have occurred in those
9

cases in which the rehydration ratio increased, reached a


peak value and then started decreasing before the
attainment of the 10 per cent moisture content (wet basis)
of fully dried potato slices. In this study, the decline

in the maximum extent of rehydration ratio for samples

tested as the drying process progressed was reasonably


used as indicator of case-hardening occurrence (figure <1.19
and 4.20). Figure 4.19 shows a typical drying condition
depicting the occurrence of case-hardening for partially
dried slices.

Cj

- b- AVG R Ratio

Drying air temperature I C) 50


An How lale (kglmS) 0 0108
Slice thickness |mm| 6

Figure 4.19 Drying condition illustrating case-hardening


♦ ♦
93

The drying time required was 6 hours and 25 minutes whereas


case-hardening occurred after a residence time of 5 hours.
Case-hardening occurred in nine out of the eighteen
drying conditions studied as shown in table 4,13.
>
Subsequently, the drying conditions depicting such a
behaviour for partially dried potato slices were considered

inappropriate. It was assumed that leaching of material


during the determination of rehydration ratio was a
constant for all drying conditions studied.

Table 4.13 Residence time for case-hardening occurrence

Drying Airflow Slice Residence Maximum


Air Rate Thick Time Rehydration
*Temper (kg/m2s) ness (min.) Ratio
ature (mm) (g H20/g dry
(°C) matter)
(mean + sd)

50 0.0108 2 120 4.14 + 0.42


50 0.0108 6 300 1.73 ± 0.11

60 0.0054 2 80 4.39 + 0.07


60 0.0108 2 100 4.47 + 0.13
60 0.0108 4 150 2.17 ± 0.09
60 0.0108 6 240 1.58 ± 0.03

70 0.0054 2 120 3.64 + 0.07

70 0.0108 2 60 6.05 ± 0.14

70 0.0054 6 240 2.01 ± 0.10

For each hydration study, the surface water of the slices



94

was removed by a filter paper (Whatman No. 41).


On the other hand, nine of the eighteen drying

conditions were successful in the production of a fully

dried product devoid of a hardened case. Figure 4.20 is

representative of such conditions. In this particular

case, the drying time is 2 hours (120 minutes).

AVG. R.Ralio

Drying air temperature |°C): 50


An How rale (kg/m*s): 0.0054
Slice thickness (mm): 2

Figure 4.20 Drying condition illustrating non-occurrence


of case-hardening.

4.5 Appropriate drying conditions

The course of dehydration of potato slices comprises


essentially two phases. The first, negligibly short at
high drying air temperatures (60°C and 70°C), during which
free moisture persists on the cut surface is referred to
as the constant rate phase. The second, in which the
* ♦
95

diffusion of moisture from the interior to the surface of


each slice becomes the limiting factor in the rate of
drying (the falling rate phase). Potato slices dri(;d
rapidly in the high-moisture range continue to dry
relatively rapidly in the low-moisture range because they

become somewhat porous, and this structure favours more

rapid moisture diffusion. However, as drying is continued

through the range of lower moisture contents, the rates of


moisture removal becomes slower, the concentration of
reactants within the potato slice is increased, and the

danger of damage to the product through nonenzymatic


browning increases. Therefore, it is desirable that
temperature of the drying air be progressively reduced as
drying proceeds through the final stages.
„ Table 4.14 shows the appropriate constant drying
conditions commensurate with the production of high
quality, storage-stable slices as realised in this study.
The fully dried product was acceptable in colour and devoid
of a hardened case.

Table 4.14: Appropriate constant drying conditions for


potato slices

Drying air Slice Airflow rate


temperature thickness (kg/m2s)
(°C) (mm)
50 2 0.0054
50 4 0.0054
50 4 0.0108
50 6 0.0054
60 4 0.0054
60 6 0.0054
70 4 0.0054
70 4 0.0108
70 6 0.0108

t

96

4.6 Optimum drying conditions

In section 4.5, drying conditions were selected with


#
the help of t an empirical critical criterion that

differentiates between the occurrence and non-occurrence


of case-hardening. In this study, modelling for

determination of optimum drying conditions (Temperature,


Slice thickness, and Airflow rate) was constrained by
unavailability of hypodermic thermocouple probes for
monitoring sample temperature during drying. The use of
dynamic optimization procedures in food processing
operations in general has been limited, and this is true

in particular of dehydration. Critical phenomena and drying


kinetics and dynamics are rarely interrelated, demanding

extensive time-consuming experimentation yielding results


of limited reliability. In this section, techniques are
reported for finding the optimal dryer temperature control
path for maximizing ascorbic acid retention and for

minimizing drying time with a constraint on resultant


ascorbic acid retention.
The primary objective of drying is the removal of
moisture, with the attendant structural, chemical, thermal,

biological and other property changes being significant as


well. Temperature is second in order of importance among

the drying process parameters, as it determines the

biological, chemical and physico-chemical changes in the

material. The extent to which the mentioned objectives are


achieved is a function of the design process parameters.
To carry out an optimization problem of the drying process,
a mathematical model of the process is required, some


97

criteria to be optimized and constraints.


A kinetic model representing ascorbic acid degradation
in potato slices as a function of moisture content and

temperature may be used to find optimal dryer-temperature


control paths for minimizing ascorbic acid loss during air
drying. Optimal dryer-temperature control may also be

determined for minimizing drying time given a specified


minimum retention of ascorbic acid. Constraints are placed
on the final moisture content and the air temperature.

Optimization may be achieved using a simulation-

optimization approach based on the complex method.

4.6.1 Models

Mathematical models are necessary for the analysis and


design of the drying process. The complexity of the:

appropriate model depends on the purpose of the application

considered. Detailed models are needed for dynamic


simulation and control, while simplified models are needed
for design or state estimation techniques. It is importeint
that the models presented here be sufficiently accurate to

warrant their use in optimization, yet easily and


efficiently evaluated (Gili et al., 1981).

4.6.1.1 Kinetic model for ascorbic acid


In general, the degradation kinetics of nutrients are
a function of moisture and temperature. Unfortunately, a
food product traverses a large range of moisture and
temperature during drying. Thus, a kinetic model must
necessarily be representative over this range. The kinetic
model representative of ascorbic acid degradation of the
potato slices may be obtained using a dynamic test approach

98

(Saguy et al., 1978; Mishkin et al., 1982). An empirical


first-order kinetic model is used,

-kC (4.37)
dt
r
where C is the concentration of ascorbic acid (normalized
with respect to initial concentration). Arrhenius
temperature dependence is used to model the first-order
rate constant,

, E.
k - k 0 exp[-^i] (4.38)

where k0 and Ea have moisture functionality,

In k 0 = + p2M + p3M 2 (4.39)

E* = Pi + P 5W + P sm2 + PiM ' (4.40)


0

M is the moisture content in g/g-solids, and Pj-Py are


parametric constants to be estimated using the proposed
modeling method. Table 4.15 shows parameter estimates for
the moisture and temperature dependence of the rate
constant associated with this model.

Table 4.15: Ascorbic acid kinetic model parameter values


for Equation (4.37) and (4.38) (Mishkin et al., 1984b)

Parameter Value

pl 16.38

P2 1.782

P3 1.890

P4 14831.0

P5 241.1

P6 656.2

------------ __________________ ' 236.8


99

4.6.1.2 Moisture model

The model representing the moisture content of the


slice during drying is essentially identical to that
presented by Mishkin et al.(1984a). 4mm thick slices of

approximately 4.0 cm diameter may be used for modeling.

These are the initial dimensions, i.e before shrinkage:


occurs. Edge effects are ignored due to the high ratio of

diameter to thickness. Fick's law for unidirectional

diffusion may be used to describe moisture transfer during


drying,

dM a ms
(4.41)
a t ax

where D is the effective diffusion coefficient (cm2/sec).


The diffusion coefficient has the following Arrhenius
temperature dependence:

D = D I0t exp [ - ^ (1 - -J_)) (4.42)


K 1 T zef

T ref = 3 1 3 K * Dref and ED have moisture functionality given

by Equation (4.43) and (4.44) (Luyben et al., 1980):

D r & f = exp [ - ^ 5* ^6 M ]
(4.43)
1 ♦ P*7 M

. = P *e + P 1, M
(4.44)
1 * ^io M

Equation (4.41) may be approximated using a finite


difference approach. The initial nodal format is
essentially the same as that of Mishkin et al., 1984a. For
100

the interior nodes the differential equation describing the


i-th volume element is Equation (4.45),

dMi Dj [M1 n - 2M t + (4.45)


dt (fi-Kj) 2

It is assumed that D is nearly uniform if the nodes are


close enough. Equation (4.45) is based on the assumption
of uniform bulk solids density. Shrinkage is assumed to be

proportional to decrease in moisture concentration (Mishkin


e t ., 1984a) ,

- ax. [P1, £ * P‘,1 (4.46)


"a

where 3x0 is the initial distance between adjacent nodes,


s
M0 is the initial moisture content, and Px3 and P x4 are

parameters estimated from drying data. On the basis of


order of magnitude analysis, the surface volume element is
assumed to be at the equilibrium moisture content (Me(J) .
Due to the symmetry of the problem, only half the slice
must be modeled as drying occurs identically from both
sides of the potato slices. Hence, for the centerplctne
node,

dM, ~d n
(4.47)
dc 2 ( 6 x n) 2

where,

t>xN - 1/2 t>x0 [P1, ^ + PJ4] (4.48)


101

and N is the number of volume elements used in the finite

difference scheme. N may be reduced to the lowest possible


without compromising accuracy, by trial and error. The
average moisture content (M) is expressed by Equation
(4.49):
>

<«,*«„>/ (4.49,
M = _ __ •** 2

Estimated parameter values for Equations (4.43) and (4.44)


are listed in Table 4.16 (Mishkin et al. 1984a). For drying
simulation Meq may be estimated using sorption data from
Gorling (1958).

Table 4.16: Moisture transfer model parameter values for


Equation (4.43) and (4.44) (Mishkin et al., 1984a)
Parameter Value
i

M
tn

20.80
PX6 85.34
PX7 7.247

PX8 23830.0
pi 45700.0
^ 9
pi 7.006
____ _ 1 D _____________

4.6.1.3 Temperature model

Ascorbic acid stability is a function of the sample


temperature and moisture content. A heat balance oh the
drying slice is used to determine the temperature of the
potato slice (T8). T 8 may be modelled using an algebraic
approximation to the heat balance (Mishkin et al., 1982).
Noting that the sensible heat term in the heat balance is
small and assuming temperature is uniform,


102

. _ Hw ms dM + r
(4.50)
s h A dt

where m B is the mass of solids in the slice (g), (dM/dt) is


the rate of vaporization of water, Tdb is the air

temperature (°C) and Hw is the latent heat of vaporization


for water (cal/g). Temperature uniformity may be verified

experimentally by inserting fine thermocouple probes at

various depths within potato slices during drying. To


prevent heat conduction into the slices and therefore
erroneous measurements, air-exposed portions of the probes
may be insulated. Because of shrinkage and other surface
effects, h and A are variable during drying and depend on
i

moisture content, as follows:

h A = A0 (P^ M + P*2) (4.51)

where h is the heat transfer coefficient (cal/min°C) , A is


surface area of sample (cm2) and A0 is initial area of
sample, (cm2). P I1 and Px2 are found by least-squares
fitting of the numerically integrated solution to the heat
balance equation to the temperature data for various drying

temperatures. For simulation purposes mB may be

approximated by Equation (4.52),

Po D d a L (4.52)
* 4 (1 ♦ M0)

where pQ is the initial slice density (g/cm3) of the potato


with diameter d (cm) and thickness L (cm).


103

4.6.2 Maximizing ascorbic acid retention

The above models may be used to simulate ascorbic acid


degradation for the desired range of drying conditions and
provide a basis for selection of control schemes which
improve ascorbic acid retention. The air temperature should

be low during the early stages of drying, in view of the


high temperature-sensitivity of ascorbic acid at elevatcsd

moisture conditions. As the moisture content is reduced,

the nutrient is rendered less sensitive to degradation and


the air temperature may be increased. This is precisely the
scheme which the optimal profiles are expected to depict..

The drying model may be used in conjunction with the


kinetic model developed for ascorbic acid kinetics to find
optimal temperature control paths for maximizing ascorbic

gcid retention in potato slices. Numerical integration of


the system of simultaneous differential equations may be
accomplished using techniques which are specifically
designed to handle stiff systems of differential equations
with increased efficiency while avoiding instability (Gear,
1971) .

The complex method (Box, 1965) may be used to find the

optimal temperature control scheme during drying using the


methods described by Mishkin et al.(l982). The dryer
temperature may be represented by the time-dependent
polynomial decision function (Mishkin et al., 1984a):

= a + btn + ctn2 + dtn3 + etn* + ft„5 + gtn* (4.53)

where tn = t/tf and the parameters a,b,c,d,e,f and g are

the decision variables, manipulated by the complex method


in order to maximize the objective function,


104

max J = C( tf) (4.54)

where C(tf) is the final ascorbic acid content of the dried


potato slices. A constraint is placed on the final average
»
moisture content (Mf <, 0.11 g/g-solids) . It is coincidental
that the usual moisture-temperature regime during hot-air
drying is naturally protective against ascorbic acid

degradation. During the early stages of the falling-rate

period, the rate of drying is sufficiently high to keep the

drying potato slice temperature low, which counteracts in


part the sensitivity of ascorbic acid to degradation at
high moisture contents. As the drying progresses, the
temperature rises but the nutrient becomes more stable
because of the falling moisture content. In view of this
naturally stable regime, use of the prescribed optimization

technique improves retention of ascorbic acid by 5-10% over


the comparable results for the constant-temperature
processes (Mishkin et al., 1982).
By specifying the functionality of the temperature

profile as a sixth order polynomial, a constraint is


imposed. It is probable that the "true" solution cannot be

modelled exactly by this function. Application of the


direct-search method with discretized decision exhibits
imposes a similar constraint. However, application of the
maximum principle to drying optimization (Mishkin et al.,

1982) does not impose such constraints on the time-


;

dependent decision profile and therefore would be expected


to find the "true" optimal path. This advantage, however,

comes at great expense in view of the complications which


accompany the method.


105

4.6.3 Minimizing drying time with a constraint on ascorbic


acid retention
A major industrial concern aside from product quality
is processing time. Specification of the drying time is not
t
necessary for the general case, but because of the kinetics
of ascorbic acid degradation, low-temperature / long-time
(LTLT) drying is favourable; therefore, it is necessary to
constraint the drying time in this case. By reducing drying

time, for example, throughput can be increased along with


productivity. A minimal-time control process may be
achieved by exploiting the utility of the complex method.
A nominal drying process is simulated at a constant
temperature Tdb, 50°C £ Tdb < 70°C for a length of time,
t*, with a final moisture content of 0.11 g/g-solids. Using
the ascorbic acid kinetic model, the final retention is
determined for the simulated process, 28%, for example.
This may be used as a lower quality constraint in the
minimum-time study as shown in Equation (4.55):

C(tf) = 0.28 (4.55)

The complex method is used with a time-dependent polynomial


temperature decision function:

T & = a + btl + c(t1)2 + d(t')3 + eft1)4 + f(t1)5 + g(t1)6 (4.56)

where t1 = t/t* and t* is the length of the nominal drying

process described above (min). Equation (4.56) may be used

also in its linear form, i.e. with only parameters "a" and

Mb". The objective function for minimizing drying time is


described by Equation (4.57):
i

♦ ♦

i
106

M i n J = tf (4.57)

For a given temperature control profile tf is defined as


the point in time during the numerical integration of the
1
drying model where the moisture constraint is satisfied,

i.e. where M = 0.11 g/g-solids.

Obviously, there are many more possible control

variables other than air dry-bulb temperature which may be

manipulated to improve the drying process. Here, the air


dry-bulb temperature is the single decision variable; the
other variables are held constant. The aim is to find the
air temperature as a function of time which will result in
maximum retention of ascorbic acid given a specific drying
time, with the constraint that the potato slice is dried
to the desired moisture content of 0.11 g/g of solids.
Potato slice thickness or air flow rate might be more
applicable in some instances. Application of the method
described in this section for such problems is straight­
forward.

The optimization of potato slice drying operation is

important for the design of both economically and

technologically efficient dryers. However, necessary effort


should be devoted to the improvement of a fundamental model
before optimization for process control are taken * into
consideration, as it were, this topic would belong to the
final problems to be resolved for the designers.


5. CONCLUSIONS

1. The rate of drying of potato slices is a function of


drying air temperature, slice thickness and air mass flow
rate. Statistical analysis of experimental data showed the
t

drying air temperature had the greatest effect on drying


rate and quality, followed by slice thickness; air mass
flow rate had little effect. The falling drying-rate

period is the most important drying regime for potato

slices.
For an overall analysis of drying behaviour the
following series of assumptions were made to facilitate the

development of mathematical models:


(a) the potato slice is a slab and no shrinkage occurs

during drying
, (b) moisture is transferred from the slice by
diffusion to the surface, followed by evaporation
(c) diffusivity remains constant during the entire
process provided the drying conditions are held
constant
(d) negligibly small constant-rate drying period,

hence the slice has a uniform moisture distribution

initially
(e) the rate of moisture transfer at the slice surface
is proportional to the rate of moisture diffusion

through the slice to the surface.

Models developed using the statistical method of

successive residuals, based on Crank's diffusion equation


produced results which are in agreement with test data.
Three-term drying models were considered adequate for

lt>7
108

practical purposes. The major variable is the drying air

temperature. For a given value of drying air temperature,

the drying rate is inversely proportional to the sguare of


thickness of slices. Thus, for each of the tempercitures
50°C, 60°C and 70°C, a thin-layer drying eguation model was

developed. The drying equations showed much variability


in the drying rate parameter k. This is expected since the

diffusion coefficient, D, is mainly affected by


temperature. These eguations were based on regression
analysis for specific drying conditions. Therefore, a
semi-empirical exponential model was developed whose drying
'constant' is a function of both slice thickness and the
.air drying temperature.
2. The limit of acceptability of visible browning (A 420

nm of 0.30) of the fully dried slices (10% moisture

content) was not exceeded in any of the drying conditions


studied. The treatment with a 0.5 per cent Sodium
Metabisulphite solution was therefore effective in
controlling the colour before and during drying. This

study was however limited in scope because it utilized only


one variety.

3. To observe the effect of the drying conditions on the


structure of the potato slices, rehydration behaviour of
the dried slices were investigated. Nine of the eighteen
constant drying conditions were successful in the
production of a fully dried product devoid of case-
hardening as measured by the rehydration ratio variation

in the course of drying. Such favourable conditions were


characterised by low drying air temperature (50°C); medium
* ♦
109

drying air temperature (60°C) in combination with thick


slices (4 mm and 6 mm) and low air flow rate (0.0054
kg/m2s ) ; or high drying air temperature (70°C) with thick

slices (4 mm and 6 mm) and the higher air flow rate (0.0108
kg/m2s ) .

4. Degradation of vitamin C followed a first-order reaction

rate for all slice thicknesses tested and was dependent on


drying air temperature and drying time. The rate constant,
k, was expressed by the Arrhenius equation, anal the

resulting activation energy, Ea, found to be 34 kj/mole.


Hence, a nonlinear predictive model developed for vitamin
C loss was found to be

In K = In K0 - — ..99.

where R is the molar gas constant and T is the absolute


drying air temperature.

A disadvantage present in mathematical models based


on the Arrhenius equation is a decrease in sensitivity.

It does not separate the effect of simultaneously changing

variables and provides only an overall information of the


behaviour of the drying system for a particular set cf
conditions. Linear models are expected to be more accurate
since they are established based on a large number of
observations.

5. This study shows that predictive models can be developed

which can form the basis for control of drying conditions


in order to maximize nutrient retention. The development
of such predictive models for potato slices and
optimization of processing conditions are potentially of
* ♦
110

great economic value. The scope of this study was limited

to the establishment of appropriate drying conditions

commensurate with the production of a high quality,


storage-stable potato slices. Further research involving
the application of optimization techniques needs to be done
K
in order to obtain the optimum drying conditions.
6. SUGGESTIONS FOR FURTHER RESEARCH

On the basis of the results of this study, the

following aspects are considered as possible areas for


future research.
1. Validation of the drying and ascorbic acid loss
t
t
equations using conventional cabinet dryers.

2. Design and development work on forced-convection driers


for potato slices.

3. A mathematical description of the drying process based


on structural and thermodynamic assumptions.

4. A different procedure to testing nutrient stability


based on the determination of isolated parameters and
integrating these effects to obtain nutrient retention
for the complete process. This procedure will offer

greater flexibility in providing a means of calculating

effects of process variations without the need to carry


s

out the complete process. Studies should be carried out.


for:
(a) development of a linear mathematical model for
ascorbic acid degradation as a function of

moisture, sample temperature and time


(b) calculation of ascorbic acid retention as a

function of time and drying conditions by

combining the mathematical models for kinetics

of degradation and drying behaviour and


(c) comparisons of predicted values and values

obtained in drying experiments.


5. Optimization of the drying operation for process control

and energy saving.


7. BIBLIOGRAPHY.

Aoland, J.D., 1971. East African Crops. Longman Group


Ltd. London.
»

Akeredolu, F ., 1987 . Shrinkage in Dehydration of Manihot


Utilissma and Dioscorea Rotunda. In: Drying
Technology, 5(1), 107-128 (1987)

Augustin, J., Swanson, B.G., Pometto, S.F., Teitzel, C.,

Artz, W.E. and Huang, C.P. 1979. Changes in

nutrient composition of dehydrated potato products


during commercial processing. J.Food Sci. 44:216-219

Ballestrem, C.G., Holler, H.J. 1977. Potato Production


in Kenya, CIP. Lima - Peru.

Barakat, Z.M; El-Wahab, M.F.A. and El-Sadr, M.M., 1955.

Action of N-Bromosuccinimide on Ascorbic Acid: A new

titrimetric method for estimation of vitamin C.


Analytical chemistry 27(4)536-540.
Bhardvaj, K.C. and Kalra, S.K. 1981. Use of simple solar

dehydrator for drying fruit and vegetable products.


J. Food Sci. Technol. (India) 18: 23-26.

Bhattacharyya, G .K. and Johnson, R .A . 1977. Statistical

Concepts and Methods. John Wiley & sons, Inc.


New York, USA.

Box, H.J. 1965. A Hew Method of Constrained Optimization


and a Comparison with Other Methods, computer

Journal,8,pp.42-52.
Brannan, J.G; Butters, J.R; Cowell, H.D; Lilly, A.E.V.

1976. Food Engineering Operations. Applied Science


Publishers Limited. London.
Brookar, D,B., Bakker-Arkemma,F.W. and Hall, C.W. 1982.
113

Drying Cereal Grains. The AVI Publishing


Company,Inc. Westport, Connecticut, USA.
Burton, W.G. 1989. The potato. Longman Group UK Limited

1989. Copublished in the United States with John


Wiley and sons, Inc., New York.

Conte, S.D. and de Boor, Carl. 1980. Elementary Numerical

Analysis. An Algorithmic Approach. Third


Edition. McGraw-Hill, Inc. New York, USA.
Crank, J. 1975. Mathematics of Diffusions, 2nd edn. Oxford

University Press, London.

Crissman. L.M. 1989. Evaluation, Choice and Use of potato


varieties in Kenya. International potato centre (CIP)
, Lima-Peru.

Durr, G. and Lorenzl, G. 1980. Potato Production and


utilization in Kenya. CIP Lima, Peru.
Ede, A.J., and Hales, K.C. 1948. The physics of drying in

heated air, with special reference to fruit and

vegetables. Great Britain, Department of Scientific


and Industrial Research, Food Investigation Board,
Special Report 53.

FAO (1996). Production Yearbook 50, FAO, United Nation*;,


Rome.

FAO (1986) . Production Yearbook, FAO, United Nation*;,


Rome.

Gear, C.W. 1971. The automatic integration of differenticl


eguations. Comm, of ACM 14(3): 176.
Gill, P.E., Murray, W., Saunders, M .A ., and Write, M.H.
1981. Aspects of mathematical modelling related to
optimization. Appl. math, modelling 5:71
♦ ♦
114

Gorling, P. 1958. Physical phenomena during the drying of

foodstuffs. In: Fundamental aspects of the


dehydration of foodstuffs. New York, U.S.A, Macmillan
_ »
Co.

Gray, D. and Hughes, J.C. 1978. Tuber quality. In:

Harris, P.M. (ed.), The potato crop. Chapman and


Hall, London.

Gupta, 8.8. 1978. Dehydration of fruits and vegetables


using solar energy. Sun, Mankind's Future Source of
Energy vol. Ill: 2080-2085.

Hadzidev, D. and Steel, L. 1979. Dehydrated mashed


potatoes - chemical and biochemical aspects. Adv.
# Food Res. 25: 55-135.
Hall, C.W., 1980. Drying and Storage of Agricultural
Crops. The AVI publishing co. Inc. Westport,
Connecticut. USA.

Harris, P.M., 1978. The potato crop: The scientific basis

for improvement. Edited by P.M. Harris. Halsted


press, John Wiley and sons, New York
Hawkes, J.G. 1990. The potato. Evolution, Biodiversity
and Genetic Resources. Belhaven press, London.

Handel, C.E., Bailey, G.F., and Taylor, D.H. 1950.


Measurements of nonenzymatic browning of dehydrated

potatoes during storage. J.Food Technl. 4:344-347.


Handel, C.E., Silveira, V.G. and Harrington, W.O. (1955)
Rates of nonenzymatic browning of white potato during
dehydration, Fd Technol. Champaign 9, 433-8
Henderson, 8.M., and Pabis. 1961. Grain drying theory, 3:.
Temperature effect on drying coefficients. J. Agr.
♦ ♦
115

Enging. Res. 6(3): 169-174.


Henderson, 8.M., and Perry, R.L. 1976. Agricultural
Process Engineering, 3rd Edn. The AVI Publishing
y
Company, Inc. Westport. Connecticut.

Holdsvorth, 8.D. 1971. Dehydration of Food Products. A


»

review. J. Food Technol.(UK) 6:331-370.

Horton, D. 1987. Potatoes: Production, Marketing and


Programmes for Developing countries. Boulder,
Colorado: Westview press.
Igbeka, J.C. 1982. Simulation of moisture movement during
drying a starch food product - cassava. J. Food
Technology v o l .17:27-36.

J/ison, A.C. 1958. A study of evaporation and diffusion


processes in the drying of fish muscle. In
Fundamental Aspects of the Dehydration of Foodstuffs.
Soc. Chem. Ind. (London) 103-125.
Kabira, J.N. 1982. Storing Ware Potatoes. In: Ministry of
Agriculture. Workshop on potato production and
marketing in Kenya, held at N.S.Q.C.S Lanet, 15th-18t.h

June, 1982.
Kabira, J.N., 1983. Storage and Processing Characteristics
of three Kenyan Potato varieties. Unpublished M.Sc.,
Thesis. University of Nairobi, Nairobi.

Kabira, J.N., 1990. Feasibility of Small Scale Production


of solar dried potato (solanum tuberosum) flour and

its potential Food Uses in Kenya. Unpublished Ph.D.,


Thesis. University of Nairobi. Nairobi.
Laing, B.M., Schlueter, D.L, and Labuza, T.P. 1978
Degradation kinetics of ascorbic acid at high
«
* ♦
116

temperature and water activity. J.Food Sci.43:1440-


1443 .

Luyben, K.CH.A.M., Liou, J.K., and Bruin, S. 1980. Enzyme


degradation during drying processes. In "Food Process

Engineering: Enzyme Engineering in Food Processing",

(Ed.) P.Linko et al. vol.2, P.192. Applied Science


Publishers, Englewood, NJ.

Maeda, E.E. and Salunkhe, D.K. 1981. Retention of ascorbic


acid and total carotene in solar dried vegetables.
J.Food Sci.46:1288-1290.
Mishkin, M., Karel, M. and Saguy, I. 1982. Applications of

optimization in food dehydration. Food Technol.


3 6 ( 7 ) : 101 .

Mishkin, M.A. 1983. Dynamic modelling, simulation and


optimization of quality changes in air-drying of
foodstuffs. Ph.D. thesis, Massachussets Institute of
Technology, Cambridge, MA.

Mishkin, M., Saguy, I. and Karel, M. 1984a. Optimization

of Nutrient Retention During Processing; Ascorbic Acid


in Potato Dehydration, Journal of Food Science, 49,

pp.1262-1266.

Mishkin, M., Saguy, I. and Karel, M. 1984b. A Dynamic Test


for Kinetic Models of Chemical Changes during

Processing: Ascorbic Acid Degradation in Dehydration


of Potatoes, Journal of Food Science, 49, pp. 1267-
1274 .

Mohsenin, N.N. 1980. Physical Properties of Plant and

Animal Materials. Gordon and Breach Science


Publishers, New York.
i
* ♦
117

Moledina, K.H. and Flink, J.M. 1982. Determination of

ascorbic acid in plant products by HPLC.


Lebensm.wiss. u. Technol. 15:351-358.
Montero, R. 1991. Potato Drying in the Peruvian Andes:
In Appropriate Technology vol.17 No.4. ITDG Peru,
Vanderghen 235, Mira flores, Lima, Peru. 1991.

Nindo, C.I. 1991. Some performance aspects of seed maize

processing. Unpublished M.Sc thesis. University of


Nairobi. Nairobi.
Petersen, J.N. 1986. Analysis of batch drying data using
SAS. In: Drying Technology, an international
journal 4(3), 319-330 (1986).
Jjeeve, R.M. 1971. One cell: The thin slice of profit,
Amer. Potato J. 48, 47-52.

Resnik, 8. and Chirife, J. 1979. Effects of moisture


content and temperature on some aspects of
nonenzymatic browning in dehydrated apple. J.
Food Science (44) 1979, 601-605.
Roberts, A.C. and McWeeny, D.J. 1972. The uses of sulphur
dioxide in the food industry. A review. J. Food
Technol (UK) 7: 221-238.

Roberts, E.H. 1978. The Potato Crop. Edited by Harris, P.M.


Chapman and Hall. A Halsted Press Book, John Wiley $
Sons. New York.

Rumsey, J.R. 1985. Optimum airflow rates for walnut

drying. ASAE paper No. 85-3045. ASAE, st. Joseph,


MI. 49085.

Saguy, I., Mizrahi, S., Villota, R., and Karel, M. 1978.


Accelerated method for determining the kinetic model

118

of ascorbic acid loss during dehydration. J. Food 55ci.


43: 1861.

Sahasrabudhe, M.R., Larmond, E. and Nunes, A.C. 1976.

Sulfur dioxide in instant mashed potato flakes. Can.

Inst. Food Sci. Technol. J. 9:207-211.

Saravacos, O.D., and Charm, S.E. 1962. A study of the


mechanism of fruit and vegetable dehydration. Food
Technol. 16(1), 78-81.

Ghakya, B.R. and Flink, J.M. 1986. Dehydration of potato


3. Influence of process parameters on drying
behaviour for natural convection solar drying
conditions. J. Food proc. preserv. 10: 127-143.
Ghakya, B.R., Moledina, K.H., and Flink, J.M. 1986.
Dehydration of Potato 4. Influence of Process
Parameters on ascorbic acid retention for natural
convection solar drying conditions. J.Food Proc.
Preserv.10:145-159.

Steel, G.D.R., and Torrie, J.H. 1980. Principles and

Procedures of Statistics. A biometrical Approach.


International Edition. McGraw-Hill Book Company.
Singapore.

Sullivan, J.F., Craig, J.C. Jr., Konstance, R.P., Egoviile,


M.J. and Aceto, N.C. 1980. Continuous explosion­

puffing of apples, J. Food Sci. 45: 1550-1555; 1558


Taylor, S.L. and Bush, R.K. 1986. Sulphites as food

ingredients. A scientific status summary by the


Institute of Food Technologists Expert panel on Food
Safety and Nutrition. Food Technol. 40: 47-51.
Van Arsdel, W.B., 1963. Food Dehydration. Volume 1 -
t
* ♦
119

Principles. Edited by Wallace B. Van Arsdel and

Michael J. Copley. The AVI publishing company, INC.


Westport, Connecticut.

Van Arsdel, W.B. 1973a. Food Dehydration. 2nd Edn.Vol.l.


Drying Methods and Phenomena. The AVI Publ. Co. Inc.
Westport Connecticut.

Van Arsdel, W.B. 1973b. Food Dehydration. 2nd Edn. Vol.2.


Practices and Applications. The AVI Publ.Co.Inc.
Westport, Connecticut.

Von Loesecke, H.W. 1943. Drying and Dehydration of Foods.

Reinhold Publishing Corporation. New York, U.S.A


Wedzicha, B.L. 1987. Review: Chemistry of sulphur

, dioxide in vegetable dehydration. Int. J. Food Sci.


Technol. 22: 433-450.

Woolfe, J.A., 1987. The Potato in the Human Diet.

Cambridge University Press, Cambridge - Londpn. New


York.


8.APPENDICES
Appendix A: Statistical Analysis of Moisture Reduction
data

A.l: The Effect of Drying Conditions on the Time of


half response(minutes).

Drying Slice Air Time of half response (minutes)


Air Thickn flow
Temp. egg rate:
Rl R2 R3 Total Mean
(°C) (mm) kg/m^s

50 2 0.0054 54.5 61.8 54.6 170.9 57.0


50 2 0.0108 66.2 67.5 70.0 203.7 67.9
50 4 0.0054 133.3 110.0 137.8 381.1 127.0
50 4 0.0108 110.0 100.0 131.1 341.1 113.7
50 6 0.0054 180.0 182.8 185.7 548.5 182.8
50 6 0.0108 180.2 183.0 194.3 557.5 185.8

60 2 0.0054 37.3 37.7 38.2 113.2 37.7


60 2 0.0108 37.7 45.9 39.1 122.7 40.9
60 4 0.0054 91.7 94.2 95.0 280.9 93.6
60 4 0.0108 83.3 86.7 81.7 251.7 83.9
60 6 0.0054 151.4 142.9 148.6 442.9 147.6
60 6 0.0108 150.0 130.9 152.7 433.6 144.5

70 2 0.0054 44.4 40.0 46.9 131.3 43.8


70 2 0.0108 46.3 46.5 43.8 136.6 45.5
70 4 0.0054 73.8 70.0 74.0 217.8 72.6
70 4 0.0108 76.3 70.0 80.0 226.3 75.4
70 6 0.0054 124.4 110.0 130.0 364.4 121.5
*70 6 0.0108 116.0 104.0 120.0 340.0 113.3

Rl, R2 and R3 are the first, second and third replications


respectively.

1 n iv E K sm ' OF N a i r o b i UBRAJ*';

1«0
121

A . 2 : Time (minutes)*of half response (MR «* 0.5) for the


drying at three temperature levels A, for three slice
thicknesses B, and two airflow rates, C.

Slice thickness = B Total


Temp. = A Airflow = b| +
/
rate = C 2 mm ■ 4 mm = 6 mm =
b2 + b3
\
b.

50 * a| 0 . 0054=c, 170.9 381.1 548.5 1100.5


0 . 0108=C2 203.7 341.1 557.5 1102.3
Total 374.6 722.2 1106.0 2202.8

" c l+ c 2

60 » aj 0.0054-c, 113.2 280.9 442.9 837.0


0.0108-Cj 122.7 251.7 433.6 808.0
Total 235.9 532.6 876.5 1645.0

■ c l*c 2

70 = a3 0 . 0054=C| 131.3 217.8 364.4 713.5


0 . 0108»C2 136.6 226.3 340.0 702.9
Total 267.9 444.1 704.4 1416.4

“ c l+ c 2

Total 0 . 0054=c, 415.4 879.8 1355.8 2651


« a( + a2 0.0108=C2 463.0 819.1 1331.1 2613.2
+ a3 Total 878.4 1698.9 2686.9 5264.2
=c,+C2 =* G
* * Each value is a total OJf three replications for
layer of slices of 0.117 m2.

The above table is necessary for the computation of the

sums of squares of the main effects and interactions.

Analysis of Variance
1. Computation of the analysis of variance without regard
to the factorial arrangement of treatments for the
experimental design used.

Correction term = C = G 2/rt = 513181.51


122

Total SS = Y Y2^ -c = 115816.83 ( 8 . 1)

v2]
? ( 8 . 2)
Block SS = -L-----C = 541.70
* t

Treatment SS (8.3)
- C - 113877.54

Error SS = Total SS - Block SS - Treatment SS = 1397.59

2. The treatment sum of squares is partitioned into


components attributable to main effects and interactions.

E (^i)2
SS (A ) = -L— --- - c (8.4)
rbc

( 2202.82 + 16452 + 1416,A2) - C = 18181.92


3(3)2

2
SS (B) (8.5)
- C
rac

( 8 7 4 . 42 + 1 6 9 8 . 92 + 2 6 8 6 . 92) - C = 91111.79
3(3)2

[(C2) - (ct) ]2 (8.6)


SS (C) = - C
rab 2 rab

12.6512. + 2^13 T2 )2 - C « (2613.2 - 2 651^ 2 26.46


3(3)3 2(3)3(3)

♦ ♦
123

E <a W 3
SS (A B ) = ------- - C - SS {A) - 55(B) (8.7)
rc

374♦62 + ,..+ 70 4 .42 - C - (18181.92 + 91111.79) 3836.85


' 3(2)

E K C* )2 (8 . 8 )
SS (AC) = -- ---- - C - SS(A) - 55(C)

1 1 0 0 . 52 + . . .+ 702.92 - C - (18181.92 + 26.46) 2 6.69


3(3)

55 (BC) (8.9)
- C - SS(B) - S5(C)
ra

415.42 +.■.+ 1331.I2 - C - (91111.79 + 26.46) = 338.00


3(3)

E iaibj°k>
- i V j , k _______________
$S(ABC) -C-SS(A) -SS (B) -55(C) -SS(AB) -SS(AC) -SS(8C)

(8.10)

170.92 + ... + 3 4 02 - C - (18181.92 + 91111.79 + 26.46


3
+ 3836.85 + 26.69 + 338.00) = 355.83


124

Appendix B: Psychrometric Representation of Ambient


Conditions for the Air used in the Drying Experiments.

Thermostat Ambient
setting Condition
for drying of Air
air
50°C at wet-bulb dry-bulb Relative
inlet (°C) (°C) Humidity
(%)
Expt. 1. 17.33 21.31 68
2. 20.56 23.04 80
3. 23.53 25.74 81
4. 24.27 25.50 90
5. 24.76 25.72 84
6. 27.21 28.19 91
60°C at
inlet
Expt. 7. 27.70 29.16 90
8. 25.99 26.23 91
9. 26.72 26.97 94
10. 25.99 26.23 91
* 11. 24.76 25.01 95
12. 25.25 25.50 95
7 0°C at
inlet
Expt. 13. 25.50 25.74 95
14. 23.04 23.28 95
15. 22.54 23.28 93
16. 22.54 23.28 95
17. 21.56 22.79 90
18. 21.56 22.54 93


125

Appendix C: Equilibrium Moisture Content.


C.1: Static and Dynamic Equilibrium Moisture Content.
Equilibrium moisture content is an important parameter
in developing mathematical expressions for the drying
process. In evaluating this parameter experimentally, one
makes the distinction between static and dynamic

equilibrium moisture contents. The former is obtained by

exposing the material to a constant volume of air, at known

temperature and humidity for a considerable duration. The


latter is achieved by passing air of known temperature and
humidity through a sample of the material until its weight
remains constant. It therefore appears that static
equilibrium moisture content is more suited for use in
storage problems whereas the dynamic equilibrium moisture
cojitent should be applied to drying experiments.
It is the drying air temperature that brings about the
variation in relative humidity. Equilibrium moisture
content is predominantly a function of relative humidity,

with temperature having a secondary effect. At drying air


temperatures of 50°C, 60°C and 70°C for the potato variety

(Desiree) studied, the dynamic equilibrium moisture content

values obtained are shown (Table C.2). During the tests,


drying of triplicate samples was carried out in the drier
for several days until there was no change in moisture
content for three consecutive daily readings.

Control of air humidity during the tests would have


required the construction of a packed bed where incoming

atmospheric air is saturated with water vapour at a


particular temperature and then heated to the required
temperature and humidity. This was considered unnecessary
due to the fact that on heating the ambient air to the
* ♦
126

operating temperatures (50°C to 70°C), the resulting


humidity values are low (less than 40%) and of the same
order of magnitude.

Tabla C.2 Experimentally determined dynamic equilibrium


moisture content# s

Equilibrium
Drying moisture
conditions content (db)
Air Slice
Temperature thickness (mm)
(°C)
50 6 0.017 ± 0.004
50 4 0.024 + 0.003
50 2 0.011 + 0.004
60 6 0.014 ± 0.001
60 4 0.006 ± 0.001
60 2 0.011 + 0.002
70 6 0.004 ± 0.001
.70 4 0.003 + 0.002
70 2 0.000 + 0.0001


127

Appendix D: Effect of Air Temperature and Thickness on


Course of Drying of Potato Slice.

0 30 60 90 120 150 180 210 240


Time, mm.
— 50 dejiees celcms 60 degiees celcius
70 degiees celdus

Air llow rale |kg/ml s): 0 0054


Slice thickness |mm): 2

Figure D.l Effect of air temperature on drying time of 2


mm thick slices.


128

--- 50 degieps celcius 60 degien celcius


70 degrees celcrus

Air How rale (kg/m*5|: 0.0108


Slice m e t ness |mm|: 2

Figure D.2 Effect of air temperature on drying time of 2


mm thick slices.


129

50 degrees celcrus — 60 degiffs cflcius


70 degrees celcius

An llow tale |kglm*s|: 0 0108


Slice thickness |mm): 4

Figure D.3 Effect of air temperature on drying time of 4


mm thick slices.


130

—— 50 degrees celcies —*— 60 degrees celcms


70 degrees celcms

An How tale |kg/mas| 0 0054


Slice thick ness |mm)' 6

Figure D 4 Effect of air temperature on drying time of 6


mm thick slices.


131

0 30 60 90 120 150 100 210 240 270 300 330 360 390
Time, mm
~~ 50 degrees celcus — 60 degrees lelcms
70 degrees celcus

A n Mow lale (tg/rrfS) 0 0108


Slice thickness |mm| 6

Figure D.5 Effect of air temperature on drying time of 6


mm thick slices.


132

Time, min
' 2 mm •— 4 mm — 6 mm
Air lemperatuie |°C) 50
Air llow rale |kg/m*$): 0.0108

Figure D.6 Effect of Slice thickness on drying curves for


50°C air temperature.


133

Time, mm
— 2 mm — A mm — 6 mm
Air tempeialuie ft): 60
Aiillow rale (kg/rriS) 0.0054

Figure D.7 Effect of Slice thickness on drying curves for


60°C air temperature.


134

— 2 mm — 4 mm —* — 6 m m

Ai r t e m p e r a l u r e ( x r 70
Ai r How rale ( kg/ nrs ) 0.0054

Figure D.8 Effect of Slice thickness on drying curves for


70°C air temperature.


135

Appendix E: Arrhenius plots illustrating influence of


temperature on rate constant, k.

2mm, 0 0108 1g/mSs

Figure E.l Arrhenius plot illustrating influence of


temperature on rate constant, k for 2 mm thick slices..

-®- 4jtvt\ 0 0108 lglm*i

Figure E.2 Arrhenius plot illustrating influence of


temperature on rate constant, k for 4 mm thick slices.
136

Reaction tale constant, 1C, Mmia.jlog)

6mrr\, 0 0108 kglrr^s

Figure E.3 Arrhenius plot illustrating influence o f


temperature on rate constant, k for 6 mm thick slices.


137

Appendix F: Raw Data used for Production of Figures in


Main Text
(all values stated are an average of three replicates).

Table F.l: Raw data for production of fig. 4.4, 4.5 and 4.6.

Residence Moisture
time content
(minutes) (db)
Fig. 4.4 Fig. 4.5 Fig. 4.6
0 4.71 3.67 4.64
30 3.48 3.15 2.83
60 2.15 2.42 1.71
90 1.12 2.25 0.17
120 0.073 2.06 0.14
150 0.058 1.79 0.10
180 0.055 1.43
210 0.048 1.05
240 0.043 0.73
270 0.59
300 0.42
330 0.35
360 0.14

Table F.2: Raw data for figure 4.12.

Residence % Retention
time of ascorbic
(minutes) acid
0.0054 kg/m2s 0.0108 kg/m2s
air flow rate air flow rate
0 100 100
20 84.4 55
40 77.5 42.3
60 68.1 33.9
80 48.7 24.8
100 40.3 20.0
120 38.7 17.9


138

Table F.3: Raw data used for production of figures 4.13 and
4.14.

Residence % Retention
time of ascorbic
(minutes) acid
Figure 4.13 Figure 4.14
0.0054 0.0108 0.0054 0.0108
kg/m2s air kg/m2s air kg/m2s air kg/m2s air
flow rate flow rate flow rate flow rate
0 100 100 100 100
30 69.1 70.3 63.5 66.8
60 57.0 45.6 40.2 38.3
90 50.5 37.4 25.7 28.7
120 45.8 31.2 20.6 25.5
150 39.8 27.5 16.8 22.7
180 35.7 25.8 14.4 20.0
210 30.3 24.0
240 25.2 21.8
270 23.5 17.3
300 22.2 14.0
330 18.8 11.5
360 17.2 7.9
'390 15.0

Table F.4: Raw data used for production of figure 4.18.

Residence time (minutes) Absorbance of colour extract


(A420 nm/lOg slices)
20 0.036
40 0.053
60 0.102
80 0.195
100 0.137
120 0.105


139

Table F .5s Raw data used for production of figure 4.19.

Residence time (minutes) Rehydration ratio (g H20/g


dry matter)
30 0.59
60 0.76
90 1.00
120 1.13
150 1.33
180' 1.40
210 1.52
240 1.62
270 1.71
300 1.73
330 1.67
360 1.55
390 1.49

Table F.6: Raw data used for production of figure 4.20.

Residence time (minutes) Rehydration ratio (g H20/g


dry matter
20 0.82
40 2.43
60 4.11
80 4.39
100 4.04
120 3.74


140

Source: Cri ssman. 1989.


141

UNIVERSITY OF NAIROBI LIBRARY

National Parks & Reserves


Main road
Secondary road
Lakes
G Land higher than 2 1 0 0 m

Map 2.
Detail of primary potato production areas
Suurco: Cr iunman, 19(30.

S-ar putea să vă placă și