Você está na página 1de 28

CONVECTIVE SOLAR DRYER WITH A WOOD WASTE BACKUP HEATER FOR

DEHYDRATION OF FOOD
Madhlopa, A. and Ngwalo G.
University of Malawi The Polytechnic, Private Bag 303,Chichiri, Blantyre 3

ABSTRACT
Appropriate technology for the conversion of solar radiation to thermal energy is vital for food
dehydration. Commercial producers of dried foods need reliable drying technologies for
continuity technologies for continuity of production. However, solar radiation is not available
whenever it is needed to drying. This limitation discourages many of these producers from
investing in solar technologies, including dryers. Consequently, it is necessary to improve the
system design to promote the application of solar dryers.
In the present study, a natural convection solar dryer with an integrated thermal mass and
sawdust backup heater has been designed and constructed, the dryer was tested in three modes of
operation by drying twelve batches of fresh pineapples (ananas comosus): solar, biomas and
solar-biomass, meteorological conditions were monitored during the dehydration process and
both fresh and dried pineapple slices analyzed for vitamin C and ash content.
Results show that the solar mode of operation was slowest (five days) in drying the samples,
with the solar biomass mode being fastest (3 days) under the prevailing meteorological
conditions (which were generally unfavorable from January through July). Samples were
successfully dried even under rainy conditions, with moisture content dropping from 669% to
11% (on dry basis db) when the dryer was operated in the solar-biomass mode. This level of
moisture content is suitable for safe storage and distribution of the dried fruit. It was also found
that 26% to 44% of the vitamin c was retained in the dried product, ash content varied between
0.5% and 0.6% (on wet bassis, wb), with no significant difference between ash content (db) in
fresh and dried samples, this shows that there was very little or no food contamination arising

from dust. The final-day efficiencies of the system were 15+1, 11+ 1 and 13+2% for the solar,
biomass and solar-biomass modes of operation respective.
It is evident that the inclusion of a biomass backup heater enables drying to proceed under all
weather conditions. It appears therefore that this solar dryer overcomes the limitation of
intermittent dehydration of a food product. However, there is need for further work to optimize
the performance of the system.

INTRODUCTION
Dehydration is a common technique for the preservation of agricultural and other products,
including fruits and vegetables .Techniques for drying including openair, fuelfired, electric
and solar dryers. The use of open-air drying is cheap but it often results in food contamination
and nutritional deterioration (Ratti and Mujumdar, 1997). Fuel-fired dryers (such as coal-or
wood-operated dryers) contribute to environmental degradation through deforestation and or
emission of gaseous and particular matter (UNEP, 1988: IPCC, 1995). The use of electric-dryers
is limited to areas where there is an electric power supply. In contrast, solar energy is not only
friendly to the environment but also available even in remote areas. Consequently, solar dryers
can be used in both urban and rural areas.
Numerous designs of solar dryers with or without thermal storage systems are reported in
(Chirarattananon et al. 1998; Goyal and Tiwari, 1997; Ayensu, 1997: Aboul-Enein et al., 2000;
Itodo et al., 2002, Madhlopa et al., 2002 and others). Thermal storage systems enable drying to
continue after sunset provided there is enough sunshine during the day. However, the intensity of
solar radiation is sometimes so low that the temperature of the thermal mass rises by a very small
(or no) margin above the ambient level. So, this still limits the continuity of the drying process in
a solar dryer with a thermal mass. It should be mentioned that discontinuity of the drying process
is worse off in a solar dryer without a thermal storage system. This reduces the rate of production
and can results in low quality of the solar-dried products.
Few studies have been conducted to overcome the limitation of the exclusive use of solar energy
in natural convention solar dryers. Bassey et al. (1987) used a sawdust burner to provide heat
2

during bad weather and at night. In their study, the burner was not integrated to the dryer but
used steam as a heat transfer medium. Bena and Fuller (2000) designed a direct solar with an
integrated biomass backup heater. Their biomass burner performed well when operated on large
pieces of hard wood.
Fuelwood is a dominant source of energy, and commonly burned using inefficient technologies
in most developing countries (Kristoferson and Bokalders, 1991; Bena and Fuller, 2002).
Unfortunately, the heavy and inefficient consumption of fuelwood is contributing to
deforestration and other environmental problems (Kristoferson and Bokalders, 1991; Hyde and
Seve, 1993). It should also be noted that the use of large pieces of wood requires felling of trees
for the task, which aggravates the problem of deforestation. In contrast, the use of waste from
timber processing as a source of thermal energy does not require felling of trees specifically for
fuelwood. This approach can therefore contribute to a sustainable exploitation of wood resources
in many developing countries.
The solar dryer reported by Bena and Fuller (2002) does not have a thermal storage system for
capture solar radiation. In view of this, their system requires backup thermal energy from the
burner for over-night dehydration of the fresh food even when solar radiation is abundant during
the day. The air temperature in the drying chamber would fall to ambient level immediately after
sunset in the absence of backup heating. In this case, the drying food has to be removed from the
dryer everyday at sunset and re-loaded into the system in the morning, to avert moisture reabsorption by the drying food product at night. This is inconvenient for a commercial producer
of dried foods.
In Malawi, open-air drying is the traditional method for preserving fruits and other fresh food
products. However, industrial type solar dryers have been applied to the drying of tobacco,
coffee, tea and fish but with limited success largely due to financial constraints and insufficient
solar data for designing appropriate crop dryers (Kafumba, 1994). Mumba (1995) designed a
photovoltaic forced-convection solar dryer for drying grain in rural areas where grid electricity
and fossil fuels are generally not available. He found that it was possible to construct a costeffective solar grain dyer using locally available materials in Malawi. More recently Madhlopa et
3

al. (2002) developed a solar dryer with composite-absorber systems for dehydration of food. The
solar collection of this dryer had a removable metallic absorber systems for dehydration of food.
The Madhlopa et al. (2002) developed a solar dryer with composite-absorber systems for
dehydration of food. The solar collector of this dryer had a removable metallic absorber plate and
a fixed wooden bottom. This design provided novel flexibility in the adjustment of the thermal
characteristics of the dryer. It was found that the dryer was suitable for preservation of mangoes
(Magifera indicus) and other fresh foods. However, both of the dryers reported by Mumba
(1995) and Madhlopa et al. (2002) did not have thermal storage and backup heating systems.
This would limit their application in poor weather.
The main objective of the present study was to develop a solar dryer with a thermal mass and
backup sawdust burner, with the following specific objectives:
a. to design and construct a solar dryer for dehydration of fruits ,
b. to determine the efficiency of the dryer, and
c. to evaluate the quality(moisture content, ascorbic acid and ash content ) of the dried food..
An indirect solar dryer has been designed and constructed with a biomass backup burner (fig. 1)
the biomass burner was made of a drum fitted with a mild steel grill and horizontal baffle. The
grill was fitted 0.20m above the bottom part of the drum while the baffle was 0.06m above the
grill. The drum was integrated to a rectangular duct with 3 vertical baffles to increase the path
length of the flue gas and the amount of heat transferred to the drying chamber.
A truncated circular removable lid was lightly-fitted on the open end of the drum through which
biomass was loaded into and ash removed from the drum, with a rectangular steel door (with
vermiculite-filled cavity, 0.025m thick) fitted on the outer side of the lid, the door was perforated
(total area of perforation = 3.55 x 10-3 m) on the bottom part for inlet of air into the combustion
chamber (drum). The area of the perforation was estimated based on the volume of air required
to burn 8kg of biomass in about 12 hrs (Ashrae, 2001). It was necessary for the combustion to
proceed for about 12 hrs to ensure that plenum temperature was higher that the ambient
temperature, and thereby sustain the drying process from sunset to sunrise the next day around
4

steel bars (0.016m diameter) were placed across the drum and duct at 0.3 m apart, and an
expanded metal placed on top of the bars to enhance absorption and storage of thermal energy,
granite rock pebbles (total mass of 360 kg, about 0.025 m pebble diameter) were packed on top
of the expanded metal, with the top part of the rock layer covered with concrete (0.025 m thick)
to make hot air from the burner chamber infiltrate through the horizontal rock matrix before
rising into the drying chamber via a rectangular hole (1.00 x 0.09 m) at the front of the concrete
absorber. The top part of the absorber was painted matt black to increase radiation absorption.
Granite was used for thermal storage because it has a high thermal diffusivity of 1.27 x 10-6 m s1

(Ayensu, 1997).

This flow pattern of hot air enables a reasonable amount of heat to be transferred to the thermal
mass (calculated thermal mass = 1.146 mj k-1). It was found necessary to have a relatively small
thickness (0.10 m) of the horizontal rock-and-concrete structure due to consideration of support
strength the flue gas exited the burner through a chimney (constructed from a galvanized iron
sheet, 6 x 10-4 m thick) which was fitted to the duct, a hole was drilled at the bottom of the knee
bend into the chimney to drain out flue-gas condensate from the system (Ashrae, 2001).

Fig. 4: A) Plan View Of The Biomass Burner, Showing Flow Of Flue Gas The Drum Through
The Rectangular Duct Into The Flue Gas Chimney, B) Longitudinal Section Of The Drum.
8

The dryer has a drying cabinet mounted on a brick wall that encloses a biomass burner. The brick
wall around the burner has cavity (0.05 m thick) which is filled with vermiculite to reduce heat
loss through the wall. The exterior dimensions of the wall are 2.52mby 1.55m drying air enters
the system through a rectangular inlet (1.00 m x 0.09m) at the bottom part of the brick wall, gets
into contact with the bare part of the drum and rises into the drying chamber through natural
convention (forming an s-shaped flow pattern). It passes through the drying bed (of thickness x =
0.3m) and exits through a solar chimney (painted matt black, height = 1.2 m, diameter = 0.18 m)
or air outlet vents. The solar chimney is fitted on top of the drying chamber to augment
thermosiphoning during solar collection, with a pressure difference (_p) given by (Ayensu, 1997)
P= g (_-_)(h1 + h2)
Where g is the acceleration due to gravity, _ and _ are air densities at the inlet (ambient) and
inside the dryer, and h1 and h2 are heights as shows in fig. 2. At the design meteorological
conditions (table 1), the pressure difference arising from natural convention is _p = 1.04 n m-2.
The drying chamber has three trays, each with a plastic mesh base, which slide horizontally
along wooden rails fixed to the vertical sides of the cabinet. These trays can be removed for
loading and cleaning the effective tray area 4.1m2 and accommodates about 20 kg of fresh
pineapple the top surface of the cabinet is opaque, and inclined at 16 to the horizontal to reduce
friction as the air rises, under natural convention, from the plenum through the drying bed, the
collector frame and drying cabinet were constructed were constructed from block board (0.02 m
thick) and covered with a painted galvanized iron sheet (6 x 10-4 m thick) to protect the wooden
components from weathering. Two wooden doors (1.20 m x 1.51 m) were fitted at the back of
the drying cabinet for accessing the drying chamber. Each door has an air vent (rectangle of 0.51
m by 0.10 m with semicircles on both short vertical sides) located at its top. A wire mesh was
fitted on the vents and collector inlet. An overhang was fitted over the outlet-air vents to prevent
rain drops from entering the drying chamber through the vents. A glass cover was fitted on the
top part of the collector and inclined at 16 to the horizontal to optimize solar collection at the
malawi polytechnic (15 48s, 35 02 e) where the system was tested. The dryer faced north,
with the flue gas chimney on the western side of the system, the drying cabinet was fixed to the
9

brick wall with metal straps mortared to the top layer of the external layer of the brick wall (figs.
1-3).
Table 1: Design parameters for a solar dryer with biomass back-up heater.
dryer

drying load and meteorological conditions

solar collector

drying products

aperture, a = 2.2m2

initial mass of load, mi = 20.0 Kg

absorber: concrete slab

initial moisture, mi = 85% (wet basis)

glass cover

final moisture, mf = 10% (wet basis)

thickness = 3mm
tilt angle = 16

meteorological conditions

concrete absorber=0.025m thick

solar radiation, h = 21.0 mj m-2 day-1

thick granite rock = 360 kg

inlet air temperature = 30 c:


plenum temperature = 40 c

drying chamber

relative humidity, rh = 80%

effective tray area = 4.1 m2


solar chimney = 1.2 height, 0.18 m
0.18 m
biomass burner
drum
length = 0.89 m
diameter = 0.0.58 m
door = 0.55 m x 0.54 m
flue gas chimney
height = 2.12 m
diameter = 0.12 m

10

EXPERIMENTATION

Sample preparation
Twelve batches of fresh pineapples (annas comosus) were procured from Limbe produce market.
These fruits were washed, peeled, sliced (disc-shaped slices, thickness of about 1 cm) and cored.
For each batch, the slices were weighed on digital top-loading balance (adam equipment co.,
model acw-6) and loaded into the dryer for dehydration.

Operational modes
The dryer was operated in three different modes: solar, biomass and a combination of the two
energy sources. For the solar mode of operation, a batch of fresh pineapple slices was weighed
(on a balance), loaded into the dryer in the morning and re-weighed after 24 hrs and on the final
day of drying. For the biomass mode of operation, drying of a batch of pineapple slices started in
the evening. The aperture of the dryer was covered with a ceiling board (a black polythene paper
was pasted on the top part of the board ) which was parallel to and extended by 0.6m beyond
the perimeter on all sides of the grass cover, with an air gap of 0.6m between the board and
transparent cover (ashrae, 1998). To block radiation when the altitude of the sun was low, a
vertical removable ceiling board (wrapped in black polythene paper) was fitted on the eastern
side in the morning and western side of the collector in the afternoon. The location of this shield
was changed around solar noon, the burner was charged with 8 kg of sawdust or wood shavings
and ignited around sunset. Again, the batch was re-weighed after 24 hrs and on the final day of
drying. Due to lack of automated facilities for recording temperatures at night, the burner was
charged with biomass fuel at sunrise, with the solar aperture covered. For the mixed mode, the
burner was again charged once with the biomass fuel around sunset and re-weighed after 24 hrs
and on the final day of drying.

Metrological parameters
Twelve mercury-in glass thermometers were inserted horizontally into the dryer (six on the
eastern side and six on the western side of the drying chamber). On each side, three of the
thermometers measured the plenum air temperature (below the drying bed) while the other three
measured the temperature of the air above the drying bed (fig 2). Ambient air temperature was
11

monitored by using a mercury-in glass thermometer placed in a room with louvered glass. The
louvers were kept open to allow free circulation of air. The flue gas temperature was also
monitored using a mercury in glass thermometer. A wet and dry bulb thermometer (kagaka
kyoeisha, model sllb) was used to measure the level of relative humidity while velocity was
measured by a casella low-speed air meter (n 1462). It was also found necessary to determine the
flow rate of drying air through the dryer in this case, the air meter was placed in the middle of
the air inlet to the drying chamber. This location was suitable for comparability of the
performance of the solar chimney and the vents on the dryer. One venting system was blocked
while measurements were taken using the other. The intensity of global solar radiation was
measured by a kipp & zonen pyranometer (cm 6b) mounted in the plane of the transparent cover,
and connected to a kipp & conen solar integrator (cc 14).

Physico-chemical analysis
The moisture content of fresh and dried samples was determined by drying an accuratelyweighed (5 g) sample in an electric oven (griffin, serial number 0518171070, accurate to 0.001
kg) at 70 c for hours (aoac, 1990). The loss in moisture was calculated as a percentage of the
mass of the dry sample. The uniformity of drying at different levels of the trays was determined
by weighing all the pineapple slices from each tray. In addition, five samples were taken from
different positions of each tray to establish the uniformity of drying across each tray (fig, 5)

12

Fig 5: position of five pineapple used for determination of drying uniformity across the bottom
tray.
Early studies indicate that vitamin c is particularly vulnerable to destruction under different
conditions (air, light and heat) and during processing (Drew and Ree. 1980: Fellows and
Hampton, 1992). So, the retention of vitamin c is used as a good indicator of the preservation of
all other nutrients (maeda and salunke, 1981).

Date

Day or Weather

Mean ambient

Mean relative

Daily solar

night

temperature (c)

humidity (%)

radiation(mj m-2)

1/6/04

Day

Partly cloudy with drizzles

17

78

16.6

1/6/04

Night

N.a

15

N.a

Nil

2/6/04

Day

Cloudy with Showers

16

76

8.4

2/6/04

Night

N.a

15

N.a

Nil

3/6/04

Day

Partly cloudy

16

77

21.6

3/6/04

Night

N.a

15

N.a

Nil

4/6/04

Day

16

82

13.5

Cloudy with showers

13

Air and flue gas temperature


The plenum air passes through the drying bed where it picks up moisture from drying food by
natural convention. Fig. 6 shows the variation of mean plenum and ambient air temperatures on
the typical days of drying. For the solar mode of operation, it is seen that the plenum and ambient
temperatures are almost equal in the morning but different during the day. In particular, the
plenum temperature is higher than the ambient temperature during the most part of the day. The
plenum-ambient temperature difference (excess temperature) was 17 c from 13:00 to 15:00 hrs
and 12 c at 17:00 hrs. This indicates that the dryer stored some of the captured solar energy for
drying to continue during periods of low insolation or at night. At a mean excess temperature 12
c at sunset, the thermal mass stored about 13.8 mj for food dehydration to continue at night.
However, the excess temperature dropped down to 0 c by the next morning. Leon et al. (2002)
report that a temperature difference (between ambient and drying air) of at least 10 c is required
for drying to continue. This shows that there was generally significant drying around solar noon.

a
Present investigation, we could not conduct a full nutritional evaluation of the food product due
to the limitation of the scope of this study, consequently, the retention of vitamin c was used as
an indicator of nutritional quality. The concentration of vitamin c in fresh (cf) and dried (cd)
pineapple samples was determined according to AQAC (1990), and expressed on dry basis ash
14

content was determined by incinerating an accurately-weighted sample (5 g) in a muffle furnace


(carbolite glm 1) at 600 c. The content of ash was expressed as a percentage of the mass of the
original sample.

Data processing
Stored thermal energy
Stored thermal energy (q) at sunset was estimated by using excess temperature (_t) in the
plenum. It should be noted that when the thermal mass is hot, the plenum air temperature is
lower than that of the heated thermal mass. This shows that the excess plenum temperature is
also lower than that of the heated thermal mass. Consequently, the stored energy computed in
this study is a minimum value on a given day.
q = m _t

(2)

Where m is thermal mass, and _t = ttm ta

retention of vitamin c
The percentage retention (r) of vitamin c was computed by:
r = 100 cd /df

(3)

Where cd is the concentration of vitamin c in dried samples on dry basis (mg)/100 g), and cf is
the concentration of vitamin c in fresh samples on dry basis (mg/100 g)

dryer efficiency
Drying efficiency varies with the nature and moisture content of the fresh product.
Unfortunately, there is lack of standard methods for testing solar dryers which adversely affects
their evaluation (Sodha et al., 1987), So, Leon et al. (2002) recommend the use of the first-day
and final day efficiencies for evaluation of drying efficiency. The efficiency () of the solar dryer
was computed
By using the equation reported by Bena and Fuller (2002) and Leon et al. (2002):

15

= wl/(ah + cm)

(4)

Where w is the mass of water evaporated (kg), 1 is the latent heat of vaporization (mj kg), a is
the aperture of the dryer (m), h is the insolation on the dryer (mj m), c is the calorific value of
biomass (mj kg-1) and m is the mass of biomass. In this study, the value of c were determined
using gallenkamp automatic adiabatic bomb calorimeter (cat. No. 15 cb 110) and corrected for
moisture content. The latent heat of vaporization was calculated using a temperature dependent
function reported by Jannot and Coulibaly (1998):
1=2.5018-0.002378 twp

(5)

Where twp is the wet bulb temperature (c) of the drying in the plenum. This temperature was
estimated using the procedure recommended by ASHRAE (1994).

RESULTS AND DISCUSSION

Weather conditions
Table 2 shows average weather conditions during the dehydration of a typical batch of fresh
pineapples slices. The weather was very unfavorable on most of the days, dominated by cloudy
skies with occasional drizzles and showers. The levels of relative humidity were relatively high,
ranging from 76% to 82%, with mean ambient air temperatures varying between 15 c and 17 c.
It is seen that the daily global solar radiation on an inclined plane (16 c to the horizontal) was
variable, with values ranging from 8.4 mj m to 21.6 mj m. So, the daily intensity solar
radiation was lower than the design value (21.08.4 mj m) on most days. These weather
conditions indicate that drying would not proceed quickly, resulting in spoilage of the pineapple
slices, if a biomass backup heater was not used. However, the daily intensity of solar radiation
was quite high on some days (typically 30.7 mj m on 22nd November 2004), which shows the
potential for using the dryer without backup heating under such weather conditions.

16

Table 2: weather conditions during one of the tests of the loaded solar dryer.

c
17

Fig 6: variation of mean plenum (tpm), flue and ambient air (ta) temperature with local time on
typical days for different modes of dryer operation: a) solar energy on 28th July, 2004, b)
biomass on 30th November, 2004 and c) solar-biomass on 24th February, 2004.
For the biomass mode of operation, the plenum temperature was initially monitored without fire
in the burner. It was found that there was insignificant differences between the plenum and
ambient temperatures during the day, which indicated that the insulation cover over the solar
collector was effective in blocking off irradiation during the day. The burner was charged with
sawdust or wood shavings. It was observed that the combustion of wood shaving was more
satisfactory than that of sawdust. Consequently, wood shavings were used in the rest of the
burner tests. With the burner in the morning, it is observed that the plenum temperature is higher
than ambient temperature immediately after 06: 00 h to 17:00 h with a maximum value of 41 c
at 12:00 h on a typical day, which is within the acceptable limits of the drying air temperature
(Bena and Fuller, 2002; leon et al., 2002). There is an excess temperature of atleast was 10 c at
17:00 h. This is attributed to the fact that the wood shavings burnt for about 10 hours with the
stored thermal energy being released into the plenum even after the completion of combustion.
In addition, it is observed that the flue gas temperature has a distinct maximum (53 c) at 09:00
h; which is relatively low. This indicates that most of the thermal energy was transferred to the
system before the gas exited the rectangular duct. Bena and Fuller (2002) report a flue gas an
excess temperature of 177 c, which was high. When the burner was charged with wood shaving
at sunset, there was an excess temperature of 6 c to 10 c the next morning but it dropped to zero
by sunset. This indicates that drying proceeded at night.
For the solar-biomas mode of operation, it is observed that the plenum temperature was higher
than ambient temperature from 06:00 h to 17:00 h, with the excess temperature increasing to
maximum value of 18 c around 09:00 h and 10:00 h, then dropping to 7 c at 17.00 h..
Moreover, the burner was re-charged with wood shavings at sunset. So, food dehydration
proceeded during both day and night.

18

Drying time and uniformity


For the solar mode of operation, it was found that the dryer reduced the moisture content of a
typical batch of fresh pineapple slices from 669 % to 16% (on dry basis, db) in 5 days, with the
daily intensity of global solar radiation varying from 18.4 to 26.8 mj m on an inclined plane
(16 to the horizontal). It should also be mentioned that the drying product was not removed
every day from the dryer after sunset because the stored thermal energy enable drying to proceed
at night. So, the dryer can be exclusively powered by solar energy to dehydrate a fresh food to
the required level of moisture under favourable weather conditions.
For the biomass mode of operation (with a non-irradiated collector), it was observed that the
dryer reduced the moisture content of a batch from 614% (db) to 13% (db) in 3 days. Prolonging
the dehydration process to 3.5 days resulted in an over-dried product. This clearly shows that the
rate of drying in this mode of dryer operation is higher than that of the solar mode, which is
attributed to the higher amount of energy supplied by the wood shavings (17.6 mj kgr)
radiation.
The solar-biomass mode of operation was used to study the drying time and uniformity. Fig. 7
shows the variation of the mass and moisture content of a typical batch. It is observed that the
rate of drying was highest during the first day (24 hours), with mass of the batch dropping from
20.02 kg to 12.28 kg. The mass of the sample was 2.90 kg by the end of the third day, with
negligible drying after this. It is seen that the moisture content dropped from of 669% to 11%
(db) in 3.0 days, with the intensity of the daily global solar radiation varying from 15.2 mj m to
21.5 mj m on an inclined plane. Lower final levels of moisture content can be achieved in
shorter drying periods with higher intensities of global radiation. The solar dryer developed by
Bena and Fuller (2002) reduced moisture content of pineapples from 559% (db) to 11 % (db) in
3.5 days, under the prevailing meteorological conditions. Leon et al. (2002) recommended a final
moisture content of 10% (wet basis, db), which is equivalent to 11% (db) for evaluation of solar
dryers. So, the final moisture level obtained after drying a batch of pineapple slices for 3 days
was within acceptable limits when the solar dryer was backed by a biomass burner.

19

Fig 7: variation of moisture content (%), db) and mass of a typical batch of pineapples (dried
from 21st to 24th June 2004) with drying time (day)
Figure 8 shows the rate of drying at different trays in the drying chamber. It is seen that samples
in the bottom tray dried fastest, with the top tray exhibiting the least drying rate. This drying
pattern is attributed to the flow of hot air which flows from the collector (in the bottom part) of
the dryer into the drying chamber in the middle part. Consequently, it is necessary to swap the
trays in order to ensure uniform drying.

20

Fig 8: variation of moisture content at different trays (28th June -1st July 2004)
The uniform of drying across the bottom tray is shown in fig 9. It is seen that samples from
positions a and b (in front of the tray) dried fastest with those at positions c and d (back of a tray)
exhibiting the least drying rate. The middle sample (position o) exhibited an intermediate rate of
drying. A similar pattern of drying uniformity was observed across the middle and top trays, with
the drying rate being most non-uniform across the top tray. These observations are attributed to
the flow pattern of the heated air which moves upwards and horizontally across the trays A
similar pattern of air flow is expected for the solar biomass operational modes. So, the front part
of the bottom tray receives the hottest air.

21

Fig 9 uniformity of drying rate across the bottom tray (28th June 1st July 2004)

Concentration of vitamin c and ash


The concentration of vitamin c varied from 22 to 34 mg/100g (wb), with a mean value of 26+4
mg/100 g (wb). These results are in close conformity with the findings (25 mg/100 g, wb) of
Osborne and Voogt (1978). The retention of vitamin c varied from 26 to 44 %, probably due to
the high sensitivity of this type of vitamins to different processes (drying, sample preparation)
and conditions (Drew and Ree, 1980; Fellows and Hampton, 1992). Ash content varied between
0.5% and 0.6% (wb) which compares well with the value of 0.4% (wb) reported by Osborne and
Voogt (1978) for raw pineapples. The minor vitamins are probably due to the season and the
maturity stage of the fruit. There was no significant difference in the concentration of ash (db) in
fresh and dried samples, which indicates that there was very little (or no) contamination arising
from dust.

Air flow rate


For diurnal operation with the loaded dryer, the estimated air flow rate varied from 0.017 m3 s-1
to 0.036 m3 s-1 (mean value of 0.024 + 0.002 m3 s-1) when the dryer was operated with solar
22

chimney, and from 0.030 m3 s-1 to 0.048 m3 s-1 (mean value of 0.037 + 0.003 m3 s-1) when dryer
was operated with the air outlet vents. The two mean values of estimated flow rate were
significantly different (p-value = 0.00). It is seen that flow rate is slightly lower when the dryer
was operated with a solar chimney than air outlet vents. This is attributed to the pressure drop
between the dryer air inlet and the chimney air which is higher than that between the dryer air
inlet and the air vents. Moreover, samples provide resistance to air flow. It was also observed
that the air meter advanced one direction when the system was operated with a solar chimney,
which indicated that there was little or no reverse flow in this mode of operation. In contrast, the
air meter could occasionally stop and then advance in the opposite direction when the system
was operated with air outlet vents open. Bena and Fuller (2002) report an average air flow rate of
0.038 m3 s-1 for a solar dryer operating in the biomass mode. These authors found considerable
variation in the flow rates for their dryer which had air outlet vents.
For nocturnal operation with an open solar chimney, it was observed that there was condensate
on a small portion of the absorber area directly below the air inlet to the chimney. This was
attributed to reverse flow at night when the chimney became cold and acted as a condenser for
the warm (and humid) air rising up from the drying bed. In view of this , the solar chimney was
blocked with card board papers (with vents opened) at sunset and opened (with vents closed) at
sunrise to avert reverse thermosiphoning. This approach proved to be effective.
It should also be mentioned that the flow rate of air through a solar chimney is affected by
several factors which include intensity of solar radiation, chimney length and cross-sectional area
(Ong and Chow, 2003). These authors found that air velocity increased with solar radiation and
air gap. It has been seen that air velocities produced by the size of the solar chimney in this study
are lower than those of the outlet air vents. Moreover, a solar chimney increases the production
cost of the dryer and is effective only during the day. Consequently, the dryer can still operate
satisfactorily with only air outlet vents.

Dryer efficiency
Table 3 shows the first-day (1) and final-day (f) efficiencies of the dryer. For the first-day
efficiency, it is seen that the solar mode of operation exhibits the highest mean value, with the
23

biomass mode showing the lowest value. This is probably due to the fact that the burner loses a
relatively high proportion of the heat though the flue gas, which results in a relatively lower
efficiency than that for the solar mode of operation. For the final day efficiency, it is again
observed that the solar mode of operation is highest with the biomass mode being the lowest.
Table 3: variation of the first () and final (f) day efficiencies.
Operational mode

Efficiency
(%)

(%)

Solar

20 + 2

15 + 1

Biomass

15 + 2

11 + 1

Solar-biomass

17 + 1

13 + 2

The calculated final efficiency of the solar mode of operation is comparable with typical
efficiency values of 10 15% for natural convention solar dryer (brenndorfer et al)., 1985). Bena
and Fuller (2002) report an efficiency value of 22% when their dyer was operated in the solar
mode. These authors state that this value is likely to be inflated because of the value of the area
of the solar aperture. Moreover, these authors to not report the first-day efficiency of their
system. Leon et al. (2000) states that the first-day efficiency may be taken as a consistent basis
for comparing the thermal performance of solar dryers. For the solar-biomass mode of operation,
it seen that the present calculated efficiency is higher than (8.6%) found by Bena and Fuller
(2002). This is attributed to the inclusion of a rectangular flue gas duct and thermal mass (with
granite rock pebbles) which increases the rate of heat transfer from the burner to the drying
chamber in the present dryer design. For the biomass mode of operation, the calculated
efficiency is again higher than that (6%) reported by Bena and Fuller (2002), for the same
reasons suggested for the soar-biomass mode of operation.

24

CONCLUSION AND RECOMMENDATIONS


A convective solar dryer has been designed and constructed with a biomass backup heater. The
system was tested in three modes of operation (solar, biomass and solar-biomass), using fresh
pineapple slices, under different weather conditions. Results show that under favorable weather
conditions, it was possible to dry pineapple slices using solar energy as the only source of
thermal energy. The solar mode of operation yielded the highest thermal efficiency, with the
biomass mode of operation exhibiting the lowest efficiency. However the dryer performed most
satisfactorily with a backup heater (operated on wood shavings), yielding an overall efficiency
13+2%, and retaining 26% to 44% of vitamin c during the dehydration process. It appears that
this solar dryer overcomes the limitation of intermittent dehydration of a food product. However,
there is need for further improvements to the design to optimize the performance of the system
and its assessment at pilot phase.
Nomenclature
A

solar aperture of the dryer (m)

calorific value of biomass (mj kg)

Cd

concentration (db) of vitamin c in solar-dried sample (mg/100 g)

Cf

concentration (db) of vitamin c in fresh sample (mg/100 g)

daily intensity of global solar radiation (mj m )

latent heat of vaporization of water (mj kg)

thermal mass (mj c)

mass of biomass (kg)

stored thermal energy (mj)

retention of vitamin c (%)

Ta

ambient temperature (c)

Twp

mean plenum temperature (c)


25

Twp

wet bulb temperature of the drying air in the plenum ( c)

mass of water evaporated (kg)

Greek letters
-

efficiency of dryer

air density (kg m )

Subscripts
A

ambient

Db

dry basis

Wb

wet basis

ACKNOWLEDGMENTS
The authors are very grateful to National Research council of Malawi for the financial support.
The Malawi Polytechnic is also acknowledged for the logistical and technical support. This work
has enabled the researchers gain some practical experience in the design, construction and testing
for solar dryers with biomass backup heaters.

REFERENCES
Aboul-enein, s., ei-sebaii, ramadan, m.r.i., el-gohary, h.g., 2000, Parametric study of a solar Air
heater with and without thermal storage for solar drying. Renewable energy 21,505522.
Aoac, 1990. Official methods of analysis of the association of official analytical chemists, Vol.2
Aoac, inc., virginia
Ashae, 1988. Methods of testing to determine the thermal performance of solar domestic water
Heating systems, standard 95-1987. American society of heating, refrigerating and
air-conditioning engineers, Atlanta.
26

Ashrae, 1994. Standard method for measurement of moist air properties, standard 41.6-94.
American society of heating, refrigerating and air-conditioning engineers Atlanta.
Ashrae 2001. Fundamentals handbook. American society of heating, refrigerating and airConditioning engineers, Atlanta, pp. 18.1-18.16
Ayensu, a., 1997. Dehydration of food crops using a solar dryer with convective heat Flow. Solar
energy 59,121-126.
Chirarattananon, s., chinporncharoenpong, c., chirarattananon, r., 1988. A steady-state model for
the forced convection solar cabinet dryer. Solar energy 41,349-360.
Bassey, m.w., whitfield, m.j. korama, e.y., 1987 problems and solutions for natural Covention
solar crop drying. In solar crop drying in Africa proceedings of food drying
workshop, dakar bassey m.w. and schmidt o.g. (eds), idrc, Ottawa, Canada.
Bena, b., fuller, r.j., 2002. Natural convention solar dryer with biomass backup Heater. Solar
energy 72,75-83
Brenndorfer, b., kennedy, I., baterman, c.o., trim, d.s., mrema, g.c., wereko-brobby, c., 1985.
Solar dryers their role in post harvest processing. Commonwealth science council,
London.
Drew, f., ree, k.i.s., 1980. Energy use, cost and product quality in preserving vegetables at home
by canning, freezing and dehydration. Food science 45,1561-1565. Fellows, P.
Hampton A., 1992. Small-scale Food Processing: A Guide to Appropriate
Equipment. Intermediate Technology Publications, London.
Goyal, R.K. Tiwari, G.N., 1997. Parametric study of reverse flat plate absorber cabinet dryer: A
new concept Solar Energy 60, 41-48..
Hyde, W.F., Seve, J.E., 1993. The economic role of wood production in tropical deforestation:
The severe example of Malawi. Forest Ecology and Management 57,283-300.
IPCC (Intergovenment Panel on Climate Change), 1995, Climate Change 1994, Cambridge
Press, Cambridge, pp.15-34.
Itodo, I.N., Obetta, S.E., Satimehin, A.A., 2002. Evaluation of a solar crop dryer for rural
applications in Nigeria, Botswana Journal of Technology 11,58-62.
Jannot, Y. and Coulibaly,Y., 1998.The evaluation capacity as a performance index for solardrier air heater Solar Energy 63, 387-391.

27

Kafumba, C.R., 1994. status of renewable energy technologies in Malawi. A paper presented at
the SEI/AFRENPREN/FWD Workshop at Naivasha, Kenya.
Kristoferson, L.K Bokalders, V., 1991. Renewable energy technologies: Their applications in
developing countries Intermediate Technology Publications, Southmpton, pp. 3-19.
Leon, M.A., kumar, S., Bhattacharya, S.C., 2002. A comprehensive procedure for performance
evaluation of solar food dryers. Renewable and Sustainable Energy Reviews 6,367 393
Madhlopa, A., Jones, S.A., Saka, J.D.K., A solar air heater with composite-absorber systems for
food dehydration. Renewable Energy 27, 27-37.
Maeda, E.E., Salunke, D.K., 1981. Retention of ascorbic acid and total carotene in solar dried
vegetables. Food Science 46, 288-1290.
Mumba, J., 1995 Economic analysis of a photovoltaic, forced-convection, solar grain drier.
Energy 20, pp. 923-928.
Osborne, D.R., Voogt, P. 1978. The Ananlysis of Nutrients in Foods. Academic Press, London,
pp.80-90, 166-167.
Ong, K.S., 2003. Performance of a solar chimney. Solar Energy 74, 1-17. Ratti, c. mujumdar,
a.s., 1997. Solar dryer of foods: modeling and numerical simulation. Solar energy 60,
151-157.
Sodha, M.S. Bansal, N.K. Kumar, A., Bansal, P.K. Malik, A.E. 1987. In solar crop drying, vol II,
CRC Press, Florida.
UNEP (United Nations Environment Programme), 1988. Environmental Perspective to the year
2000 and beyond. Penshurst Press Ltd, Kent, pp. 13-16

28

Você também pode gostar