Sei sulla pagina 1di 11

PII: S0043-1354(97)00295-9

Wat. Res. Vol. 32, No. 3, pp. 781791, 1998


# 1997 Elsevier Science Ltd. All rights reserved
Printed in Great Britain
0043-1354/98 $19.00 + 0.00

QUALITY OF WATER TREATED BY COAGULATION


USING MORINGA OLEIFERA SEEDS
M
ANSELME NDABIGENGESERE and K. SUBBA NARASIAH**

Department of Civil Engineering, University of Sherbrooke, Sherbrooke, Quebec, Canada J1K 2R1
(First received October 1996; accepted in revised form August 1997)
AbstractA model turbid water was treated by coagulation-occulation and sedimentation, with
Moringa oleifera seeds as a coagulant, using jar tests. The quality of the treated water was analyzed
and compared with that of the water treated with alum. Experiments were conducted at various dosages
of the crude 5% water extract of both dry, shelled and non-shelled Moringa oleifera seeds. Measurements of pH, conductivity, alkalinity, cation and anion concentrations, showed that coagulation with
Moringa oleifera seeds did not signicantly aect the quality of the treated water. However, concentration of organic matter in the treated water increased considerably with the dosage of Moringa solution. Since this organic matter might exert a chlorine demand and also act as precursor of
trihalomethanes during the disinfection with chlorine, it is suggested that Moringa oleifera seeds be
used as a coagulant in water and wastewater treatment, only after an adequate purication of the active
proteins. # 1998 Elsevier Science Ltd. All rights reserved
Key wordscoagulation-occulation, natural coagulants, Moringa oleifera, cationic proteins, vegetable
oil

INTRODUCTION

Coagulation-occulation followed by sedimentation,


ltration and disinfection, often by chlorine, is used
worldwide in the water treatment industry before
the distribution of treated water to consumers
(Edzwald et al., 1989; Kawamura, 1991a;
Degremont, 1989; AWWA, 1990; Desjardins, 1988;
Viessman and Hammer, 1985). Aluminum salts are
by far the most widely used coagulants in water
and wastewater treatment (Bratby, 1980). However,
recent studies have pointed out several serious
drawbacks of using aluminum salts, such as
Alzheimer's disease and similar health related problems associated with residual aluminum in treated
waters (AWWA, 1990; Miller et al., 1984;
Letterman and Driscoll, 1988; Qureshi and
Malmberg, 1985), besides production of large
sludge volumes (James and O'Melia, 1982). There is
also the problem of reaction of alum with natural
alkalinity present in the water leading to a reduction of pH (Degremont, 1989), and a low eciency in coagulation of cold waters (Haarho and
Cleasby, 1988; Morris and Knocke, 1984). A signicant economic factor is that many developing
countries can hardly aord the high costs of
imported chemicals for water and wastewater
treatment (Schultz and Okun, 1983, 1984;
Ndabigengesere, 1995).
*Author to whom all correspondence should be addressed.
781

Current research is oriented toward producing


more eective trivalent aluminum coagulants, such
as polyaluminum chloride (PAC) and polyaluminum silico sulphate (PASS) (Jolicoeur and Haase,
1989; Berrak, 1992). Although these new coagulants
have improved the coagulation process considerably, they have not corrected all the drawbacks
mentioned earlier.
Ferric salts and synthetic polymers have also
been used as coagulants but with limited success,
because of the same disadvantages as in the case of
aluminum salts (Bratby, 1980; Haarho and
Cleasby, 1988; Boisvert, 1992; Letterman and Pero,
1990; Goppers and Straub, 1976; Aizawa et al.,
1990; Biessinger et al., 1976; Biessinger and Stokes,
1986; Mallevialle et al., 1984). Therefore, it is desirable that other cost eective and more environmentally acceptable alternative coagulants be developed
to supplement if not replace alum, ferric salts, and
synthetic polymers. In this context, natural coagulants present a viable alternative (Kawamura,
1991b).
Natural coagulants of vegetable and mineral origin were in use in water treatment before the advent
of chemical salts, but they have not been able to
compete eectively because of the fact that a scientic understanding of their eectiveness and mechanism of action was lacking. Thus far, use of
natural coagulants has been discouraged without
any scientic evaluation. They have succumbed progressively under modernization and survived only in

782

Anselme Ndabigengesere and K. Subba Narasiah

remote areas of some developing countries


(Hespanhol and Selleck, 1975, Jahn, 1981).
Recently however, there has been a resurgence of
interest in natural coagulants for water treatment in
developing countries (Jahn, 1981, 1986, 1988).
A water soluble extract of the dry seeds of
Moringa oleifera is one of the natural coagulants
(Ndabigengesere et al., 1995). Moringa oleifera is a
tropical plant belonging to the family of
Moringaceae. Up to fourteen dierent species have
so far been identied, all of them possessing varying
degrees of coagulation activity (Jahn, 1988).
Moringa oleifera is the most widespread species,
which grows quickly at low altitudes in the whole
tropical belt, including arid zones (Morton, 1991;
Verdcourt, 1985; Jahn, 1986). It is generally known
in the developing world as a vegetable, a medicinal
plant and a source of vegetable oil. However, in the
Sudan it has been traditionally used in water purication (Jahn and Hamid, 1979; Jahn, 1981). These
multiple uses of the Moringa oleifera plant have
greatly promoted its widespread application. It has
been the favoured plant to grow in the vicinity of
homes for its esthetic beauty, as fence and for providing shading.
Numerous laboratory studies have so far shown
that Moringa oleifera seeds possess eective coagulation properties (Jahn, 1986, 1988; Karerwa, 1986;
Ndabigengesere, 1995; Muyibi and Evison, 1995a)
and that they are not toxic to humans nor animals
(Grabow et al., 1985; Berger et al., 1984). They are
quite ecient in reducing turbidity and microorganisms from raw waters (Jahn, 1986; Barth et al.,
1982; Sutherland et al., 1989; Olsen, 1987), in water
softening (Muyibi and Evison, 1995b), and also in
sludge conditioning (Ademiluyi, 1988). The active
agents of coagulation are dimeric cationic proteins
of molecular weight of approximately 13 kilodaltons (kDa) having an isoelectric point between 10
and 11 (Ndabigengesere et al., 1995; Fink, 1984).
As a coagulant, the crude water extract of Moringa
oleifera seeds compares favourably with alum, and
its use as an eective coagulant has been recommended in developing countries to reduce the
exorbitant cost of water treatment (Jahn, 1981,
1986, 1988; McConnachie, 1993).
Using this natural coagulant in developing
countries could eectively alleviate their economic
situation and allow further extension of water
supply to rural areas. However, except in traditional use and in some laboratory or pilot studies,
no large scale exploitation of Moringa oleifera in
water treatment has been reported so far. This
rejection may be explained by the presentation of
Moringa as a low technology appropriate only to
developing countries (Olsen, 1987; Jahn, 1988). One
way to improve acceptance of Moringa as a coagulant all over the world is to show clearly its advantages over conventional coagulants and apply

modern technology to supply it to water and wastewater treatment industry at cheaper cost.
Previous studies have focused mainly on the eciency of Moringa oleifera seeds as a coagulant, but
there is a need for a systematic evaluation of the
quality of the water treated by the crude water
extract of Moringa oleifera seeds. The purpose of
the present study is hence to examine various parameters of the quality of the water treated by coagulation using Moringa oleifera seeds, and
compare them with that of the water treated with
alum.

MATERIALS AND METHODS

Model turbid water used


Coagulation experiments were conducted on a model
turbid water, prepared with laboratory tap water, from
the city water supply of Sherbrooke, Quebec, Canada. The
tap water was rst collected in a plastic tank of 200 litres
and kept for at least three days before use in order to
obtain a uniform quality throughout the experimental
study. The quality parameters of the tap water were determined using Standards Methods (APHA et al., 1992);
Table 1 shows a few essential characteristics of the tap
water.
The model turbid water was prepared by adding kaolin
particles to the tap water (Morris and Knocke, 1984;
Ndabigengesere, 1995; Olsen, 1987; McConnachie, 1993).
Laboratory grade kaolin (5 g) (Anachemia, AC 5302),
were added to one litre of water and the suspension was
stirred for 30 min using a magnetic stirrer and then
allowed to settle for 24 h. The supernatant was then carefully removed and stored in a plastic bottle. The maximum
particle size remaining in the kaolin suspension was estimated using Stokes law to be around 2 mm. The kaolin
suspension was diluted using tap water to produce an initial turbidity of 105 NTU. It was obtained by diluting
400 ml of the kaolin suspension in 20 litres of tap water.
The model turbid water was agitated continuously to prevent sedimentation and change of initial turbidity. Then, a
few quality parameters of this model turbid water were
determined using Standard Methods (APHA et al., 1992).
Table 2 shows a typical analysis of the model turbid
water. The quality of this water was kept constant
throughout the study.
Moringa oleifera seeds and alum used
The Moringa oleifera seeds used as coagulant in this
study were obtained from Burundi in Central Africa. The
dry pods of Moringa oleifera were harvested in June 1991.
The bark enveloping the seeds in the pods was removed
Table 1. Quality parameters of the Sherbrooke tap water
Parameter
pH
Turbidity
Conductivity
Acidity
Alkalinity
Total hardness
Calcium (Ca2+)
Magnesium (Mg2+)
Chloride (Cl)
Sulphate (SO2
4 )
Nitrate (NO
3)
Phosphate (PO3
4 )

Tap water
7.3
1.0 NTU
150.0 m mho cm1
4.4 mg l1
52.0 mg l1 as CaCO3
46.0 mg l1 as CaCO3
48.0 mg l
21.0 mg l1
11.0 mg l1
7.4 mg l1
0.4 mg l1
0.5 mg l1

Coagulation using Moringa oleifera seeds


Table 2. Quality parameters of the model turbid water
Parameter

Synthetic turbid water

pH
Turbidity
Conductivity
Acidity
Alkalinity
Total hardness
Calcium (Ca2+)
Magnesium (Mg2+)
Chloride (Cl)
Sulphate (SO2
4 )
Nitrate (NO
3)
Phosphate (PO3
4 )

7.6
105 NTU
154.0 m mho cm1
4.7 mg l1
53.0 mg l1 of CaCO3
59.0 mg l1 of CaCO3
52.0 mg l1
22.0 mg l1
12.0 mg l1
8.0 mg l1
0.4 mg l1
0.5 mg l1

and the non-shelled seeds were transported to Canada and


stored until use in the Environmental Engineering
Laboratory of the University of Sherbrooke. The grains
were stored in dry conditions at room temperature of
232 18C and relative humidity of 82%. The storage
period tested varied from 6 weeks up to 4 years.
The aluminum sulphate [Al2(SO4)3.18H2O] used in the
present study was of laboratory grade (Anachemia AC405). The coagulation experiments with alum were performed only for comparison purposes. A 5% solution of
alum in tap water was used (5 g of alum in 100 ml). The
alum powder was totally soluble in the water. A fresh solution was prepared every day for reliable results.
Preparation of Moringa oleifera seeds as coagulant
Dry Moringa oleifera seeds were used as a coagulant in
two forms, namely shelled and non-shelled. For the shelled
seeds, the husk enveloping each seed was rst removed
manually, good quality seeds were then selected, and the
kernel was ground to a ne powder using an ordinary
electric kitchen blender. The non-shelled seeds were
directly ground to a ne powder using the same domestic
blender. The active agents of coagulation were then
extracted from the powder using tap water. The powder
was weighed and was added to the appropriate volume of
tap water. A concentration of 5% (5 g of powder in
100 ml) was used throughout this study after several trials
varying from 0.5 to 8%. The whole mixture was stirred
for 30 min at room temperature (20218C) using a magnetic stirrer. The suspension was ltered rst through a
Whatman no. 42 lter paper, and then through a 0.45 mm
membrane. The resultant ltrate solution was then used as
a coagulant. This is the crude water extract of Moringa
oleifera seeds. The quality parameters of this water extract
of Moringa oleifera were analyzed using Standard
Methods (APHA et al., 1992). Table 3 shows typical
characteristics of the crude water extract of both shelled
Table 3. Characteristics of the crude water extract of Moringa oleifera seeds
Parameter
pH
Conductivity m mho cm1)
Alkalinity (mg l1 as CaCO3)
Ca2+ (mg l1)
Mg2+ (mg l1)
Na+ (mg l1)
K+ (mg l1)
Fe3+ (mg l1)
Cl (mg l1)
1
SO2
4 (mg l )
1
NO
3 (mg l )
1
PO3
4 (mg l )
TOC (mg l1)
COD (mg l1)
TKN (mg l1)

Shelled seeds

Non-shelled seeds

6.4
1500.0
246.0
14.5
47.9
13.4
42.9
3.0
19.0
9.0
110.0
208.0
4760.0
15000.0
1193.0

5.8
1700.0
60.0
15.2
30.6
24.4
63.6
5.0
11.0
8.0
140.0
187.0
3678.0
9630.0
802.0

783

and non-shelled dry Moringa oleifera seeds. In order to


prevent any aging eects, such as change in pH, viscosity
and coagulation activity due to microbial decomposition
of organic compounds during storage (Jahn, 1988;
Ndabigengesere, 1995), a fresh solution was prepared for
each sequence of experiments.
To explore the possibility of more ecient ways to use
Moringa oleifera seeds as a coagulant, vegetable oils were
extracted rst from the powder using petroleum ether solvent, and then coagulating agents were extracted using tap
water. The extraction with petroleum ether was carried
out as described above in the case of water. The coagulating proteins were then puried from the water extract by
precipitation, dialysis, ion-exchange and lyophilization as
described previously in Ndabigengesere et al. (1995).
Figure 1 summarizes the protocol used to purify these
active proteins from the dry Moringa oleifera seeds. This
protocol consists of barks and husks removal, grinding,
extraction of the vegetable oil using petroleum ether or
any other suitable solvent, and then extraction of water
soluble proteins with water.
The proteins were precipitated from the aqueous extract
using ammonium sulphate [(NH4)2.SO4]. Ammonium sulphate salt was added to the solution until a saturation of
80 to 100% was reached. The precipitated proteins were
separated by ltration or centrifugation, dissolved in
water, and dialysed through a membrane with a molecular
weight cut-o of 12 00014 000 kDa to remove salts and
impurities of low molecular weights. The active proteins
remained inside of the dialysis tube.
The concentrated protein solution was then poured on a
carboxymethyl cellulose (CM) column. The cationic proteins were adsorbed and retained in the column. Next,
they were eluted from the column using 1 M NaCl solution. The eluted solution was dialysed again through the
12 00014 000 kDa membrane to remove salts. Finally, the
inside solution was lyophilized and the end-product was a
stable white protein powder.

Coagulation experiments
Jar test is the most widely used method for evaluating
and optimizing the coagulation-occulation processes
(Kawamura, 1991a; Bratby, 1980). This study consists of
batch experiments involving rapid mixing, slow mixing
and sedimentation. The Phipps and Bird jar test apparatus
was used in all the coagulation experiments. Glass beakers
of 1 litre lled with the model turbid water were used. The
apparatus allowed six beakers to be agitated simultaneously, and rotational speed could be varied between 0
and 100 rotations per minute (RPM), thus allowing simulation of dierent mixing intensities and resulting occulation process.
In a typical run, beakers were lled with one litre of the
model turbid water, placed on the oc illuminator and agitated at the preselected intensity of rapid mixing. During
rapid mixing, the coagulant dosage was added into each
beaker using Eppendorf pipettes. The duration of rapid
mixing was controlled with a stopwatch. After rapid mixing, the preselected intensity of slow mixing was quickly
established and its duration was controlled again with the
stopwatch. After slow mixing, the beakers were carefully
removed from the oc illuminator and were placed in a
safe place for the sedimentation phase to take place. In
this study, the intensity and duration of both rapid mixing
and slow mixing were xed respectively at 100 RPM for
2 min in the case of rapid mixing and 40 RPM for 20 min
in the case of slow mixing. The duration of sedimentation
was kept constant at 30 min. All experiments were run at
room temperature (20218C) and no pH control was exercised.

784

Anselme Ndabigengesere and K. Subba Narasiah

Fig. 1. Flow diagram showing the processing protocol of dry Moringa oleifera seeds.
Water quality parameters and equipment
Quality parameters of the water, treated by coagulation
either with Moringa oleifera seeds or with alum, were
measured following the sedimentation phase using
Standard Methods (APHA et al., 1992). The parameters
analyzed were turbidity, pH, conductivity, alkalinity, cation concentration (Ca2+, Mg2+), anion concentration
3

(Cl, SO2
4 , PO4 , NO3 ), organic compounds measured by
Chemical Oxygen Demand (COD) and Ultraviolet (UV)
absorbance at 280 nm, and sludge volume produced.
Turbidity measurements were conducted using a Hellige
turbidimeter. The pH was measured using a Hach-One pH
meter. Conductivity was measured using a Hach conductivity meter. Cation and anion concentrations were analyzed using a Hach spectrophotometer, Model DR 2000.
COD was measured using the same spectrophotometer
after a digestion period of 2 h in a Hach tube digester at
1508C. UV absorbance was measured using a UV/Visible
Cecil spectrophotometer, Model CE 599.
The same equipment was used to characterize the crude
water extract of Moringa oleifera seeds, as shown in
Table 3. If needed, dilution with distilled water was used.
Total Organic Carbon (TOC) was measured using a carbon analyzer apparatus, Dohrmann, Model DC-80. K+,
Na+, and Fe3+ were determined using an atomic absorp-

tion apparatus, Varian, Model AA 1275. Total Kjeldahl


Nitrogen (TKN) was measured using an automatic digester and titrator, Buchi, Model B 343, and lastly, sludge
volume was measured using plastic Imho cones.
RESULTS AND DISCUSSION

The experimental results on the quality of the


water treated both by shelled and non-shelled dry
Moringa oleifera seeds as well as alum as coagulant
are presented graphically in Fig. 2 through 5.
Turbidity removal and optimum dosage
The primary goal of coagulation-occulation is
turbidity removal. As shown in Fig. 2, dry Moringa
oleifera seeds, both shelled and non-shelled, demontrate an impressive coagulating capacity. The optimal dosage, which is the minimum dosage
corresponding to the lowest residual turbidity, is
equal to that of alum in the case of shelled seeds. It
was respectively 1 ml l1 for both alum and shelled
Moringa oleifera seeds (SMOS) and 10 ml l1 for

Coagulation using Moringa oleifera seeds

785

Fig. 2. Turbidity removal by coagulation with non-shelled Moringa oleifera seeds (NSMOS), shelled
Moringa oleifera seeds (SMOS), and alum (ALUM).

non-shelled Moringa oleifera seeds (NSMOS), corresponding to 50 and 500 mg l1, assuming that all
the seed material and the alum powder were entirely
solubilized and actively participated in the coagulation process. This is true for alum, but for
Moringa oleifera seeds, only 25% of the mass was
dissolved during extraction. For this reason, it was
decided to express the coagulant dosage in ml l1
throughout this study.
At optimum dosage, turbidity decreased from 105
to 10 NTU, corresponding to a turbidity removal of
90% in the case of alum, SMOS and NSMOS.
Compared with USEPA Standards (1991), which
state that the turbidity of drinking water should be
less than 1 NTU, the value of 10 NTU is quite
excessive. This means that sedimentation alone is
not enough to eliminate the ocs and that some
sort of ltration too has to be used for treating this
water. Flocs formed during rapid mixing were visible to the naked eye during slow mixing reaching a
diameter of approximately 1 to 2 mm. They settled
rapidly during the 30 min sedimentation period.
The dierence in optimum dosage between
shelled and non-shelled dry Moringa oleifera seeds
suggests that the active proteins are contained only
in the kernel. The average weight of the dry
Moringa oleifera seeds was 0.23 g and the kernel
represented 70 percent of the total mass, leaving
30% for the husk. As already shown in Table 3, the
crude water extract of Moringa oleifera seeds contained 1193 mg l1 of TKN in the case of shelled
seeds, and only 802 mg l1 of TKN in the case of
non-shelled ones. It may be noted that the protein
concentration is generally proportional to the TKN
concentration. As the optimum dosage was 10 times
greater in the case of non-shelled seeds, it is possible that some active proteins were denatured by
adsorption on the husk particles during extraction,
when the non-shelled seeds were processed. Thus,

shelled seeds are more suitable for further processing than non-shelled ones.
Quality of the treated water and sludge volume
Figure 3 to 5 show various parameters analyzed
on the treated water. As in the case of turbidity,
results of each parameter for non-shelled Moringa
oleifera seeds (NSMOS), shelled Moringa oleifera
seeds (SMOS) and aluminum sulphate (ALUM) are
plotted on the same graph for better comparison.
Each point in the graphics represents the average of
at least three measurements.
As can be seen from Fig. 3(a), Moringa oleifera
seeds did not aect signicantly the pH value,
which remained almost constant at 7.6 for all
dosages tested. In contrast, the pH value decreased
from 7.6 to 4.2 for alum, which means that in practical terms, further chemical addition is necessary in
order to correct the pH of the nished water to
values between 6.5 and 8.5 (USEPA, 1991).
From Fig. 3(b), it can be seen that Moringa oleifera seeds, both shelled and non-shelled, did not signicantly change the conductivity of the treated
water, which remained constant at 150 m mho cm1
for all dosages tested.
In contrast, the conductivity increased considerably form 150 to 842 m mho cm1 with the increasing dosage of alum. This increase in conductivity is
caused by sulphate ions remaining in the treated
water. As can be seen, the lowering of pH and
increase in conductivity in the case of alum puts
this chemical coagulant at a relative disadvantage
compared to Moringa oleifera.
Figure 3(c) shows the variation of alkalinity of
the treated water as a function of coagulant dosage.
Alkalinity remained almost constant in the case of
Moringa oleifera seeds, whereas it decreased rapidly
for 53 to 0 mg l1 (as CaCO3) in the case of alum.
The decrease in pH, alkalinity, and increase in ionic
strength are closely related and could cause an

786

Anselme Ndabigengesere and K. Subba Narasiah

Fig. 3. Quality of water treated by coagulation with dry Moringa oleifera seeds and alum: (a) pH value,
(b) conductivity, (c) alkalinity, and (d) sludge volume.

imbalance in the water chemistry thus initiating corrosion problems in the distribution network
(Degremont, 1989). It is common practice in water
treatment using alum to add alkalinity in the form
of bicarbonate or lime which increases the sludge
volume as well as the treatment costs. With
Moringa oleifera, it is not necessary to add any
other chemicals, as can be concluded from the experimental results.
Figure 3(d) shows the variation of the sludge
volume produced as a function of coagulant dosage.
At the optimal dosage of 1 ml l1 the shelled
Moringa oleifera seeds produced 1.5 ml of sludge,
whereas alum produced 7.6 ml. NSOMS produced
an even lower sludge volume of 1 ml at its optimum
dosage of 10 ml l1. The larger sludge volume in the
case of alum (4 to 5 times), compared to Moringa
oleifera, can be explained by the production of
aluminum hydroxide as a precipitate. In the case of
Moringa oleifera, only initial suspended particles

are agglomerated into larger and settleable ocs,


but no additional precipitate is formed. Besides
being voluminous, the alum sludges are gelatinous,
acidic, and dicult to dewater and to dispose of in
the environment (AWWA, 1990; Degremont, 1989).
The cost of sludge treatment and disposal being
proportional to the sludge volume, it may be
argued that Moringa sludges would be more economical to treat than alum sludges. A further advantage in the sludge problem lies in the fact that all
Moringa by-products are biodegradable organics, so
that the sludge can be used as a fertilizer provided
that no heavy metals are present in the water being
treated. The dierence in the sludge volume produced by SMOS and NSMOS can be explained by
the dierence in residual turbidity. The optimal
dosage for SMOS was 1 ml l1 and the residual turbidity was 7 NTU, whereas the optimal dosage for
NSMOS was 10 mg l1 and its residual turbity was
10 NTU. This dierence in residual turbidity

Coagulation using Moringa oleifera seeds

787

Fig. 4. Quality of water treated by coagulation with dry Moringa oleifera seeds and alum: (a) orthophosphates, (b) sulphates, (c) chlorides, and (d) nitrates.

suggests that SMOS removed more particles than


NSMOS. This is why SMOS produced slightly
higher sludge volume than NSMOS.
Figure 4(a) shows orthophosphates in the treated
water as a function of coagulant dosage. It can be
seen that alum slightly decreases orthophosphate
concentration, whereas both SMOS and NSMOS
increased their concentration signicantly, from 0.4
to 1.6 mg l1 with increasing dosage. Table 3 shows
that the crude water extract of SMOS and NSMOS
contains respectively 208 and 187 mg l1 of orthophosphate. However, orthophosphate is not directly
involved in the coagulation reaction because of its
soluble nature, and therefore a signicant fraction
remains in the treated water.
Although phosphate removal is not of great concern in the potable water treatment industry, it is of
importance in wastewater treatment, where euents
containing more than 1 mg l1 of total phosphorus
can cause eutrophication of lakes and streams
(Ryding and Rast, 1994). Alum on the other hand can

even be used in chemical dephosphatation of wastewaters (Weber, 1972). In this respect, it may be disadvantageous to use Moringa oleifera as it increases the
orthophosphate content of the treated water.
Concentration of sulphate ions in the treated
water is shown in Fig. 4(b) as a function of coagulant dosage. Neither SMOS nor NSMOS aected
the sulphate concentration, whereas alum increased
it considerably from 12 to 90 mg l1. USEPA
Standards (1991) for drinking water limit the concentration of sulphates at 250 mg l1 due to taste as
well as laxative considerations of sulphates in
water. The implication of sulphate ions in the
increase of conductivity has already been mentioned. Higher sulphate ion concentration means
also an increase in non-carbonate hardness in the
nished water which runs the risk of scale formation in boilers and other warm water appliances
(Degremont, 1989).
Figure 4(c) shows chloride ions in the treated
water as a function of coagulant dosage. They

788

Anselme Ndabigengesere and K. Subba Narasiah

remained almost at the same level with increasing


dosage of all the three coagulants. Therefore,
neither alum nor both shelled or non-shelled
Moringa oleifera seeds signicantly aect the chloride concentration. We know that chloride ions are
limited to 250 mg l1 in drinking water standards
due to associated taste and corrosion problems
(USEPA, 1991).
Concentration of nitrate ions in the treated water
is shown in Fig. 3(d) as a function of coagulant
dosage. Alum does not signicantly aect the concentration of nitrates, whereas both SMOS and
NSMOS increased their presence slightly, from 0.4
to 1 mg l1. It is known that nitrates must be limited to 10 mg l1 in drinking water because of the
risk of methaemoglobinaemia in infants (USEPA,
1991). As can be seen, this limit has not been
exceeded here, but if the water to be treated
requires a high dosage of coagulant, it is quite possible that it might be exceeded. As with orthophosphates, the aqueous extract of SMOS and NSMOS

contains signicant amounts of nitrates, namely


110 and 140 mg l1 respectively as shown in Table 3.
These nitrates do not participate in the coagulation
reaction and hence remain in the treated water.
This is a drawback of Moringa oleifera as far as
nitrates are concerned.
Figure 5(a) shows the concentration of magnesium ion in the treated water as a function of coagulant dosage. The concentration of magnesium
ion is almost independent of the coagulant dosage,
but it is slightly lower for alum than for SMOS and
NSMOS. The same information for calcium ion is
shown in Fig. 5(b). The concentration of calcium
ion is independent of the coagulant dosage.
Magnesium and calcium ions cause hardness of
water. It has been reported by Muyibi and Evison
(1995a) that Moringa oleifera is capable of softening
hard water. As shown in Fig. 5(a) and (b), the present results do not support their conclusion. It
should however be noted that compared to our
model water, they used harder waters and obtained

Fig. 5. Quality of water treated by coagulation with dry Moringa oleifera seeds and alum: (a) magnesium, (b) calcium, (c) chemical oxygen demand, and (d) absorbance at 280 nm.

Coagulation using Moringa oleifera seeds

a signicant hardness removal only at relatively


higher dosages. It may be possible that the optimum dosage for turbidity removal is lower than
that for water softening. As we are primarily interested in turbidity removal and in the improved
quality of the settled water, we did not try higher
dosages of the crude water extract of the dry
Moringa oleifera seeds to verify this. It would be an
additional asset if Moringa oleifera seeds could to
reduce hardness from waters.
Figure 5(c) and (d) show the concentration of organic matter in the treated water as a function of
coagulant dosage, as measured respectvely by COD
and UV absorbance at 280 nm. COD is normally
used to assess organic matter in water and wastewater. It does not however, provide information on
the dierent organic compounds present in the
sample. Proteins absorb ultraviolet rays principally
at 280 nm (Scopes, 1994). Hence we believe that the
UV measurement constitutes an easy tool to determine protein concentrations. It may be added that
other organic compounds can also absorb at that
wavelength and interfere in the measurement
(Scopes, 1994).
It can be seen that alum does not signicantly
reduce the organic matter, whereas both shelled and
non-shelled Moringa oleifera seeds tend to increase
it considerably in the treated water. As already
shown in Table 3 the crude water extract of both
SMOS and NSMOS contains signicant concentrations of organic matter present as COD, TOC
and TKN. After the coagulation process, some of
the organic matter remains in the treated water,
thus increasing its COD and its UV absorbency.
Parameters like COD and UV absorbency do not
allow identication of any specic organic compounds that are present. Thus, it is not possible to
base our discussion on the concentration of each
organic compound present in the drinking water
standard as we did previously with other parameters. Instead, we present in general the possible
eects of organic matter in the water treatment process as well as during the storage of the treated
water.
Toxicological studies conducted so far reported
that Moringa oleifera seeds do not constitute a
serious health hazard (Grabow et al., 1985; Berger
et al., 1984). Eilert and Nahrstedt (1981) isolated a
very active antimicrobial agent (4m-4-rhamnosyloxybenzyl-isothiocynate) from Moringa oleifera seeds.
It is readily soluble in water but its behaviour has
not yet been fully investigated. Some recent studies
reported the presence of mutagenic, genotoxic and
diuretic compounds in roasted dry Moringa oleifera
seeds (Villasenor et al., 1989a,b; Caceres et al.,
1991, 1992). Whether these compounds are also present in non-roasted seeds has not yet been studied.
Even though Moringa oleifera seeds are already in
use as food in some countries, further toxicological

789

studies are yet to be conducted on the use of these


seeds in water treatment.
Besides potential health risks, organic matter in
treated water can cause colour, taste and odour
problems, and these problems can be amplied by
the storage of treated water for longer periods.
Previous studies recommended storage periods of
no longer than 24 h (Jahn, 1986). This is feasible
only at the individual house level in rural areas,
where water is collected from natural sources in
small quantities once or twice a day. However, in
modern water distribution networks, the residence
time of water can easily exceed 24 h in pipes and
reservoirs. The same problem arises with the crude
water extract itself. When it is stored for periods
longer than two days, objectionable odours might
develop due to microbial decomposition of its organic compounds. In our case, although the solution was quite active as a coagulant, we preferred
to prepare a fresh extract every day as already mentioned.
In the water treatment industry, coagulation-occulation is followed by oc separation by sedimentation and/or ltration, and then by disinfection,
mostly by chlorine (AWWA, 1990; Degremont,
1989; Desjardins, 1988). During the disinfection
process by chlorine, organic matter can act as a precursor of trihalomethanes, which may be carcinogenic (Babcock and Singer, 1979; Moore and
Ramamoorthy, 1984; AWWA, 1990; USEPA,
1991). In order to use Moringa oleifera seeds even
in modern water treatment systems therefore, the
above problem should be addressed and suitable
methods of its preparation in safer and more stable
form should be developed. The foregoing investigations suggest that Moringa oleifera seeds should
be used in the water and wastewater treatment
industry only after an adequate purication of the
active proteins.
Results of the coagulation experiments with the
puried proteins showed that the optimal dosage
was 0.5 to 1 mg l1, which is 50 to 100 times lower
than the optimal dosage for alum or the crude
water extract of the shelled dry Moringa oleifera
seeds. With the puried extract, no increase of organic matter, phosphates or nitrates, in the treated
water was noticed as in the case of the crude water
extract. The protein powder is totally soluble in
water and it is stable. The protein powder remained
equally active in coagulation even after a storage
period of one year in a plastic bottle without any
special precaution. Hence, this puried protein form
is far more suitable to application in the water and
wastewater treatment industry.
Value of the vegetable oil from Moringa oleifera
seeds was recognized as early as the eighteenth century, but it has never been handled on a commercial
scale. When rened, the oil is slightly yellowish,
odourless and it has a sweet and pleasant avour. It
is suitable for edible purposes, in cosmetics and as a

790

Anselme Ndabigengesere and K. Subba Narasiah

lubricant for delicate machinery (Vaughan, 1970;


Said et al., 1974). The kernel contains 35% of oil
and 37% of proteins. Our earlier studies have
shown that petroleum ether dissolved 34% of the
total mass whereas water dissolved 25%
(Ndabigengesere, 1995). The vegetable oil from
Moringa oleifera seeds is called ben oil or behen oil.
In earlier times, the oil was exported to Europe
from India and it was used as a lubricant in delicate
machinery (Eckey and Miller, 1954).
The barks and husks of Moringa oleifera seeds
can be pyrolysed to produce activated carbon, valuable as an adsorbent in water treatment to remove
colour, taste and organic matter, as proposed by
Pollard et al. (1995). The residual solids may be
used as a feed to livestock and poultry, or used as a
fertilizer.
The cost of the puried active protein will probably be higher than the cost of alum at present. An
economic study of the purication process has not
yet been conducted. Many factors could help outweigh the cost of purication such as low optimal
dosage, environmental advantages of using this
natural coagulant, exploitation of its vegetable oil
and the valorization of the activated carbon and residual solids, etc. In order to exploit Moringa oleifera seeds as described, it must be cultivated on a
large scale say like coee and tea, which are cash
crops grown only in tropical regions but consumed
all over the world. On the other hand the cultivation
and the processing industries could generate employment and hence constitute a new source of income.
Moringa oleifera is a fast growing tree and dry pods
can be harvested after just one year of growing
(Jahn, 1986). Other species of Moringa may be considered too. Gene cloning of the Moringa proteins in
other plants should also be investigated.
CONCLUSIONS

. Both shelled and non-shelled dry Moringa oleifera


seeds can be used as a coagulant, but shelled
seeds are more eective.
. Analysis of the treated water shows that Moringa
oleifera seeds do not signicantly aect the pH,
conductivity, alkalinity, cation and anion concentrations, except for orthophosphate and nitrates,
which are increased. The crude water extract
increased considerably the concentration of organic matter, causing odour, colour and taste
problems during treated water storage. Organic
matter might consume additional chlorine in the
treatment plant and can act as a precursor of trihalomethanes during the disinfection process. For
these reasons, it is suggested that Moringa oleifera seeds be used in water and wastewater treatment only after a suitable purication of the
cationic active proteins.
. It is possible and even advantageous to exploit
simultaneously Moringa oleifera seeds as a coagulant as well as a vegetable oil source.

. Compared to alum, Moringa oleifera seeds do not


need pH and alkalinity adjustments, and do not
result in corrosion problems. They produce a
much smaller volume of sludge, which is not
hazardous.
. Moringa oleifera seeds present a viable alternative
coagulant to alum not only in developing
countries but worldwide.
AcknowledgementsThe authors gratefully acknowledge
the nancial support provided by the Programme
Canadien des Bourses de la Francophonie to carry out
this study. Thanks are also due to sta of the
Environmental Engineering and the Molecular Biology
Laboratories, Department of Civil Engineering and
Department of Biology respectively, University of
Sherbrooke, for their assistance in the laboratory works.

REFERENCES

Ademiluyi A. (1988) Sludge conditioning with Moringa


seeds. Envir. Int. 14, 5963.
Aizawa, T., Magara, Y. and Musashi, (1990) Problems
with introducing synthetic polyelectrolytes coagulants
into water purication. In Coagulation, Flocculation,
Sedimentation, Filtration and Flotation, International
Water
Supply
Association
(IWSA)/International
Association of Water Pollution Research and Control
(IAWPRC), Joint Specialised Conference, Jonkoping,
Water Supply, 8, 1117.
American Public Health Association (APHA), American
Water Works Association (AWWA) and Water
Pollution Control Federation (WPCF) (1992) Standard
methods for the examination of water and wastewater.
18th edition, APHA Publication Oce, Washington
D.C.
American Water Works Association (AWWA) (1990)
Water quality and treatment: a hand book of community
water supplies. McGraw Hill Publishing Company, 4th
edition, New York.
Babcock D.B. and Singer J.C. (1979) Chlorination and coagulation of humic and fulvic acids. J. Am. Wat. Wks
Ass. 71, 149152.
Barth V.P., Habs M., Klute R., Muller S. and Tauscher
B. (1982) Trinkwasseraufbereitung mit sames von
Moringa oleifera. Chem. Z. 106, 7578.
Berger M.R., Habs M., Jahn S.A.A. and Schmahl
D. (1984) Toxicological assessment of seeds from
Moringa oleifera and Moringa stenopetala, two highly
ecient primary coagulants for domestic water treatment of tropical waters. East Afr. Med. J. 61, 712717.
Berrak A. (1992) Inuence de quelques electrolytes sur le
comportement de oculants communs et d'alun basiques.
Memoire de Ma|trise, Universite de Sherbrooke,
Sherbrooke.
Biessinger K.E. and Stokes G.N. (1986) Eects of synthetic polyelectrolytes on selected aquatic organisms. J.
Wat. Pollut. Control Fed. 58, 207213.
Biessinger K.E., Lemke A.E., Smith W.E. and Tyo
R.M. (1976) Comparative toxicity of polyelectrolytes to
selected aquatic animals. J. Wat. Pollut. Control Fed.
48, 183187.
Boisvert J. P. (1992) Etude comparative de divers coagulants utilises pour le traitement physico-chimique des
eaux. Memoire de Ma|trise, Universite de Sherbrooke,
Sherbrooke.
Bratby J. R. (1980) Coagulation and occulation, with
emphasis on water and wastewater treatment. Uplands
Press Ltd., Croydon.

Coagulation using Moringa oleifera seeds


Caceres A., Saravia A., Rizzo S., Zibala L., De Leon E.
and Nave F. (1992) Pharmacological properties of
Moringa oleifera: II. Screening for antispasmodic, antiinammatory and diuretic activity. J. Ethnopharmacol.
36, 233237.
Caceres A., Cabrera O., Morales O., Mollinedo P. and
Mendia P. (1991) Pharmacological properties of
Moringa oleifera: I. Preliminary screening for antimicrobial activity. J. Ethnopharmacol. 33, 213216.
Degremont, (1989) Momento technique de l'eau. neuvieme
edition, tome 1 et tome 2, Paris.
Desjardins R. (1988) Le traitement des eaux. Edition de
l'Ecole Polytechnique de Montreal, deuxieme edition
revue, Montreal.
Eckey E. W. and Miller L. P. (1954) Vegetable fats and
oils, Reinhold Publishing Corporation, New York.
Edzwald J., James K. and Dempsey B.A. (1989)
Coagulation as an integrated water treatment processes.
J. Wat. Wks Ass. 81, 7278.
Eilert B.W. and Nahrstedt A. (1981) The antibiotic principle of seeds of Moringa oleifera and Moringa stenopetala. Planta Medica 42, 5561.
Fink W. (1984) Identizierung, Reindarstellung un
Strukturaufklarung ockungsaktiver wirkstoe aus hoheren planzen zur wasserreinigung, Doctoral thesis,
University of Heidelberg, Heidelberg.
Grabow W.O.K., Slabert J.L., Morgan W.S.G. and Jahn
S.A.A. (1985) Toxicity and mutagenicity evaluation of
water coagulated with Moringa oleifera seed preparations using sh, protozoan, bacterial, enzyme and
Ames Salmonella assays. Wat. SA 11, 914.
Goppers V. and Straub C.P. (1976) Polyelectrolyte persistence in a municipal water supply. J. Am. Wat. Wks Ass.
76, 319321.
Haarho J. and Cleasby J.L. (1988) Comparing aluminum
and iron coagulants for in-line ltration of cold waters.
J. Am. Wks Ass. 80, 168175.
Jahn S.A.A. (1988) Using Moringa seeds as coagulants in
developing countries. J. Am. Wat. Wks Ass. 90, 4350.
Jahn S. A. A. (1986) Proper use of African coagulants for
rural water supply: Research in the Sudan and a guide for
new projects. Deutsche Gesellschaft fur Technische
Zusammenarbeit (GTZ), Manual 191, Eschborn.
Jahn S. A. A. (1981) Traditional water purication in developing countries: Existing methods and potential application.
Deutsche
Gesellschaft
fur
Technische
Zusammenarbeit (GTZ), Manual 117, Eschborn.
Jahn S.A.A. and Hamid D. (1979) Studies on natural
water coagulants in Sudan, with special reference to
Moringa oleifera seeds. Wat. S.A. 5, 9097.
James C. and O'Melia C.R. (1982) Considering sludge
production in the selection of coagulants. J. Am. Wks
Ass. 74, 158251.
Jolicoeur C. and Haase D. (1989) Les aluns basiques dans
le traitement physico-chimique de l'eau: Survol de proprietes et evolution recente. Sci. Techn. Eau 22, 3146.
Karerwa T. (1986) Etude sur le captage des eaux de pluie
et le traitement des eaux de rivieres par les coagulants et
occulants naturels, Projet de n d'etudes, Universite du
Burundi, Bujumbura.
Kawamura S. (1991a) Integrated design of water treatment
facilities. John Wiley and Sons, New York.
Kawamura S. (1991b) Eectiveness of natural polyelectrolytes in water treatment. J. Am. Wat. Wks Ass. 83, 88
91.
Letterman R.D. and Pero R.W. (1990) Contaminants in
polyelectrolytes used in water treatment. J. Am. Wat.
Wks Ass. 82, 8797.
Letterman R.D. and Driscoll C.T. (1988) Survey of residual aluminum in ltered water. J. Am. Wat. Wks
Ass. 80, 154158.

791

Mallevialle J., Bruchet A. and Fiessinger F. (1984) How


safe are organic polymers in water treatment. J. Wat.
Wks Ass. 76, 8793.
McConnachie G. L. (1993) Water treatment for developing countries using baed channel hydraulic occulation. In: Proc. Instn Civ. Engrs Wat., Marit. and Energy,
101, 5561.
Miller R.G., Koper F.C., Kelty K.C., Stober J.A. and
Ulmer N.S. (1984) The occurrence of aluminum in
drinking water. J. Wat. Wks Ass. 76, 8491.
Moore J. K. and Ramamoorthy S. (1984) Organic chemical in natural waters: Applied monitoring and impact
assessment. Springer-Verlag, New York.
Morris J.K. and Knocke W.R. (1984) Temperature eects
on the use of metal - ion coagulants for water treatment. J. Am. Wat. Wks Ass. 66, 7479.
Morton J.F. (1991) The horseradish tree, Moringa pterygosperma (Moringaceae)a boon to arid lands. Econ.
Bot. 45, 318333.
Muyibi S.A. and Evison L.M. (1995a) Moringa oleifera
seeds for softening hardwater. Wat. Res. 29, 10991104.
Muyibi S.A. and Evison L.M. (1995b) Optimizing physical
parameters aecting coagulation of turbid water with
Moringa oleifera seeds. Wat. Res. 29, 26892695.
Ndabigengesere A. (1995) Etude du mecanisme et optimisation de la coagulation-occulation de l'eau avec le
Moringa Oleifera. These de Doctorat, Universite de
Sherbrooke, Sherbrooke.
Ndabigengesere A., Narasiah K.S. and Talbot B.G. (1995)
Active agents and mechanism of coagulation of turbid
waters using Moringa oleifera. Wat. Res. 29, 703710.
Olsen A. (1987) Low technology water purication by
bentonite clay and Moringa oleifera seeds occulation as
performed in Sudanese villages: eects on Schistosoma
mansoni cercariae. Wat. Res. 21, 517522.
Pollard S.J.T., Thompson F.E. and McConnachie
G.L. (1995) Microporous carbons from Moringa oleifera
husks for water purication in less developed countries.
Wat. Res. 29, 337347.
Qureshi N. and Malmberg R.G. (1985) Reducing aluminum residuals in nished water. J. Am. Wat. Wks Ass.
77, 101108.
Ryding S. O. and Rast W. (1994) Le controle de l'eutrophisation des lacs et des reservoirs, Masson, Paris.
Said S.I., Ismail M., Guirgis S., Kamel E. and Tahani
E.A. (1974) Benseeds: a potential oil source. Agric. Res.
Rev. 42, 4750.
Schultz C. R. and Okun D. (1984) Surface water treatment
for communities in developing countries. John Wiley and
Sons, New York.
Schultz C.R. and Okun D. (1983) Treating surface waters
for communities in developing countries. J. Am. Wat.
Wks Ass. 75, 212219.
Scopes K. R. (1994) Protein purication: Principles and
practice. Springer-Verlag, New York.
Sutherland J.P., Folkard G.K. and Grant W.D. (1989)
Seeds of Moringa species as naturally occurring occulants for water treatment. Sci., Technol. and Dev. 7,
191197.
USEPA (1991) Is your drinking water safe? Oce of Water
(WH-550), EPA 570/9-91-005, Washington DC.
Vaughan J. G. (1970) The structure and utilization of oil
seeds, Chapman and Hall Ltd., London.
Verdcourt B.A. (1985) A synopsis of the Moringaceae.
Kew Bull. 40, 123.
Villasenor I.M., Lim-Sylianco C.Y. and Dayrit F. (1989a)
Mutagens from roasted seeds of Moringa oleifera. Mut.
Res. 224, 209212.
Villasenor I.M., Finch P., Lim-Sylianco C.Y. and Dayrit
F. (1989b) Structure of mutagen from roasted seeds of
Moringa oleifera. Carcinogenesis 10, 10851087.
Weber W. J. (1972) Physicochemical processes for water
quality control. Wiley-Interscience, New York.

Potrebbero piacerti anche