You are on page 1of 7

Journal of Water Process Engineering 1 (2014) 84–90

Contents lists available at ScienceDirect

Journal of Water Process Engineering


journal homepage: www.elsevier.com/locate/jwpe

Biological treatment of wastewater contaminated with p-cresol using


Pseudomonas putida immobilized in polyvinyl alcohol (PVA) gel
Riham Surkatti, Muftah H. El-Naas *
Chemical and Petroleum Engineering Department, UAE University, P.O. Box 15551, Al-Ain, United Arab Emirates

A R T I C L E I N F O A B S T R A C T

Article history: In this study, the biodegradation of simulated wastewater containing p-cresol has been carried out using
Received 22 December 2013 Pseudomonas putida immobilized in PVA gel, in batch and continuous reactors. The effects of initial
Received in revised form 23 March 2014 concentration, temperature, pH and PVA volume fraction on the biodegradation of p-cresol were
Accepted 24 March 2014
evaluated in a batch SBBR. Continuous experiments were also carried out to study the effect of other
Available online 24 April 2014
operating parameters such as air flow rate and residence time on the biodegradation efficiency. The batch
experimental results indicated that the biodegradation capabilities of P. putida are highly affected by
Keywords:
temperature, pH and PVA vol.% and optimum performance were obtained at 35 8C, 8 and 40%,
Biodegradation
P-Cresol
respectively. The biomass did not seem to be inhabited by high concentration of p-cresol and the
Spouted Bed Bioreactor (SBBR) biodegradation data fitted well the Monod non-inhibitory Model. The kinetics data obtained from the
P. putida Monod model were utilized in modeling the continuous biodegradation process and gave very good
Immobilization agreement with experimental results.
Continuous flow ß 2014 Elsevier Ltd. All rights reserved.

1. Introduction formation of hazardous by-products [12,13]. Biological treatment


has been gaining attention as an efficient technology that is
Phenolic compounds such as cresols are major toxic pollutants versatile, environmental friendly, inexpensive and can offer
that are commonly observed in the effluent of various industrial complete mineralization of organic pollutants [13,14]. Many
wastewaters. p-Cresol is widely used in the production of microorganisms have shown high potential in utilizing cresol as
antioxidants, light-resistance, dyes, pigments and antiseptics sole source of carbon and energy, under aerobic and anaerobic
[1,2]. It is highly toxic, potentially carcinogenic, and can cause, operating conditions, even at relatively high concentrations [15–
even at low concentrations, adverse effect on the nervous, 19]. Aerobic microorganisms are more favorable in the degradation
cardiovascular and respiration systems [2–4]. p-Cresol has been of toxic compounds since they grow faster and can achieve
classified as pollutant of Group C (possible human carcinogens) complete conversion of organic pollutant to inorganic compounds
and listed as a priority pollutant by the US Environmental (CO2, H2O) [20]. Pseudomonas putida is well known Gram-negative,
Protection Agency (EPA) [5–7]. The World Health Organization aerobic bacterium that grows optimally at room temperature [21].
(WHO) recommends 0.001 mg/l as the acceptable p-cresol It is a member of the fluorescent pseudomonad group that has great
concentration in potable water. Therefore, there is an urgent need potential in the areas of bioremediation and biocatalysts [22]. The
to reduce the concentrations of p-cresol to the acceptable limit use of free bacteria in wastewater treatment usually creates
before discharging contaminated wastewater to any water body. numerous problems; consequently cell immobilization in biologi-
Traditionally, complex and expensive treatment techniques cal wastewater treatment is gaining attention by many investi-
have been employed for wastewater treatments including: carbon gators [23–25]. The immobilization of biomass has several
adsorption [8], ion exchange [9], activated solvent extraction [10], advantages including increasing the biodegradation rate; enhanc-
and chemical oxidation [11]. These alternatives are usually ing the control of the bioprocess; improving biocatalyst stability
associated with numerous drawbacks such as high cost and the and advancing the tolerance against harsh environmental condi-
tions [26,27]. Among immobilization carrier, PVA is a type of
polymer that has been widely used in the area of biocatalysts [28].
It is inexpensive and non-toxic synthetic polymer, which can be
* Corresponding author. Tel.: +971 3 713 5188.
E-mail address: Muftah@uaeu.ac.ae (M.H. El-Naas).
physically cross-linked by iterative freezing–thawing process,

http://dx.doi.org/10.1016/j.jwpe.2014.03.008
2214-7144/ß 2014 Elsevier Ltd. All rights reserved.
R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90 85

producing a gel that has a fibril network with highly porous jet injected through an orifice in the bottom of the reactor [34]. A
structure, high mechanical strength, and elastic nature [24,29,30]. schematic diagram of the Spouted Bed Bioreactor (SBBR) is shown
The porous structure increases the diffusion of the substrate and in Fig. 1.
oxygen to enhance the biodegradation process. It is well known
that this method also decreases the toxicity and increases 2.3. Reagents
mechanical strength for the prepared matrix [31,32].
The aim of this study is to evaluate the biodegradation of p- Analytical grade p-cresol was purchased from Sigma–Aldrich in
cresol by P. putida immobilized in poly vinyl alcohol (PVA) gel and a powder form. All other chemicals and PVA powder were of
to investigate the effect of several operating conditions on the analytical grade and were obtained from BDH, UK.
biodegradation process. Batch and continuous experiments were
carried out in a specially designed Spouted Bed Bioreactor (SBBR) 2.4. Synthetic wastewater preparation
to evaluate the effect of initial p-cresol concentration, operating
temperature, solution pH, PVA volume fraction, air flow rate and Synthetic wastewater of p-cresol was prepared for the desired
liquid flow rate on the biodegradation rate. concentrations in mineral nutrient solution before each experi-
mental run. The solutions were always kept in a brown flask to
2. Materials and methods avoid light oxidation of the p-cresol.

2.1. Preparation and acclimatization of immobilized bacteria in PVA 2.5. Analytical methods
gel
Analysis for p-cresol was carried out using Chrompack Gas
A special strain of the bacterium P. putida was obtained in an Chromatograph, Model CP9001 with flame ionization detection
AMNITE cereal form (P300) from Cleveland Biotech Ltd. The cereal (GC/FID). The GC was equipped with capillary column (Stabilwax,
contained a consortium of microorganisms, with P. putida as the 30 m, 0.25 mm ID) and a flame ionization detector which was set at
dominate strain. Bacterial cells were extracted from the cereal, 250 8C. A sample of 1 ml was filtered through syringe filter with
immobilized in PVA gel by freezing–thawing process, and pore size of 0.45 mm and injected into the GC. The temperature
acclimatized to phenol solution with concentrations up to program started at 100 8C and increased at a rate of 20 8C min1 to
200 mg/l. More details about the bacterial extraction and 180 8C. The analyses were conducted in duplicates, and a standard
immobilization were reported in previous studies by El-Naas solution was used to recheck the accuracy of the GC.
et al. [28,33]. The immobilized bacteria were then gradually
acclimatized to p-cresol concentration up to 200 mg/l over a period 3. Results and discussions
of 7 days by placing the PVA particles in p-cresol synthetic solution
with mineral nutrients in Spouted Bed Bioreactor (SBBR). 3.1. Batch biodegradation of p-cresol

2.2. Spouted Bed Bioreactor (SBBR) Batch experiments were carried out in order to study the effect
of initial concentration, temperature, pH and PVA vol.% on the
The Spouted Bed Bioreactor was made of Plexiglas with a total biodegradation process. All experimental results reported in the
volume of 300 ml and fitted with a surrounding jacket for next sections were based on averaging results of repeated
temperature control. Air was continuously introduced at a experimental runs (duplicates).
specified flow rate into the reactor to enhance mixing and at
the same time provide excess oxygen to maintain aerobic 3.1.1. Effect of p-cresol initial concentration
conditions [14]. The temperature of the reactor content was Batch experiments were carried out at various initial p-cresol
controlled to the desired value by the surrounding jacket. The SBBR concentrations (25, 50, 100, 150 and 200 mg/l) to evaluate the
is characterized by a systematic intense mixing due to the cyclic effect of initial concentration on the biodegradation rate. From the
motion of particles within the bed, that is generated by a single air start, a yellow color was observed for a short period of time during

Fig. 1. A schematic diagram of the spouted bed bioreactor (SBBR).


86 R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90

250 320
25 mg/l
p-cresol Concentration (mg/l)

50 mg/l
100 mg/l
200 150 mg/l 300

Biodegradation Rate ( mg/l.h)


200 mg/l

150 280

100 260

240
50

220
0
0 10 20 30 40
Time (min) 200
20 25 30 35 40 45 50
Fig. 2. Variation of p-cresol concentration with time; PVA vol.% = 30% and T = 30 8C. Temperature (ºC )

Fig. 4. Variation of biodegradation rate with temperature. Initial pH 7; PVA


volume = 30% of total volume; initial p-cresol concentration = 200 mg/l.
360

340
Biodegradation Rate (mg/l.h)

320 higher than 40 8C or lower than 25 8C tend to have negative effect


on the biodegradation rate. Low temperatures usually result in
300
slowing down the activity of the bacteria, while raising the
280 temperature to high values (higher than 40 8C) leads to deactiva-
tion of the main biodegradation enzymes [37,38].
260

240 3.1.3. Effect of pH


Being proteins, enzymes are stabilized by weak hydrogen bonds
220 and are generally affected by the variation of pH [39]. In order to
200 study the effect of pH on the biodegradation, experiments were
conducted at pH ranging from 5.0 to 8.0. Initial p-cresol
180 concentration and temperature were fixed at 200 mg/l and
0 20 40 60 80 100 120 140 160 180 200 220
30 8C. It is believed that most organisms cannot tolerate pH levels
p-cresol Concentration (mg/l) below 4.0 or above 9.5, and the optimum pH for most
Fig. 3. Biodegradation rate at different p-cresol initial concentrations; PVA microorganisms lies in this range. As shown in Fig. 5, the
vol.% = 30% and T = 30 8C. biodegradation rate of p-cresol is diminished at pH 5 and increases
sharply at higher values. The experimental results indicate that the
biodegradation of p-cresol reaches optimum at a pH value of 8.
all experimental runs. This is believed to be due to the formation of However, any pH variation between 6 and 8 does not seem to have
2-hydroxymuconicsemialdehyde, which has been produced from the any negative effect on the biodegradation rate. Similar results were
metabolizing of 3-methyl catechol by meta pathway [35,36]. The reported by Singh et al. [40] for the biodegradation of p-cresol using
concentration of p-cresol at different time intervals was measured Gliomastix indius.
quantitatively by GC (FID) as mentioned in Section 2.5. Fig. 2 shows
the experimental data of the reduction in p-cresol concentration.
Clearly the reduction seems to be linear with time, which indicates 340
constant biodegradation rates. The biomass did not seem to be
330
inhibited by the p-cresol as indicated by the exponential increase of
the biodegradation rate with initial concentration (Fig. 3). The lack 320
BiodegradationRate (mg/l.h)

of inhibitory effect even at high concentrations of 200 mg/l could 310


be attributed to the biomass immobilization in PVA gel, which acts 300
as a protective shelter against substrate toxicity [23]. However, the 290
inhibitory effect of p-cresol has been reported at 400 mg/l by
280
Basheer and Farooqi [4] in the biodegradation of p-cresol using
270
activated sludge.
260
3.1.2. Effect of temperature 250
Experiments were carried out at temperatures ranging from 25 240
to 45 8C to study the effect of temperature on the degradation of p- 230
cresol. Other conditions such as initial concentration and pH were
220
kept constant at 200 mg/l and 7, respectively. Fig. 4 presents the 4 5 6 7 8 9
experimental results of the effect of temperature on the
pH
biodegradation rate. Increasing the operating temperature from
25 to 30 8C seems to enhance the biodegradation capability of P. Fig. 5. Effect of the initial solution pH on the biodegradation rate; PVA Vol.% = 30;
putida, reaching an optimum between 30 and 40 8C. Temperatures initial p-cresol concentration = 200 mg/l; T = 30 8C.
R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90 87

500 50 mg/l
200 100 mg/l

p-cresol Concentration (mg/l)


150 mg/l
450
Biodegradation Rate (mg/l.h)

200 mg/l

400 150

350
100
300

250 50

200
0
0 50 100 150 200 250
150
15 20 25 30 35 40 45 Time (min)
PVA%
Fig. 7. p-Cresol concentration as a function of time for different initial
Fig. 6. Variation of biodegradation rate with PVA% volume; T = 35 8C; initial p-cresol concentrations. Reactor temperature = 35 8C; pH = 8; air flow rate 2 l/min; liquid
concentration = 200 mg/l; pH = 7. flow rate = 5 ml/min.

3.1.4. Effect of PVA% in the total operating volume the same as that of the outlet. Samples were collected from the
The amount of PVA pellets can be considered as an indirect effluent stream and analyzed at different time intervals. The
measure of the amount of active biomass available for the reduction of p-cresol concentration as a function of time at different
biodegradation of p-cresol, and consequently it is one of the most initial concentrations is shown in Fig. 7, and indicates that the
important factors that can affect the biodegradation process. biodegradation rate is highly dependent on the initial concentra-
Experiments were conducted to study the effect of PVA volume, as tion. The residence time inside the reactor (60 min) was sufficient
a fraction of the total operating reactor volume, on the to completely consume the substrate at low concentration (50 mg/
biodegradation rate. The initial p-cresol concentration, solution l). However, at higher concentrations, the substrate was not
pH and temperature were kept at 200 mg/l, 7, and 30 8C, completely degraded, but steady state was achieved within one
respectively. The total operating volume of the bioreactor was residence time. It must be mentioned here that during the startup
maintained at 300 ml. Fig. 6 shows the biodegradation rate for of the biodegradation process, sharp reduction in p-cresol
three PVA volume fractions (20, 30 and 40%). It is noticeable that concentration was observed, followed by a period of slow
the biodegradation rate of p-cresol tends to increase linearly with reduction before stabilizing. A faster sorption step due to the
increasing the PVA volume fraction in the reactor. This is expected physiochemical interactions between the organic chemicals and
as the amount of PVA is directly related to the amount of bacteria in microbial cell walls is believed to have resulted in the first stage of
the bioreactor, and hence the presence of more PVA particles the biodegradation process [41,42]. During this phase, the removal
means the presence of more bacterial cells. In addition, the volume by the biodegradation is less significant. However, in the second
fraction of the p-cresol solution is reduced. Although larger PVA phase, the entire exposed surface of the PVA pellets gets saturated
volume fractions were not tested, it is expected that more PVA with the substrate or reach closed to saturation, then the removal
particles in the reactor will hinder the movement and mixing of the by adsorption becomes less significant. In this phase, the
pellets, and consequently, reduce the biodegradation rate. contribution of biodegradation becomes significant. The overall
percent removal is plotted as a function of initial concentration in
3.2. Continuous biodegradation for p-cresol Fig. 8. In all cases, the percent removal reached more than 85%,
which indicates that the immobilized bacteria have very high
Although the batch experimental work provided essential data potential for the biodegradation of p-cresol.
regarding the effect of certain operating parameters on the
biodegradation rate, continuous operation is vital in assessing
the potential industrial application of the biodegradation process.
Experiments were carried out to study the continuous biodegra-
dation process by feeding a synthetic wastewater containing p- 100
cresol with different initial concentrations to the bioreactor using a
peristaltic pump (Watson Marlow, Model 323) for a period of 4 h,
or until the biodegradation rate reaches steady state. The reactor
temperature and pH were kept constant at 8 and 35 8C, which are 90
% Removal

the optimum conditions obtained in the batch study. In all


experiments the volume fraction of the PVA pellets was kept at 30%
of the total operating volume. Again, all experiments were carried
out in duplicates and average values are reported in the following 80
sections. The standard deviation ranged between 2 and 5%.

3.2.1. Effect of initial concentration


The effect of p-cresol concentration on the biodegradation rate 70
0 50 100 150 200
was studied at four initial concentrations: 50, 100, 150 and
p-cresol Initial Concentration (mg/l)
200 mg/l. Both liquid and air flow rates were fixed at 5 ml/min and
2 l/min, respectively. Since the SBBR was operated under well- Fig. 8. p-Cresol removal % as a function of initial concentration. Reactor
mixed conditions, the substrate concentration in the reactor was temperature = 35 8C; air flow rate 2 l/min; liquid flow rate = 5 ml/min.
88 R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90

160 160
5 ml/min 1 L/min
6.5 ml/min 140 2 L/min
140 3 L/min
10 ml/min

p-cresol Concentration (mg/l)


p-cresol Concentration (mg/l)

120 120

100
100
80
80
60
60
40
40
20
20
0
0 50 100 150 200 250
0
0 50 100 150 200 250 Time (min)
Time (min)
Fig. 10. Concentration of p-cresol as a function of time for different air flow rates;
Fig. 9. Concentration of p-cresol in the reactor as a function of time for different Initial p-cresol concentration = 150 mg/l; reactor temperature = 35 8C; liquid flow
liquid flow rates (LF). Initial p-cresol concentration = 150 mg/l; air flow rate 2 l/min. rate 5 ml/min.

oxygen for the aerobic biodegradation, it does not seem to provide


3.2.2. Effect of liquid flow rate the smooth mixing needed for good mass transfer. Based on visual
The effect of liquid flow rate on the biodegradation efficiency observation, high flow rates lead to slugging making the small
was evaluated for three different flow rates (5, 6.5 and 10 ml/min), Spouted Bed Bioreactor lose its systematic cyclic motion and good
while keeping the initial concentration and airflow rate constant at particle movement. Similar observations were reported by El-Naas
150 mg/l and 2 l/min (Fig. 9). It is often anticipated that increasing et al. [43] and Salehi et al. [45], who noted that after a certain air
the liquid flow rate would reduce the residence time inside the flow rate the biodegradation rate could no longer be improved.
bioreactor and hence reduce the biodegradation rate. In fact
increasing the liquid flow rate may have two opposing factors. On 3.3. Modeling of p-cresol biodegradation
one hand, the higher flow velocity around the PVA pellets may
improve external mass transfer, especially for low concentrations Modeling of the biodegradation process involves relating the
of p-cresol. On the other hand, the reduced residence time will specific growth rate of the biomass to the consumption rate of the
decrease the contact between the immobilized bacteria and the substrate. Based on material balance, the specific consumption rate
substrate [43]. The latter seems to be the dominate factor and of the substrate is expressed as follows:
hence the observed reduction in biodegradation efficiency. Shetty
et al. [44] suggested that increasing the liquid velocity may result dS m
qs ¼ ¼ (1)
in higher erosion of the biomass from the biofilm surface, and Xdt Y
reduce the biodegradation rate. This may be valid for biofilms but where qs is the specific consumption rate (h1); X is the biomass
not for bacterial cell immobilized in PVA particles. These cells are concentration (mg/l); S is the substrate concentration (mg/l); Y is
usually very well nested inside the porous structure of the PVA and the cell mass yield (g/g); In order to represent the degradation
rarely get stripped away in significant numbers by the high liquid kinetics of p-cresol, two models (Haldane and Monod) were used to
flow. fit the experimental data obtained from the batch degradation
experiments (described in Section 3.1.1). The first model considers
3.2.3. Effect of air flow rate p-cresol as non-inhibitory compound and neglects the inhibitory
Aerobic biodegradation processes require systems which effect, while the second one takes into account the inhibitory effect
ensure the adequate oxygen supply for maintaining the growth of p-cresol. The Monod model can be expressed in terms of the
of the microorganisms. Air flow rate plays an important role in the degradation rate, q (mg/l h) as:
biodegradation process using aerobic bacteria such as P. putida. In
addition to providing enough oxygen for the biodegradation qmax S
q¼ (2)
process, it enhances mixing and particles movement in the ks þ S
bioreactor. The effect of air flow rate on the continuous
The Haldane model is expressed by:
biodegradation process was investigated for three different values:
1, 2 and 3 l/min. The initial p-cresol concentration and liquid flow qmax S
rate were fixed at 150 mg/l and 5 ml/min, respectively. q¼ (3)
ks þ S þ ðS2 =ki Þ
The variation of p-cresol concentration with time for the three
flow rates is presented in Fig. 10. It seems that the biodegradation where So is the initial substrate concentration (mg/l), qmax is the
process is optimized at air flow rate of 2 l/min, which can provide maximum consumption rate (mg/l h), Ks is the half saturation
sufficient amount of oxygen for the immobilized bacteria to coefficient (mg/l) and KI is the substrate inhibition constant (mg/l).
complete the aerobic biodegradation process. At a lower air flow These two models can be used to predict the variations of the
rate (1 l/min), however, the distribution of PVA pellets was not biodegradation rate (q) with initial p-cresol concentrations,
uniform through the bioreactor. Visual observation indicated that utilizing the relation in Eq. (1) and assuming that Y is constant
this flow rate could not provide enough air lift to give good over the concentration range. This assumption is valid if the
spouting of the particles and consequently, most of particles were concentration is much higher than Ks (S  Ks) [33]. The degrada-
settling at the bottom of the bed with negligible movement. tion rate q was determined from a plot of the substrate
Although higher flow rate (3 l/min) may provide more than enough concentration S versus time for each initial concentration So. From
R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90 89

Table 1 160
Degradation kinetics parameters for Monod and Haldane Models. Model
140 100 mg/l
Model qmax (mg/l h) Ks (mg/l) KI (mg/l) R2 150 mg/l

p-cresol Concentration (mg/l)


Monod 357.9 21.64 – 0.9748 120
Haldane 357.9 21.64 1.903  108 0.9748
100

the values of q and So, the values of kinetics parameters for both 80
models were obtained using SigmaPlot non-linear regression
60
which uses the Marquardt–Levenberg algorithm. Parameters
values for both models are presented in Table 1. Although both
40
models give acceptable fit for the data with the same R2 value
(0.975), The Monod non-inhibitory model clearly gives a much 20
more acceptable fit between the data and model equation. Large KI
value was obtained in Haldane Model, which indicates that the 0
0 50 100 150 200 250
culture is less sensitive to the substrate inhibition. Monod model is
compared with the experimental data in Fig. 11. The biodegrada- Time (min)
tion kinetics were obtained from fitting the batch experimental
Fig. 12. Comparison of the continuous experimental data with fitted model
data to the Monod model; kinetics parameters according to Monod (So = 100 and 150 mg/l).
non-inhibition model was determined to be qmax and ks are
357.9 mg/l h and 21.64 mg/l respectively.
The global biodegradation rate was calculated based on the where So is the initial substrate concentration in (mg/l), S is the
specific consumption rate, where p-cresol is the limiting substrate, concentration at desired time (mg/l), (rs) is the rate of removal of
while oxygen and other nutrients are in excess. Assuming perfect substrate (mg/l h), V is the reactor volume (l), t is the time (h) and Fs
mixing in the bioreactor, the mass balance of the continuous flow is the volumetric flow rate (l/h). The model was applied for
reactor can be expressed as follows: different p-cresol concentrations in the effluent and the variation of
the substrate concentration with time (given by Eq. (8)) was
dN A
¼ F A ðC Ao  C A Þ  r A V (4) evaluated using (E–Z Solve). The model predictions are compared
dt with the experimental data in Fig. 12. The figure shows very good
Dividing the equation by V; agreement between the model predications and the experimental
data and confirms the validity of the proposed model for
dS F s simulating the continuous biodegradation of p-cresol in SBBR.
¼ ðSo  SÞ  r S (5)
dt V
The substrate uptake rate is given by: 4. Conclusions
r s ¼ q (6)
The results of the batch biodegradation study showed that p-
Assuming constant yield in the bioreactor as mentioned above, cresol biodegradation process was highly dependent on the initial
the substrate consumption can be described as follows: concentration, temperature, pH and the amount of the bacteria in
the bioreactor. The biodegradation process was optimized at
qmax S
r s ¼ (7) temperature and pH values of 35 8C and 8, respectively. In addition,
ks þ S
the biodegradation rate was found to increase with the initial
Combining Eqs. (5) and (7), the change of p-cresol concentration concentration and PVA volume fraction. Continuous biodegrada-
in the reactor can be evaluated as: tion results indicated that P. putida had high potential for the
biodegradation p-cresol up to 200 mg/l, with a removal efficiency
dS F s q S of more than 85%. The biodegradation efficiency was affected by
¼ ðSo  SÞ  max (8)
dt V ks þ S other parameters such as air flow rate and residence time, with an
optimal air flow rate of 2 l/min.
360 Modeling of the batch experimental data indicated that p-cresol
340 was non-inhibitory substrate for P. putida, and the biodegradation
process was successfully described by Monod model with
Biodegradation Rate (mg/l.hr)

320 determination coefficient (R2) of 0.98. The obtained kinetics data


were utilized for modeling the continuous biodegradation process
300
and showed very good fit to the experimental results.
280
Acknowledgements
260

240 The authors would like to acknowledge the financial support


provided by the Japan Cooperation Center, Petroleum (JCCP) and
220
the technical support by JX Nippon Research Institute Co., Ltd (JX-
200
Experimental NRI).
Monod model
180
0 50 100 150 200 References

p-cresol Concentration (mg/l) [1] D. Bergé-Lefranc, C. Vagner, R. Calaf, H. Pizzala, R. Denoyel, P. Brunet, et al., In vitro
elimination of protein bound uremic toxin p-cresol by MFI-type zeolites, Micro-
Fig. 11. Comparison of the batch experimental data with fitted model (Monod). porous Mesoporous Mater. 153 (2012) 288–293.
90 R. Surkatti, M.H. El-Naas / Journal of Water Process Engineering 1 (2014) 84–90

[2] M. del Olmo, C. Dı́ez, A. Molina, I. de Orbe, J.L. Vı́lchez, Resolution of phenol, o- [25] M. Szczesna-Antczak, E. Galas, Bacillus subtilis cells immobilised in PVA-cryogels,
cresol, m-cresol and p-cresol mixtures by excitation fluorescence using partial Biomol. Eng. 17 (2001) 55–63.
least-squares (PLS) multivariate calibration, Anal. Chim. Acta 335 (1996) 23–33. [26] L.-s. Zhang, W.-z. Wu, J.-l. Wang, Immobilization of activated sludge using
[3] J.M. Sanders, J.R. Bucher, J.C. Peckham, G.E. Kissling, M.R. Hejtmancik, R.S. Chhabra, improved polyvinyl alcohol (PVA) gel, J. Environ. Sci. 19 (2007) 1293–1297.
Carcinogenesis studies of cresols in rats and mice, Toxicology 257 (2009) 33–39. [27] Y.J. Liu, A.N. Zhang, X.C. Wang, Biodegradation of phenol by using free and
[4] F. Basheer, I.H. Farooqi, Biodegradation of p-cresol by aerobic granules in se- immobilized cells of Acinetobacter sp. XA05 and Sphingomonas sp. FG03, Biochem.
quencing batch reactor, J. Environ. Sci. 24 (2012) 2012–2018. Eng. J. 44 (2009) 187–192.
[5] A.y.-e. Hamitouche, Z. Bendjama, A. Amrane, F. Kaouah, D. Hamane, R. Ikkene, [28] M.H. El-Naas, A.-H.I. Mourad, R. Surkatti, Evaluation of the characteristics of
Biodegradation of p-cresol by mixed culture in batch reactor – effect of the three polyvinyl alcohol (PVA) as matrices for the immobilization of Pseudomonas putida,
nitrogen sources used, Proc. Eng. 33 (2012) 458–464. Int. Biodeterior. Biodegrad. 85 (2013) 413–420.
[6] D. Arya, S. Kumar, S. Kumar, Biodegradation dynamics and cell maintenance for [29] W. Yujian, Y. Xiaojuan, L. Hongyu, T. Wei, Immobilization of Acidithiobacillus
the treatment of resorcinol and p-cresol by filamentous fungus Gliomastix indicus, ferrooxidans with complex of PVA and sodium alginate, Polym. Degrad. Stab. 91
J. Hazard. Mater. 198 (2011) 49–56. (2006) 2408–2414.
[7] R.K. Singh, S. Kumar, S. Kumar, A. Kumar, Development of parthenium based [30] R. Hernández, A. Sarafian, D. López, C. Mijangos, Viscoelastic properties of
activated carbon and its utilization for adsorptive removal of p-cresol from poly(vinyl alcohol) hydrogels and ferrogels obtained through freezing–thawing
aqueous solution, J. Hazard. Mater. 155 (2008) 523–535. cycles, Polymer 45 (2004) 5543–5549.
[8] P. Khare, A. Kumar, Removal of phenol from aqueous solution using carbonized [31] C.M. Hassan, J.H. Ward, N.A. Peppas, Modeling of crystal dissolution of poly(vinyl
Terminalia chebula-activated carbon: process parametric optimization using con- alcohol) gels produced by freezing/thawing processes, Polymer 41 (2000) 6729–6739.
ventional method and Taguchi’s experimental design, adsorption kinetic, equi- [32] O. Tepe, A.Y. Dursun, Combined effects of external mass transfer and biodegra-
librium and thermodynamic study, Appl. Water Sci. 2 (2012) 317–326. dation rates on removal of phenol by immobilized Ralstonia eutropha in a packed
[9] C. Flox, J.A. Garrido, R.M. Rodrı́guez, F. Centellas, P.-L. Cabot, C. Arias, et al., bed reactor, J. Hazard. Mater. (2008) 9–16.
Degradation of 4,6-dinitro-o-cresol from water by anodic oxidation with a [33] M.H. El-Naas, S.A. Al-Muhtaseb, S. Makhlouf, Biodegradation of phenol by Pseu-
boron-doped diamond electrode, Electrochim. Acta 50 (2005) 3685–3692. domonas putida immobilized in polyvinyl alcohol (PVA) gel, J. Hazard. Mater. 164
[10] J. Zhou, W. Duan, J. Xu, Y. Yang, Experimental, Simulation study on the extraction (2009) 720–725.
of p-cresol using centrifugal extractors, Chin. J. Chem. Eng. 15 (2007) 209–214. [34] S. Al-Zuhair, M. El-Naas, Immobilization of Pseudomonas putida in PVA gel
[11] Y. Zheng, D.O. Hill, C.H. Kuo, Destruction of cresols by chemical oxidation, J. particles for the biodegradation of phenol at high concentrations, Biochem.
Hazard. Mater. 34 (1993) 245–260. Eng. J. 56 (2011) 46–50.
[12] S.-Y. Tsai, R.-S. Juang, Biodegradation of phenol and sodium salicylate mixtures by [35] P.Y. Ahamad, A. Kunhi, S. Divakar, New metabolic pathway for o-cresol degrada-
suspended Pseudomonas putida CCRC 14365, J. Hazard. Mater. 138 (2006) 125–132. tion by Pseudomonas sp. CP4 as evidenced by 1H NMR spectroscopic studies,
[13] K. Lika, I.A. Papadakis, Modeling the biodegradation of phenolic compounds by World J. Microbiol. Biotechnol. 17 (2001) 371–377.
microalgae, J. Sea Res. 62 (2009) 135–146. [36] S. Masunaga, Y. Urushigawa, Y. Yonezawa, Biodegradation pathway of o-cresol by
[14] M.H. El-Naas, S. Al-Zuhair, S. Makhlouf, Batch degradation of phenol in a spouted heterogeneous culture: Phenol acclimated activated sludge, Water Res. (1986)
bed bioreactor system, J. Ind. Eng. Chem. 16 (2010) 267–272. 477–484.
[15] M. Claußen, S. Schmidt, Biodegradation of phenol and p-cresol by the hypho- [37] J.E. Bailey, D.F. Ollis, Biochemical Engineering Fundamentals, second ed.,
mycete Scedosporium apiospermum, Res. Microbiol. 149 (1998) 399–406. McGraw-Hill Book Co., Singapore, 1986.
[16] N. Pazarlioglu, Y. Kaymaz, A. Babaoglu, Biodegradation kinetics of o-cresol by [38] I. de Ory, L. Enrique Romero, D. Cantero, Modeling the kinetics of growth of
Pseudomonas Putida DSM 548 (pJP4) and o-cresol removal in a batch-recirculation Acetobacter aceti in discontinuous culture: influence of the temperature of
bioreactor system, Electron. J. Biotechnol. 15 (2011) 93–97. operation, Appl. Microbiol. Biotechnol. 49 (1998) 189–193.
[17] M. Maeda, A. Itoh, Y. Kawase, Kinetics for aerobic biological treatment of o-cresol [39] D.L. Nelson, M. Cox, Lehninger Principles of Biochemistry, Macmillan Worth
containing wastewaters in a slurry bioreactor: biodegradation by utilizing waste Publishers, New York, 2000.
activated sludge, Biochem. Eng. J. 22 (2005) 97–103. [40] R.K. Singh, S. Kumar, S. Kumar, A. Kumar, Biodegradation kinetic studies for the
[18] G.S. Veeresh, P. Kumar, I. Mehrotra, Treatment of phenol and cresols in upflow removal of p-cresol from wastewater using Gliomastix indicus MTCC 3869, Bio-
anaerobic sludge blanket (UASB) process: a review, Water Res. 39 (2005) 154–170. chem. Eng. J. 40 (2008) 293–303.
[19] I.H. Farooqi, F. Basheer, T. Ahmed, Global NEST 10 (2008) 39–46. [41] J. Wu, H.-Q. Yu, Biosorption of 2,4-dichlorophenol by immobilized white-rot
[20] J. Bai, J.-P. Wen, H.-M. Li, Y. Jiang, Kinetic modeling of growth and biodegradation of fungus Phanerochaete chrysosporium from aqueous solutions, Bioresour. Technol.
phenol and m-cresol using Alcaligenes faecalis, Process Biochem. 42 (2007) 510–517. 98 (2007) 253–325.
[21] O. Bouallègue, R. Mzoughi, F.X. Weill, N. Mahdhaoui, Y. Ben Salem, H. Sboui, et al., [42] M. Tsezos, J.P. Bell, Comparison of the biosorption and desorption of hazardous
Outbreak of Pseudomonas putida bacteraemia in a neonatal intensive care unit, J. organic pollutants by live and dead biomass, Water Res. 23 (1989) 561–568.
Hosp. Infect. 57 (2004) 88–91. [43] M.H. El-Naas, S. Al-Zuhair, S. Makhlouf, Continuous biodegradation of phenol in a
[22] K.-C. Loh, B. Cao, Paradigm in biodegradation using Pseudomonas putida—a review spouted bed bioreactor (SBBR), Chem. Eng. J. 160 (2010) 565–570.
of proteomics studies, Enzyme Microb. Technol. 43 (2008) 1–12. [44] K. Vidya Shetty, R. Ramanjaneyulu, G. Srinikethan, Biological phenol removal
[23] Y. Wang, Y. Tian, B. Han, H.-b. Zhao, J.-n. Bi, B.-l. Cai, Biodegradation of phenol by free using immobilized cells in a pulsed plate bioreactor: effect of dilution rate and
and immobilized Acinetobacter sp. strain PD12, J. Environ. Sci. 19 (2007) 222–225. influent phenol concentration, J. Hazard. Mater. 149 (2007) 452–459.
[24] A. Idris, N.A.M. Zain, M.S. Suhaimi, Immobilization of Baker’s yeast invertase in [45] Z. Salehi, H. Yoshikawa, R. Mineta, Y. Kawase, Aerobic biodegradation of p-
PVA-alginate matrix using innovative immobilization technique, Process Bio- nitrophenol by acclimated waste activated sludge in a slurry bubble column,
chem. 43 (2008) 331–338. Process Biochem. 46 (2011) 284–289.

You might also like