You are on page 1of 8

J.

of Supercritical Fluids 57 (2011) 50–57

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


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

Hydrothermal gasification of olive mill wastewater as a biomass source in


supercritical water
Ekin Kıpçak, Onur Ö. Söğüt, Mesut Akgün ∗
Chemical Engineering Department, Yıldız Technical University, Davutpasa Campus, No. 127, 34210 Esenler, Istanbul, Turkey

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 hydrothermal gasification of biomass in supercritical water is investigated. The work
Received 1 October 2010 is of peculiar value since a real biomass, olive mill wastewater (OMW), is used instead of model biomass
Received in revised form 7 February 2011 compounds. OMW is a by-product obtained during olive oil production, which has a complex nature char-
Accepted 8 February 2011
acterized by a high content of organic compounds and polyphenols. The high content of organics makes
OMW a desirable biomass candidate as an energy source. The hydrothermal gasification experiments for
Keywords:
OMW were conducted with five different reaction temperatures (400, 450, 500, 550 and 600 ◦ C) and five
Biofuels
different reaction times (30, 60, 90, 120 and 150 s), under a pressure of 25 MPa. The gaseous products
Olive mill wastewater
Hydrothermal gasification
are mainly composed of hydrogen, carbon dioxide, carbon monoxide and C1 –C4 hydrocarbons, such as
Supercritical water methane, ethane, propane and propylene. Maximum amount of the gas product obtained is 7.71 mL per
mL OMW at a reaction temperature of 550 ◦ C, with a reaction time of 30 s. The gas product composition is
9.23% for hydrogen, 34.84% for methane, 4.04% for ethane, 0.84% for propane, 0.83% for propylene, 49.34%
for carbon dioxide, and 0.88% for minor components such as n-butane, i-butane, 1-butene, i-butene,
t-2-butene, 1,3-butadiene and nitrogen at this reaction conditions.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction high solubilities and complete miscibility with supercritical water.


Chemical reactions with high efficiencies can be obtained in the
Energy production based on fossil fuels causes high import case of a water–organic mixture without interfacial transport limi-
expenditures, environmental problems and a major decrease in tations [5,6]. Hence, supercritical water offers a control mechanism
the world’s nonrenewable energy reserves. Hence, a growing inter- depending on solubility, a low viscosity, excellent transport proper-
est has emerged on the development of new and efficient energy ties based on its high diffusion ability and new reaction possibilities
sources. Biomass energy is among these sources, having the poten- for hydrolysis or oxidation [7].
tial to provide the gradually increasing energy demand of the world. Using the aforementioned desirable properties of supercritical
As a renewable and environmental friendly energy source, it can be water, a lot of research has been made on the field of hydrother-
utilized with various energy conversion technologies. One of these mal gasification, as the method simply involves the hydrothermal
technologies, which have shown a great potential for the conver- conversion of a material into gaseous products at conditions that
sion of biomass with high moisture content to produce hydrogen exceed the critical point of water. Hydrothermal gasification pro-
and other gases, is hydrothermal gasification under supercritical cesses are currently being investigated both as a waste treatment
conditions. process and as a means of energy and materials recovery from
Water, being an important solvent in nature, has very unique biomass and biodegradable organic wastes. These processes have
characteristics as a reaction solvent under supercritical conditions. numerous advantages, one of which is water’s role as a reaction
The critical temperature and pressure of water are 374.8 ◦ C and solvent, eliminating the necessity for drying the biomass. This, in
22.1 MPa, respectively. At temperatures of near-critical and super- consequence, increases the thermal efficiency of the gasification
critical point, both the H3 O+ and the OH− ions are formed due to process [8]. Another advantage is supercritical water’s ability to
self-dissociation of water. Therefore, water can be considered to dissolve the organic biomass components, allowing for hydroly-
be both acidic and alkaline, so behaves as a catalytic precursor for sis to break down the polymeric biomass structure as opposed
acidic or basic reactions [1–4]. Besides, organic compounds have to pyrolysis used in conventional gasification. Also, a high level
of solid conversion is achieved, resulting in the formation of low
amounts of chars and tars [9–11]. Extensive investigations have
∗ Corresponding author. Tel.: +90 212 383 4759; fax: +90 212 383 4725. been conducted in the recent years on the hydrothermal gasifica-
E-mail address: akgunm@yildiz.edu.tr (M. Akgün). tion of biomass under supercritical conditions. The studies covered

0896-8446/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.supflu.2011.02.006
E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57 51

model compounds, such as glucose [6,12,13], cellulose [14–16], Table 1


The characteristics of the olive mill wastewater.
lignin [9,17] and some real biomass compounds [6,18–21].
In this study, the hydrothermal gasification of biomass in super- Parameters Value
critical water is investigated. The work is of peculiar value, as a pH 4.36
real biomass, namely olive mill wastewater is used in the experi- COD (g/L) 24
ments instead of model compounds. Olive mill wastewater (OMW) TOC (mg/L) 6138
is a by-product obtained during the production of olive oil. It is IC (mg/L) 30.1
TC (mg/L) 6168.1
generally composed of the water content of the olive fruit, water
TN (mg/L) 99.0
used to wash and process the olives, soft tissues from the olive TSM (g/L OMW) 10.35
pulp and a very stable oil emulsion [22]. The annual OMW pro- Ash (g/L OMW) 2.88
duction in the Mediterranean countries is estimated to range from Phenols (mg/L OMW) 3.70
Ca (mg/L OMW) 86.80
10 to 30 million m3 . This quantitative diversity results from dif-
Fe (mg/L OMW) 0.77
ferent factors such as the olive oil extraction method, cultivation K (mg/L OMW) 452.20
soil, the use of pesticides and fertilizers, olive harvesting time, Mg (mg/L OMW) 14.21
degree of ripening, duration of aging, olive variety and climatic Na (mg/L OMW) 22.27
conditions [23]. These factors cause qualitative variations in OMW Color Dark brown

composition as well. The typical OMW composition by weight is


83–96% water, 3.5–15% organic compounds and 0.5–2% mineral 2.2. Experimental procedure
salts. The organic fraction includes sugars, tannins, pectins, lipids,
organic acids, nitrogen compounds, polyalcohols and polyphe- The hydrothermal gasification of OMW was performed in a
nols. The inorganic fraction, on the other hand, includes chloride, coiled tubular reactor system (280 cm length × 4.35 mm i.d.), which
sulfate and phosphoric salts of potassium as well as calcium, is placed into a PID controlled split furnace (Protherm, model SPT
iron, magnesium, sodium, copper and traces of other elements 11/70/750). From the feeding pumps to the gas–liquid separator, all
[24]. wetted parts of the system, such as tubing, fittings, etc., were made
The complex composition stated above makes OMW a signifi- of 316 SS. The wastewater is pumped into the furnace using a high
cant pollution potential. Its polyphenol content is responsible for pressure pump (Autoclave Engineers, Series III pump). After exit-
several biological effects, as phenolic compounds of low molecu- ing the furnace, the effluent is cooled rapidly by passing through a
lar weight show toxicity on seed germination, aquatic organisms heat exchanger, and the reaction stops immediately. The possible
and bacteria. What is more, the dark color of OMW is attributed solid particles in the effluent, which would be formed, are removed
to these phenolic compounds, which can still be detected visu- with a 0.5 ␮m inline filter before the stream is depressurized by a
ally after a 1/1000 dilution in water [25]. The extremely high back-pressure regulator (BPR) (GO Regulator Inc.). The system pres-
content of organics is considered as a negative feature as well. sure is kept at a constant value (±0.1 MPa) by the BPR set at the
The chemical oxygen demand (COD) of OMW is 80–200 g/L, exit of the reactor and monitored through a digital pressure indica-
whereas its biological oxygen demand (BOD) is 12–63 g/L. It tor. The system temperature is also kept stable (±1 ◦ C) through the
should be noted that these values are around 200–400 times PID-controlled split furnace. The products leaving the BPR are sepa-
higher than those of a typical municipal sewage [26]. All of these rated by a gas–liquid separator. The gaseous effluent taken from the
characteristics, combined with the high amount of organics con- top portion of the separator is transported for quantitative analysis
tent can make OMW an energy source for energy production by a gas chromatograph capable of online sampling. The gas chro-
during its treatment by supercritical water oxidation (SCWO) matography instrument used in this work is a Clarus 500 model
[7]. PerkinElmer Arnel. On the other hand, the samples taken from the
liquid effluent leaving the system are sent for TOC–TN analysis,
with the aim of determining their organics content (Fig. 1).
2. Materials and methods
2.3. Analytical methods
2.1. Materials
The concentrations of the wastewater and liquid phase reactor
The olive mill wastewater that is used in the experiments was effluents were characterized in terms of total organic car-
supplied from an olive oil producing plant in Turkey, which was bon concentration (TOC). TOC analyses were performed using
roughly filtered and used without any dilution. The characteristics a total organic carbon–total nitrogen analyzer (HACH-LANGE
of the OMW used in this work are presented in Table 1. IL550 TOC–TN), which is based on combustion catalytic oxida-

Fig. 1. The schematic demonstration of the experimental setup (PI: pressure Indicator, FI: flow indicator).
52
Table 2
Experimental conditions and results.

Pressure Temperature Reaction Feed Exit TOC Gas CH4 (mol%) H2 (mol%) C2 H6 (mol%) C3 H8 (mol%) CO (mol%) CO2 (mol%) Other gases Closed C Energy (J/mL
(bar) (◦ C) time (s) flowrate (mg/L) product (mol%) balance (%) OMW)
(mL/dk) (mL/mL
OMW)

250 400 30 7.18 4,230.0 <dl 1.13 4.07 1.28 0.12 16.31 64.74 12.36 68.91 0.00
250 400 60 6.7 3,142.5 <dl 1.42 5.65 1.53 0.12 14.08 69.70 7.49 51.20 0.00
250 400 90 4.47 4,640.0 0.27 2.20 6.12 2.30 0.12 0.29 77.74 11.23 77.81 147.0

E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57


250 400 120 3.35 4,500.0 0.35 3.01 8.30 2.17 0.12 0.72 76.47 9.22 76.18 158.8
250 400 150 2.68 4,610.0 0.38 3.12 7.96 2.18 0.09 0.00 83.01 3.64 78.22 163.1
250 450 30 4.91 4,945.0 0.44 6.39 6.57 3.06 0.36 2.02 74.06 7.55 84.17 187.5
250 450 60 4.58 3,130.0 1.60 18.89 8.19 2.38 0.39 0.00 68.60 1.55 64.11 1,397.6
250 450 90 3.06 2,955.0 1.33 13.35 8.92 2.69 0.38 0.00 72.64 2.02 59.05 873.5
250 450 120 2.29 2,665.0 1.30 12.98 9.30 2.95 0.37 0.00 72.74 1.66 54.08 928.8
250 450 150 1.83 3,140.0 1.60 15.32 9.15 3.41 0.27 0.00 69.90 1.95 64.28 1,172.0
250 500 30 7.58 4,313.3 1.89 28.62 6.98 2.04 0.45 2.31 57.09 2.50 85.77 2,559.3
250 500 60 3.79 3,580.0 3.37 29.25 7.81 1.96 0.47 0.02 59.24 1.25 85.96 4,613.9
250 500 90 2.53 2,750.0 2.15 28.69 7.06 2.80 0.40 1.21 56.54 3.30 62.43 2,971.5
250 500 120 1.89 1,960.0 3.81 31.44 7.43 3.11 0.75 0.00 55.81 1.45 63.18 5,962.3
250 500 150 1.52 1,735.0 4.09 32.16 8.26 2.57 0.58 0.00 54.76 1.67 61.81 6,311.0
250 550 30 6.61 1,570.3 7.71 34.84 9.23 4.04 0.84 0.00 49.34 1.71 88.81 13,740.2
250 550 60 3.31 1,075.0 6.16 34.25 9.06 4.21 0.68 0.00 50.82 0.98 68.03 10,629.5
250 550 90 2.21 870.8 5.53 33.68 8.92 4.29 0.95 0.00 50.35 1.82 59.54 9,913.1
250 550 120 1.65 759.2 5.40 29.15 9.06 4.70 0.86 0.00 54.70 1.52 56.65 8,827.7
250 550 150 1.32 716.3 5.99 27.62 8.80 5.10 0.98 0.00 54.31 3.19 60.79 9,887.6
250 600 30 5.94 1,015.0 4.85 23.11 10.59 0.76 0.76 0.00 56.70 1.89 56.31 7,016.3
250 600 60 2.97 840.5 5.91 22.94 10.79 0.72 0.72 0.00 56.91 2.20 62.16 8,990.4
250 600 90 1.98 770.1 2.28 22.34 10.33 0.69 0.69 0.00 58.58 2.16 31.24 3,327.1
250 600 120 1.48 510.8 2.87 18.34 9.69 0.55 0.55 0.00 66.13 1.22 31.86 2,934.2
250 600 150 1.19 431.4 6.88 17.14 9.07 0.56 0.56 0.00 68.22 1.15 63.45 7,802.6
100 550 60 1.18 874.3 3.89 14.40 9.28 6.18 0.65 0.33 66.99 2.17 46.14 3,781.2
150 550 60 1.84 903.5 4.71 18.47 7.29 6.03 0.52 0.00 66.20 1.49 53.34 6,062.2
200 550 60 2.55 1,080.0 5.59 22.73 8.52 5.59 0.73 0.00 60.27 2.16 63.44 8,345.8
250 550 60 3.31 1,230.0 5.56 23.93 7.70 5.93 0.84 0.00 60.36 1.24 65.63 8,189.7
300 550 60 4.11 1,270.0 5.20 24.61 8.51 5.37 0.45 0.00 60.21 0.85 63.33 7,245.6

dl: detection limit.


E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57 53

Fig. 2. Compositions of gaseous effluent during hydrothermal gasification of the OMW at 30 s of the reaction time.

tion method, using a highly sensitive multi-channel non-dispersive times of 30, 60, 90, 120 and 150 s, under a constant pressure of
infrared detector (NDIR). Standard solutions for the calibration 250 bar as described in detail in Table 2. To examine the effect
were prepared by using potassium hydrogen phthalate (Acros). All of pressure on the gas yield and composition, five runs were
the reagents were pure for analytical use. In order to provide pre- performed in the pressure range of 100–300 bar at a constant tem-
cise data, the samples were analyzed in triplicate, and the averages perature of 550 ◦ C and reaction time of 60 s (Table 2). The reaction
are reported as results. Total phenolic content of the OMW was times at supercritical conditions were calculated as below:
determined using the Folin–Ciocalteau method [27]. Physicochem-
ical properties of the wastewater such as COD, total solids and ash VReactor SC (P, T )
= (1)
in Table 1 were determined according to Standard methods [28]. F L
The gas samples were analyzed according to ASTM D1945, ASTM
where VReactor is the reactor volume, SC (P,T) and L are the fluid
D1946, ASTM D2597 and DIN 51872-4 methods by PerkinElmer,
densities in g mL−1 under reaction conditions and at feed pump
Clarus 500 RGA-1115 model GC coupled with one FID and two TCD
conditions, respectively. F is the volumetric flow rate in mL s−1 of
detectors. Ca, Mg, Na, K and Fe contents of the OMW were measured
the OMW. The reaction media at every experimental run consisted
with ICP-OES instrument (PerkinElmer, Optima 2100 DV) which
of dilute mixtures of organic wastes in water, and the fluid mixture
was equipped with an autosampler (PerkinElmer AS-93), Mea-
densities in the feed and inside the reactor were assumed to be
surement conditions: power 1.45 kW, plasma flow 15.0 L min−1 ,
those of pure water [29].
auxiliary flow 0.8 L min−1 , nebulizer flow 1 L min−1 .
It was observed that the reaction temperature had an important
role on the hydrothermal decomposition of the OMW contain-
3. Results and discussion ing organic compounds in supercritical water. While some organic
compounds in the OMW are transformed into intermediates, some
The hydrothermal gasification experiments were carried out in organic compounds and intermediates are directly destroyed into
the temperature range of 400–600 ◦ C, with five different reaction the final products in the reactor. During the hydrothermal decom-
54 E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57

Fig. 3. The effect of reaction temperature and time on the gas yield per mL OMW Fig. 4. The change of the TOC removal with reaction temperature and time.
fed to the reactor.

position, the organic molecules in the wastewater are converted


into smaller intermediate products (such as hydrogen, carbon
monoxide, methane, ethylene, ethane, propylene and propane) and
final products (such as carbon dioxide, nitrogen and water), which
leave the system in gas phase. These intermediate products in the
gas phase effluent at 400, 450, 500 and 550 ◦ C are shown compar-
atively in Fig. 2. Although the hydrothermal reaction is carried out
in the absence of the oxidant, mol% of carbon dioxide is very high
as shown in Fig. 2. This situation shows that some of the water in
supercritical conditions dissociates to H+ and OH− ions, and reacts
with the organics in the wastewater. A possible mechanism for the
reaction can be written as below:
m 
Cn Hm + nH2 O  nCO + + n H2 (2)
2
C + H2 O  CO + H2 (3)

CO + H2 O  CO2 + H2 (4)

CO + 3H2  CH4 (or C2 H6 , C2 H4 , C3 H8 , C3 H6 ) + H2 O (5)

The maximum gaseous effluent flow rate was obtained as


Fig. 5. The effect of reaction temperature and time on the total biofuel content in
51 mL/min at 550 ◦ C with a reaction time of 30 s. However, the
the gas products.
measured gas flow rates on the system can be quite misleading,
as the OMW flow rates fed to the reactor were calculated consid-
ering the water density at supercritical conditions, in order to keep phase. However, while a part of this organic carbon which trans-
the reaction time constant (Eq. (1)). In other words, the OMW feed forms into gas products that consist of combustible components
flow rates are greater at lower temperatures. Hence, it is more con- called as biofuel, a part of it transforms into carbon dioxide. On
venient to evaluate the gaseous effluent amounts with respect to the other hand, the carbon balance closed within 32–86% in the
the unit amount of OMW fed to the system (Fig. 3). As it can be hydrothermal gasification of the OMW, as shown in Table 2. Devia-
seen from Fig. 3, the amount of gaseous effluent (and consequently tion of carbon balance changes between 14% and 68%. As indicated
the efficiency of the system) increases with increasing tempera- by Klingler and Vogel [30], the missing carbon in the balance may
ture. This situation can also be verified when the change of total be owing to by-products such as sticky brown humin polymers
organic carbon (TOC) is taken into account. The TOC content of formed or solid carbon formed because of carbonization, but cannot
OMW (which is initially 6138 mg/L) decreased with temperature be analyzed, because carbon nano-particles were observed during
and reaction time. This is an expected occasion, as the organic con- cleaning of the filter after each run. The sticky brown humins were
tent of OMW reacts more and gets through the gaseous phase, as generally in the liquid effluent, and got smeared on the walls of the
the reaction temperature and duration increase. This situation can sample tubes.
also be seen through Fig. 4, which shows the change of TOC con- Fig. 5 shows changing of total mol% of combustible components
version at different reaction conditions. At elevated temperatures with increasing temperature and reaction time in the gaseous efflu-
and reaction times, more than 90% of the organic carbon present ent. While total mol% of combustible components increases up to
in OMW is removed from the wastewater enabling its treatment. about 50% with increasing temperature at a reaction time of 30 s, it
In other words, the organic carbon is transferred into the gaseous decreases slightly with increasing reaction time. These results show
E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57 55

Fig. 6. The change of the methane content in the gas products with reaction tem-
Fig. 8. The change of the ethane content in the gas products with reaction temper-
perature and time.
ature and time.

that the gasification reaction has to be carried out at higher tem-


the highest amount methane (35%) is obtained at a reaction temper-
peratures than 500 ◦ C and lower reaction times than 30 s in order
ature of 550 ◦ C and a reaction time of 30 s. At low temperatures, the
to produce much more combustible matter.
methane percentage increases with increasing reaction time. How-
When the gas composition is evaluated, though it may differ
ever, at higher temperatures, the amount of methane is inversely
at different reaction conditions, is seen that the content mainly
proportional with the reaction time and temperature above 550 ◦ C.
includes methane, ethane, hydrogen, carbon dioxide and carbon
This situation indicates the degradation of the formed methane into
monoxide. Besides these products, scarce amounts of propane,
carbon dioxide and water with time and elevated temperatures. A
proylene, nitrogen, n-butane, i-butane, 1-butene, i-butene, t-2-
similar approach can also be made for hydrogen and ethane pro-
butene and 1,3-butadiene were also obtained. Besides, products
duction as well. From Figs. 7 and 8, it is seen that the highest amount
such as an oil product, char, water-soluble products, and water
of hydrogen and ethane is obtained at 600 ◦ C, at a reaction time of
formed in liquid effluent, which were not analyzed. Figs. 6 and 7
30 s.
show the change of the two main gaseous products as biofuel,
Fig. 9 shows the change of carbon dioxide amounts in the
methane and hydrogen, at different reaction conditions, respec-
gaseous product at different reaction conditions. From Fig. 9, it is
tively.
seen that the amount of carbon dioxide produced is maximized at
As it can be seen from Fig. 6, the amount of methane in the gas
effluent increases dramatically with increasing temperature, and

Fig. 7. The change of the hydrogen content in the gas products with reaction tem- Fig. 9. The change of the carbon dioxide content in the gas products with reaction
perature and time. temperature and time.
56 E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57

Fig. 10. The effect of reaction pressure on the TOC removal and gas product yield during hydrothermal gasification of the OMW at 550 ◦ C for 60 s of reaction time.

lower temperatures and longer reaction times. As the temperature


increases the carbon dioxide percentage begins to decrease, having
a minimum value at 550 ◦ C. When the quantity of carbon monoxide
is taken into account, it is seen that the most favorable temperature
is 400 ◦ C (Table 2). As the temperature or the reaction time begins
to increase, carbon monoxide formation decreases dramatically, or
CO is an intermediate product, and transforms quickly into CO2 at
higher temperatures.
Same results has also been obtained by D’jesus et al. [19],
although biomass source used is different. They have reported that
the methane and hydrogen production increase with increasing
temperature, whereas the CO2 and CO content in the gas effluent
decrease. They have also stated that this increase in methane and
hydrogen concentrations with temperature could be explained by
the formation of methane from both CO/CO2 and directly from the
reacting biomass. On the contrary, there are also some studies that
the methane decreases with increasing temperature, whereas the
CO2 and CO contents increase in the gas effluent [31,32].
From the aforementioned results, it can be concluded that the Fig. 11. The effect of reaction pressure on the composition of the gas effluent during
most favorable reaction temperature for the hydrothermal gasifi- hydrothermal gasification of the OMW at 550 ◦ C for 60 s of reaction time.
cation of olive mill wastewater under supercritical conditions is
550 ◦ C, with a beneficial reaction time of 30 s. The energy values of
produced gases at various reaction conditions are calculated and ent also decreases with decreasing system pressure, against what
given in Table 2. These energy values also confirm that the most would be expected (Fig. 10). The reason of this is that smut forms
favorable reaction condition is 550 ◦ C for a reaction time of 30 s, because of carbonization due to decreasing the fluid density at low
and it forms a gas composition which has an energy content up to system pressures. Hence, the carbon content in the OMW settles
13.74 kJ per mL OMW at this reaction conditions. to the reactor inner wall as smut, instead of transforming into gas
products. This situation causes system cloggages with time. On the
other hand, decreasing system pressure causes the carbon dioxide
3.1. Pressure effect on the hydrothermal gasification of OMW
content to increase and the methane content to decrease in the
gas effluent (Fig. 11). This tendency has also been mentioned by
To examine the effect of reaction pressure on the hydrothermal
Yan et al. [33] and Gadhe and Gupta [34] for the gasification of the
gasification of the OMW, a series of experiments was performed
model glucose and methanol solution in SCW, respectively.
at 550 ◦ C for a residence time of 60 s. As shown in Fig. 10, the TOC
content of liquid effluent decreases with decreasing reaction pres-
sure from 300 to 100 bar. This situation can be explained with fluid 4. Conclusion
density which is a function of temperature and pressure, as dis-
cussed in detail at a previous work [7]. However, although the TOC Supercritical water as a reaction medium appears to be
content of the liquid effluent decreases, the flow rate of gas efflu- attractive for the treatment and the gasification of industrial or
E. Kıpçak et al. / J. of Supercritical Fluids 57 (2011) 50–57 57

agricultural wastewaters, which contain high amount of organic [14] X. Hao, L. Guo, X. Zhang, Y. Guan, Hydrogen production from catalytic gasi-
carbon. This technology enables the production of combustible fication of cellulose in supercritical water, Chemical Engineering Journal 110
(2005) 57–65.
gases, while the treatments of the aforementioned wastewaters [15] Y. Matsumura, H. Nonaka, H. Yokura, A. Tsutsumi, K. Yoshida, Co-liquefaction
are carried out in the same process. In this study, the gasification of coal and cellulose in supercritical water, Fuel 78 (1999) 1049–1056.
of organics in the OMW, which is detrimental for the soil and the [16] M. Watanabe, H. Inomata, K. Arai, Catalytic hydrogen generation from biomass
(glucose and cellulose) with ZrO2 in supercritical water, Biomass and Bioenergy
aquatic life, was performed at sub- and supercritical conditions of 22 (2002) 405–410.
water. The experiments show that the amount of the gas prod- [17] A. Yamaguchi, N. Hiyoshi, O. Sato, K.K. Bando, M. Osada, M. Shirai, Hydrogen
uct increases with increasing system temperature. However, long production from woody biomass over supported metal catalysts in supercritical
water, Catalysis Today 146 (2009) 192–195.
reaction times cause a decrease in the combustible compounds of
[18] M.J. Antal, S.G. Allen, D. Schulman, X.D. Xu, R.J. Divilio, Biomass gasification in
the gas effluent, since these materials transform into carbon diox- supercritical water, Industrial and Engineering Chemistry Research 39 (2000)
ide and water at high temperatures. The most favorable reaction 4040–4053.
condition is 550 ◦ C for reaction time of 30 s, and a gas composition [19] P. D’Jesus, N. Boukis, B. Kraushaar-Czarnetzki, E. Dinjus, Gasification of corn
and clover grass in supercritical water, Fuel 85 (2006) 1032–1038.
that has energy content up to 10 kJ per mL OMW is formed at this [20] Y.J. Lu, H. Jin, L.J. Guo, X.M. Zhang, C.Q. Cao, X. Guo, Hydrogen production by
reaction conditions. biomass gasification in supercritical water with a fluidized bed reactor, Inter-
national Journal of Hydrogen Energy 33 (2008) 6066–6075.
[21] B. García-Jarana, J. Sánchez-Oneto, J.R. Portela, E. Nebot, E.M. de la Ossa, Super-
Acknowledgement critical water gasification of industrial organic wastes, Journal of Supercritical
Fluids 46 (2008) 329–334.
This work has been supported by The Scientific and Technolog- [22] F. Raposo, R. Borja, E. Sanchez, M.A. Martin, A. Martin, Performance and kinetic
evaluation of the anaerobic digestion of two-phase olive mill effluents in reac-
ical Research Council of Turkey (TUBITAK, project no: 108M546). tors with suspended and immobilized biomass, Water Research 38 (2004)
2017–2026.
References [23] R. Borja, B. Rincon, F. Raposo, Anaerobic biodegradation of two-phase olive
mill solid wastes and liquid effluents: kinetic studies and process per-
formance, Journal of Chemical Technology and Biotechnology 81 (2006)
[1] N. Akiya, P.E. Savage, Roles of water for chemical reactions in high-temperature 1450–1462.
water, Chemical Reviews 102 (2002) 2725–2750. [24] G. Tziotzios, S. Michailakis, D.V. Vayenas, Aerobic biological treatment of olive
[2] G. Brunner, Near critical and supercritical water. Part I. Hydrolytic and mill wastewater by olive pulp bacteria, International Biodeterioration and
hydrothermal processes, Journal of Supercritical Fluids 47 (2009) 373–381. Biodegradation 60 (2007) 209–214.
[3] A. Kruse, E. Dinjus, Hot compressed water as reaction medium and reactant. 2. [25] A. Fiorentino, A. Gentili, M. Isidori, M. Lavorgna, A. Parrella, F. Temussi, Olive oil
Degradation reactions, Journal of Supercritical Fluids 41 (2007) 361–379. mill wastewater treatment using a chemical and biological approach, Journal
[4] M.D. Bermejo, M.J. Cocero, Supercritical water oxidation: a technical review, of Agricultural and Food Chemistry 52 (2004) 5151–5154.
AIChE Journal 52 (2006) 3933–3951. [26] K. Al-Malah, M.O.J. Azzam, N.I. Abu-Lail, Olive mills effluent (OME) wastewater
[5] Y. Calzavara, C. Joussot-Dubien, G. Boissonnet, S. Sarrade, Evaluation of biomass post-treatment using activated clay, Separation and Purification Technology
gasification in supercritical water process for hydrogen production, Energy 20 (2000) 225–234.
Conversion and Management 46 (2005) 615–631. [27] V.L. Singleton, J.A. Rossi, Colorimetry of total phenolics with
[6] P.T. Williams, J. Onwudili, Subcritical and supercritical water gasification phosphomolybdic–phosphotungstic acid reagents, American Journal of
of cellulose, starch, glucose, and biomass waste, Energy & Fuels 20 (2006) Enology and Viticulture 16 (1965) 144–158.
1259–1265. [28] A.D. Eaton, L.S. Clesceri, A.E. Greenberg, Standard Methods for the Examina-
[7] H Erkonak, O.Ö. Söğüt, M. Akgün, Treatment of olive mill wastewater by super- tion of Water and Wastewater, 19th ed., American Public Health Association,
critical water oxidation, Journal of Supercritical Fluids 46 (2008) 142–148. Washington, DC, 1995.
[8] A.D. Taylor, G.J. DiLeo, K. Sun, Hydrogen production and performance of nickel [29] W. Wagner, A. Prub, The IAPWS formulation 1995 for the thermodynamic prop-
based catalysts synthesized using supercritical fluids for the gasification of erties of ordinary water substance for general and scientific use, Journal of
biomass, Applied Catalysis B: Environmental 93 (2009) 126–133. Physical and Chemical Reference Data 31 (2002) 387–535.
[9] T. Furusawa, T. Sato, H. Sugito, Y. Miura, Y. Ishiyama, M. Sato, N. Itoh, N. [30] D. Klingler, H. Vogel, Influence of process parameters on the hydrothermal
Suzuki, Hydrogen production from the gasification of lignin with nickel cata- decomposition and oxidation of glucose in sub- and supercritical water, Journal
lysts in supercritical water, International Journal of Hydrogen Energy 32 (2007) of Supercritical Fluids 55 (2010) 259–270.
699–704. [31] Y. Lu, L. Guo, X. Zhang, Q. Yan, Thermodynamic modeling and analysis of
[10] Y. Matsumura, M. Sasaki, K. Okuda, S. Takami, S. Ohara, M. Umetsu, T. Adschiri, biomass gasification for hydrogen production in supercritical water, Chemical
Supercritical water treatment of biomass for energy and material recovery, Engineering Journal 131 (2007) 233–244.
Combustion Science and Technology 178 (2006) 509–536. [32] A. Kruse, D. Meier, P. Rimbrecht, M. Schacht, Gasification of pyrocatechol in
[11] J. Yanik, S. Ebale, A. Kruse, M. Saglam, M. Yüksel, Biomass gasification in super- supercritical water in the presence of potassium hydroxide, Industrial and
critical water. Part 1. Effect of the nature of biomass, Fuel 86 (2007) 2410–2415. Engineering Chemistry Research 39 (2000) 4842–4848.
[12] X.H. Hao, L.J. Guo, X. Mao, X.M. Zhang, X.J. Chen, Hydrogen production from [33] Q. Yan, L. Guo, Y. Lu, Thermodynamic analysis of hydrogen production from
glucose used as a model compound of biomass gasified in supercritical water, biomass gasification in supercritical water, Energy Conversion and Manage-
International Journal of Hydrogen Energy 28 (2003) 55–64. ment 47 (2006) 1515–1528.
[13] P.T. Williams, J. Onwudili, Composition of products from the supercritical water [34] J.B. Gadhe, R.B. Gupta, Hydrogen production by methanol reforming in super-
gasification of glucose: a model biomass compound, Industrial and Engineering critical water: suppression of methane formation, Industrial and Engineering
Chemistry Research 44 (2005) 8739–8749. Chemistry Research 44 (2005) 4577–4585.

You might also like