Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 14

Persistent Organic Pollutants

History, Properties and Removal


Methods of Phenolic Compounds

Yesim GUCBILMEZ, Department of Chemical Engineering/ESTU, Eskisehir,


Turkey, ygucbilmez@eskisehir.edu.tr

Abstract

In this review article, phenol and chlorophenols are investigated in terms of their
production histories, physiochemical properties, pollution resources and removal
methods. It is seen that both phenol and chlorophenols are highly toxic compounds,
produced from natural and anthropogenic sources, which are hazardous to both humans
and the environment even at very low concentrations. The typical industries which
produce phenol and chlorophenol pollution are petrochemical, textile, plastics, resin
manufacturing, dye, pharmaceutical, iron and steel, pulp and paper industries as well as
the petroleum refineries, and coal gasification operations. Phenol is a highly corrosive
and nerve poisoning agent. It causes harmful side effects such as sour mouth, diarrhea
and impaired vision. It is also toxic for the eco-system with toxicity levels ranging
between 10-24 mg/L for humans, 9-25 mg/l for fish and lethal blood concentration
between 150-mg/100 ml. Chlorophenols found in natural waters or drinking water also
cause serious health problems such as histopathological alterations, genotoxicity,
mutagenicity, and carcinogenicity among others. Due to the aforementioned reasons, the
phenolic compounds in wastewaters or in drinking water must be removed using a
suitable wastewater treatment method such as adsorption, extraction, electrochemical
oxidation, biodegradation, wet air oxidation photo catalysis or enzyme treatment among
others.

Keywords: Phenol, chlorophenols, wastewater treatment, phenolic compounds,


phenolics, organic pollutants

1. Introduction

Due to the advancement of technology and rapid industrial growth, water resources
all over the world are under threat. In general, developed countries suffer from chemical
discharge problems while developing countries suffer from agricultural sources. In most
wastewaters and in drinking waters, one or more of the following toxic organic pollutants
may occur [1]: Organochlorines, chlorobenzenes (CBs), polychlorinated dibenzo-p-
dioxins (PCDDs) and polychlorinated dibenzofuranes (PCDFs), phenols, chlorophenols
and other phenol-derived compounds.
Phenolic compounds are phytochemicals found in nature which cannot be synthesized
in the human body. Thus, they are mainly taken from food and medicinal herbs. They are
found in most plant tissues, including fruits and vegetables [2]. They exist in water
systems due to the discharge of contaminated streams from industrial, agricultural and
domestic activities as well as a result of natural phenomena. They cause both severe and
long‐lasting effects on both humans and animals including damage to red blood cells and
the liver. Their interaction with microorganisms, inorganic and other organic compounds
in water can produce substituted compounds or other moieties, which may be as toxic as
the original phenolic compounds [3].
Chlorophenols (CPs) are common contaminants which can be found in surface,
ground, and drinking water [4-8]. Thermal and chemical degradation of chlorophenols
lead to the formation of harmful substances which cause public health problems such as
genotoxicity, mutagenicity, and carcinogenicity among others. The transformation of
chlorophenols also can lead to the formation of electrophilic metabolites that may bind
and damage the DNA or gene products [5, 6].

2. History of Phenol and Chlorophenols

The academic and industrial chemist F. F. Runge was born on February 8, 1794 in
Billwerder near Hamburg. Runge isolated phenol from coal tar in 1834 in impure form
and called it “carbolic acid” (Karbols¨aure) or “coal oil acid” (Kohlen¨ols¨aure) [7-9].
Phenol was first used as a disinfectant in 1865 by the British surgeon Joseph Lister at
the Glasgow University, Scotland, for sterilizing wounds, surgical dressings and
instruments [10]. However, the phenol sprays used during the surgeries were also
dangerous to the mucous membrane of the lungs when inhaled. Thus, by 1890, the phenol
spray was abandoned by the medical community [11].
During the Boer War, England placed an embargo on phenol, which caused a
shortage on the continent. As a result, Raschig Process was developed in Germany in the
year 1940 for the manufacture of synthetic phenol on a large scale [12, 13].
During World Wars I and II, the Sulfonation Process was further improved and other
processes such as Chlorination Process and Raschig Process were first commercialized
[14]. In 1924, Dow Chemical started to commercialize synthetic phenol based on the
direct chlorination of benzene to chlorobenzene [11].
Since the mid-20th century, patents related to phenol production have primarily
focused on the improvements to the existing processes rather than developing alternative
processes. One example is the European patent filed by the Mitsui Petrochemical Limited
(LTD) in 1989 which proposed a method to reduce the byproduct acetone through the
integration of a recycle loop [15, 16].
Presently, phenol is used in the production of phenolic and epoxy resins [17, 18],
plastics [19], plasticizers [20], polycarbonates [21], nylon [22], dyes [23, 24]
disinfectants [25], herbicides [26], polymers, drugs, pesticides [27] wood preservatives
[28] and fungicides [29].
Chlorophenols are produced by the electrophilic halogenation of phenol with chlorine.
There are 5 basic types and 19 different chlorophenols. Some chlorophenols are used as
pesticides and herbicides. Others are used as antiseptics and disinfectants [30].
Chlorophenols are very toxic and their presence is dangerous for humans as well as
aquatic life. They are found in the wastewaters of textile, pharmaceutical, petrochemical,
pesticide, paper and other industries [31].
The following processes have been mainly used in the industry for the production of
phenol [10, 32] up to date: Sulphonation of benzene and production of phenol by heating
the benzenesulphonate in molten alkali hydroxide [33], Chlorination of benzene and
alkaline hydrolysis of the chlorobenzene, C-Chlorination of benzene and catalytic
saponification by Cu in the steam hydrolysis of the chlorobenzene [34, 35] (Raschig
process, Raschig–Hooker Process), Alkylation of benzene with propene to
isopropylbenzene (cumene), Oxidation of cumene to the corresponding tert-
hydroperoxide and cleavage to phenol and acetone (Hock Process), Toluene oxidation to
benzoic acid and subsequent oxidizing decarboxylation to phenol (Dow Process),
Dehydrogenation of cyclohexanol–cyclohexanone mixtures [10].
Among these processes, only the Hock and Toluene Oxidation Processes are
important industrially. The other processes were discarded for economic reasons [10].
The global phenol market size reached a value of US Dollars (USD) 23.17 billion in the
year 2020. The phenol market is further expected to grow at a Compound Annual Growth
Rate of (CAGR) of 5.3% between the years 2021 and 2026 to reach to a value of about
USD 30 billion by the year 2026 [36].
In the case of chlorophenols, their production and usage have led to the entry of
persistent toxic components into the water systems which are resistant to biodegradation.
However, microorganisms exposed to these pollutants have evolved the ability to
biodegrade some of them and the routes by which a given compound is degraded are
determined by the physical, chemical, and microbiological components of the water
system under question. Understanding the genetic basis of catabolism of chlorophenols
may improve the efficiency of naturally occurring microorganisms or help to raise new
microorganisms which are able to degrade chlorophenols [37].

3.Physiochemical Properties of Phenol and Chlorophenols

Phenol is an odorous chemical compound, found either as a colourless liquid or


white solid at room temperature and may be highly toxic and caustic. Phenols are similar
to alcohols but form stronger hydrogen bonds. Thus, they are more soluble in water than
alcohols and have higher boiling points [38]. They are widely used as a raw material in
the manufacture of phenolic resins, automotive parts, nylon, epoxy resins, polycarbonate
engineering thermoplastics, wood preservatives, heavy duty surfactants, pharmaceuticals,
disinfectants, tank linings and coating materials among others. They are also widely used
in household products as well; for instance, phenol-derived n-hexylresorcinol, is used in
cough drops and other antiseptic applications and butylated hydroxytoluene (BHT) is a
common antioxidant in foods. Also, in the dye industry, substituted phenols are used to
make intensely coloured azo dyes [38, 39].

Chlorophenols, produced by chlorinating phenol or hydrolysing chlorobenzenes,


contain the benzene ring, the –OH group and chlorine atoms. All chlorophenols, except
2-chlorophenol, are solids with their melting points between 33 and 191 oC. They are
weakly acidic and their acidity is slightly lower than phenols. In reactions with alkaline
metals such as sodium and potassium in the aquatic environment, they yield salts which
are highly soluble in water. Their toxicity depends on the degree of chlorination and the
position of the chlorine atoms relative to the hydroxyl group decreasing with the number
of chlorine substituents [40].

4.Sources of Phenol and Chlorophenol Pollution

The existence of phenolic compounds in wastewaters can be attributed to both


natural and anthropogenic activities. Natural sources include decomposition of dead
plants and animals, synthesis by microorganisms and plants in the aquatic environment.
Anthropogenic sources, on the other hand, include industrial, domestic, agricultural and
municipal activities [10].

Phenol enters water systems in effluents from major industries such as


petrochemical, textile, plastics, resin manufacturing, dye, pharmaceutical, iron and steel,
pulp and paper as well as petroleum refineries, and coal gasification operations. It is very
important to remove phenols and aromatic compounds from industry effluents before
discharging them because of their toxicity to aquatic organisms [41, 42].
The CPs are among the most important environmental pollutants. They are
commonly used in the production of paper and pesticides but also as intermediates in the
production of dyes, plastics and pharmaceuticals [43-45]. These kinds of applications
often lead to wastewater and groundwater contaminations. In addition, as a result of tap
water chlorination treatment, CPs have also been detected in drinking water [46].

5.Removal Methods of Phenol and Chlorophenols

Phenol is a highly corrosive and nerve poisoning agent. It causes harmful side
effects such as sour mouth, diarrhea and impaired vision. It is also toxic for fish and the
toxicity levels usually range between 10-24 mg/L for humans and 9-25 mg/l for fish
while the lethal blood concentration is around 150-mg/100 ml [47].
The available removal methods used for phenol can be separated into two main
groups: conventional and advanced. Conventional methods include steam distillation,
extraction, adsorption, ion exchange [42] and advanced methods include wet air
oxidation, catalytic wet air oxidation, ozonation, membrane processes, electrochemical
oxidation, biological processes/biodegradation and enzymatic treatment among others
[42, 48].
It is possible to carry out steam distillation processes, based on the relative
volatility of phenol, for phenol removal from aqueous solutions. The phenol–water
mixture has a minimum azeotrope at 9.21% (w/w) phenol [42, 49-51]. Thus, azeotropic
distillation or steam stripping can purify the water from phenol impurities and an effluent
concentration as low as 0.01 mg/L can be obtained [48].
Liquid-liquid extraction (LLE), is a standard, non-destructive process for the
treatment of phenolic compounds, suitable over a wide range of phenol concentrations
and is cost-effective in some cases. Benzene and butyl acetate have been used as solvents
in this process in the past, but the most common solvent at present is di-isopropyl ether
(DIPE), which is used in the phenosolvan process [97, 98]. For the extraction process, the
solubility of the preferred solvent in water should be tolerable eliminating the need for
further purification steps. The selectivity of the solvent depends mainly on the type of
the solvent, temperature of the system, and the concentration of the phenol in the
wastewater [97].

Adsorption method for the removal of phenols from water is effective on a large
scale from low concentrations to high concentrations, depending on the economics and
recycling and the adsorbent. Activated carbon (AC) is the most used adsorbent in the
industry. It is expensive but has been proven to be effective for the removal of even trace
levels of organic compounds. Therefore, new options are being developed for AC
including impregnation with nanoparticles; using different sources of carbon, different
activation methods, Carbon NanoTubes (CNTs), graphene‐based materials as well as
substitution with low cost biosorbents such as chitin/chitosan [98–100].

Phenol can also be removed by using strong-base anion exchangers [101, 102].
In a sample literature study, it was observed that the studied amberlite adsorption
capacity was higher than that of known adsorbents. A theoretical model based on
molecular adsorption and ion exchange was developed to fit the experimental data and to
obtain the parameters related to both phenomena. It was seen that uptake onto Amberlite
occured by “only” adsorption at acidic pH and by both “both” adsorption and ion
exchange at alkaline pH. The phenol removal increased at pH values from 9 to 14 and
remained constant below 8. The developed theoretical model allowed the phenol removal
to be determined depending on the value of pH and the amount of resin [102].

Chemical oxidation processes can decompose phenolic compounds into smaller


molecules which are less toxic and easier to treat or mineralize [103, 104]. Among the
chemical oxidation processes, advanced oxidation processes (AOPs) such as the Fenton
process, ozonation, photolysis or combinations of these are recommended for low
concentrations. Incineration, on the other hand, is suitable for very high concentrations
(for Chemical Oxygen Demand (COD)≥100 g/L). However, it is not encouraged anymore
since it is not eco-friendly [105].

The wet air oxidation (WAO) reaction is recognized as the cleanest technique of
all because no additive is added to cause secondary pollution. At reasonable temperatures
(175°C–320°C) and high pressures of 2.17–20.71 MPa, wastewater is kept in the liquid
phase and the organics are oxidized into small organic acids which are likely to be
biodegradable [106]. Usually, it is suitable for a wide range of wastewaters with initial
COD values between 20-200 g/L [107]. It is also recognized as a potential pre-treatment
before biodegradation (BOD) to detoxify hazardous pollutants and satisfy the
biodegradation requirements [108].

Catalytic Wet Air Oxidation (CWAO) offers an alternative solution to treating


refractory wastewaters. CWAO gained a lot of interest over the past two decades due to
its ability to oxidize toxic wastewaters and complete mineralization [109-113]. Moreover,
it is a heterogeneous process, so an additional stage of separation of the catalyst is not
required in most instances making the process more economic [114].

In literature studies, it was shown that, during the process, phenol breaks down
to form aromatics (hydroquinone, catechol, and benzoquinone) and carboxylic acids
(oxalic and formic acid). Furthermore, traces of malonic, maleic and fumaric acid were
also detected [115]. Eftaxias et al. [116] showed that the Langmuir Hinshelwood Model
could be used in order to define the reaction mechanism. Using this model, all
experimental concentration profiles were accurately described with mean deviations less
than 8%. The use of Trickle Bed Reactors (TBR) is common in the CWAO of phenolic
compounds; however liquid maldistribution may cause hot spot formation, reactor
thermal runaway and deactivation of the catalyst [115].

Chemical adsorption and oxidation are the most common methods used for
phenol removal from wastewaters but still they have some drawbacks like higher energy
consumption, high operating cost, and being unsustainable. As an alternative, ozone-
based technology has been widely used to degrade organic pollutants since ozone has a
high oxidizing capability. However, single ozonation exhibits very low efficiency,
requires a large amount of ozone, has low solubility in water, low reaction rate with
organic compounds, and a slow mineralization rate. Hence, in order to overcome this
problem, a catalyst must be used in the process in order to increase the degradation
efficiency and to reduce the amount of ozone used. The process can be either
homogeneous or heterogeneous. Heterogeneous catalytic ozonation has the advantage of
producing no secondary pollutants and also it is easier to regenerate the catalyst. The
removal of phenol by heterogeneous catalytic ozonation has been studied by many
researchers. Of the many types of catalysts that have been used, zeolites have a high
potential due to their large specific surface areas, good thermal and chemical stabilities
and controllable hydrophilic/hydrophobic properties [117].

In late years, researchers have shifted their focus towards the advanced
membrane processes because these processes are eco-friendly, cost-effective and no
chemical additives are needed. However, the main drawback of these processes is the
flux decline caused by membrane fouling [89]. Membrane fouling is the phenomenon
caused by the accumulation of biological inorganic, organic and colloidal species onto
the membrane surface or within its pores. It results in flux decline and a rapidly
increasing transmembrane pressure during operation, and the possible deterioration of
mechanical strength [118, 119].
Polymers and ceramics are commonly used in the fabrication of membranes
which are used for filtration. Polymer membranes are relatively cheap, but have lower
tolerance to harsh conditions, whereas ceramic membranes have higher mechanical
strength which results in greater resistance to harsh environments and longer lifespan
[120, 121]. Researchers are exploring effective ways to combine the advantages of the
two membranes in order to produce a composite material called as the Mixed Matrix
Membrane (MMM) [122].

Many researchers have worked on the electrochemical oxidation of phenolic


compounds [123, 124]. In this process, the electrode material should be electrochemically
stable, economically feasible and very effective for the removal of the organic pollutants
[125]. There are various studies performed using several anodic materials such as
Ti=SnO2, Pt [126], vitreous carbon [127] and PbO 2 [128]. Among them, PbO2 electrodes
have gained popularity due to their high electrical conductivity, strong oxidizing ability,
low cost, and ability to be used as viable anodic materials for the degradation of organic
contaminants [129, 130].

Photocatalysis can also be used to remove phenol from wastewaters. For


instance, a model employing a Parabolic Trough Collector (PTC) can be utilized in order
to carry out the studied photocatalytic reactions. The solar photo-catalytic system can be
tested in terms of the %phenol removal (%RPhenol) and the %Total Organic Carbon
(TOC) mineralization efficiency (%MTOC) by using a catalyst such as FeSO 4 and a solar
photo such as oxalic acid to enhance the process efficiency. Such a study was performed
by Abid et al. (2019) where the researchers showed that the effectiveness of FeSO 4 and
H2O2 in phenol removal was strongly dependent on the pH of the solution and the
efficiency was higher in acidic environments than in basic environments. The optimum
experimental value of pH for maximum phenol removal was found to be 4. A quadratic-
second order polynomial equation was seen to approximate the photo-catalytic system
well with the correlation coefficient values 95.26% and 95.2% for the parameters of
%RPhenol and %MTOC, respectively [131].

Biological treatment is the most commonly applied treatment for aqueous


phenols. The treatment is inexpensive, simple to design and maintains and transforms the
phenolic solutions into simple end products. Compounds such as Bisphenol A (BPA) can
also be efficiently removed with biological treatments such as activated sludge [132–
135].

Due to the significant challenges associated with conventional phenol


removal methods, research studies have focused on finding alternative methods which
can be applicable to a wide range of reaction conditions with little or no environmental
impacts. Enzymatic treatment is such a method and is seen as a potential alternative to
conventional methods. Enzymes can act on specific recalcitrant pollutants to remove
them by precipitation or transform them to other products. They can also change the
characteristics of a given waste to render it more amenable to treatment or aid in
converting waste material to value-added products [136]. The enzymatic method is
suitable for treating high concentration wastewaters because the process can be
completed without dilution within a short time [137].

Several methods are also available for the removal of chlorinated phenols from
water. Among them, biodegradation [14], oxidation by AOPs [138] and adsorption [139,
140] are the most widely used [141].
6.Comparison of Different Methods for the Removal of Phenol and
Chlorophenols

Although various traditional and advanced methods are possible to apply for the
removal of phenol from wastewaters; the two important parameters which define the
suitable method are the initial and final phenol concentrations as seen in Table 1 and
Table 2 [89]:

Further discussions can also be considered. Extensive research continues to be


carried out on phenol removal from water, from conventional methods like adsorption on
activated carbon to new technologies such as enzymatic treatment. Improvements to
overcome low efficiencies and high operational costs of conventional techniques provide
attractive alternatives such as low-cost adsorbents or chemical modification on the
adsorbents to increase the surface area. Enzymatic treatment has proven to be an effective
method with more than 95 % phenol removal using different peroxidases such as soybean
and horseradish. New developments in the use of additives substantially mitigate enzyme
inactivation which is the most important drawback in the application of enzymatic
treatments [97].

As for chlorophenols adsorption, biodegradation and oxidation by advanced


oxidation processes seem more widely used than others [138-141]. Advanced Oxidation
Processes (AOPs) are based on the generation and use of powerful hydroxyl radicals
(OH-) by means of chemical, photochemical or photocatalytic methods [139]. Adsorption
has proven to be one of the most attractive and effective methods for removing
chlorophenols from water, due to its low maintenance cost, high efficiency, simplicity of
operation as well as no or lower generation of toxic substances [141].

References

[1] Tang W, Physicochemical Treatment of Hazardous Wastes. 1 st ed. Florida: Routledge


Taylor & Francis Group, CRC Press; 2016. ISBN: 9781566769273

[2] Taamalli A, Contreras M, Abu-Reidah I, Trabelsi N, Ben Youssef N. Quality of


Phenolic Compounds: Occurrence, Health Benefits, and Applications in Food Industry.
Journal of Food Quality. 2019;2019:1-2.

[3] Laura A, Moreno-Escamilla J, Rodrigo-García J, Alvarez-Parrilla E. Phenolic


compounds. In: Yahia E, Carillo-Lopez A, ed. by. 1st ed. Woodhead Publishing; 2018.
Ceccaroli B, Lohne O. Solar grade silicon feedstock. In: Luque A, Hegedus S, editors.
Handbook of Photovoltaic Science and Engineering. 2nd ed. Chichester: Wiley; 2011. p.
169-217. DOI: 10.1002/978047974704.ch5.

[4] Anku W, Mamo M, Govender P. Phenolic compounds in water: Sources, reactivity,


toxicity and treatment methods. In: Soto-Hernández M, Palma-Tenango M, García-
Mateos R, ed. by. Phenolic Compounds. London: InTechOpen; 2017. DOI:
10.5772/67213.

[5] Davı̀ M, Gnudi F. Phenolic compounds in surface water. Water Research.


1999;33(14):3213-3219.
[6] Olaniran A, Igbinosa E. Chlorophenols and other related derivatives of environmental
concern: Properties, distribution and microbial degradation processes. Chemosphere.
2011;83(10):1297-1306.

[7] Runge F. Ueber einige Produkte der Steinkohlendestillation. Annalen der Physik und
Chemie. 1834;107(5):65-78.

[8] Runge F. Ueber einige Producte der Steinkohlendestillation. Annalen der Physik und
Chemie. 1834;108(20-25):328-333.

[9] Reichenbach. Meinung von dem Kyanol, der Karbolsäure u. s. f. des Hrn. Runge.
Annalen der Physik und Chemie. 1834;107(32):497-512.

[10] Nguyen, M.T., Kryachko, E.S., Vanquickenborne, L.G. General and theoretical
aspects of phenols. In: Rappoport Z et al., editor. The Chemistry of Phenols. Chichester:
Wiley; 2013. p.6. DOI: 10.1002/0470857277.

[11] Reddy N, Elias A. Chlorine and the Chemistry of Disinfectants. Resonance.


2021;26(3):341-366.

[12] Kenyon R, Boehmr N. Phenol by sulfonation. Industrial and Engineering Chemistry.


1950;42(8):1446-1455.

[13] Rahman M, Bin Abdul Razzak P, Binti Surani N, Binti Musdafa Kamal N, Faudzi S,
Mohammad N et al. Mini Design Project (Production of Phenol). Universiti Teknologi
Mara; 2018.

[14] Shinohara Y. Development in the Manufacturing Process of phenol. Journal of


Synthetic Organic Chemistry, Japan. 1977;35(2):138-146.

[15] Mitsui Petrochemical Industries. Preparation of phenol through cumene.


EP0371738A2, 1989.

[16] Daowdat B, Hoeltzel G, Tannenbaum R. Direct route to phenol from benzene


University of Pennsylvania; 2017. Available from:
https://repository.upenn.edu/cgi/viewcontent.cgi?article=1089&context=cbe_sdr.

[17] Kandola B, Horrocks A. Composites. In: Horrocks A, Price D, editors. Fire


Retardant Materials. 2001.

[18] Yang G, Rohde B, Tesefay H, Robertson M. Biorenewable Epoxy Resins Derived


from Plant-Based Phenolic Acids. ACS Sustainable Chemistry & Engineering.
2016;4(12):6524-6533.

[19] Cygan M, Szemień M, Krompiec S. Statistical screening analysis of the chemical


composition and kinetic study of phenol-formaldehyde resins synthesized in the presence
of polyamines as co-catalysts. 2018;13(5):e0195069.
[20] Kim W, Joshi U, Lee J. Making Polycarbonates without Employing Phosgene: An
Overview on Catalytic Chemistry of Intermediate and Precursor Syntheses for
Polycarbonate. Industrial & Engineering Chemistry Research. 2004;43(9):1897-1914.

[21] Wang B, Wang L, Zhu J, Chen S, Sun H. Condensation of phenol and acetone on a
modified macroreticular ion exchange resin catalyst. Frontiers of Chemical Science and
Engineering. 2013;7(2):218-225.

[22] Vedamurthy T, Murugesan M. Synthesis, characterization, and evaluation of the


hydrophobic, dielectric properties of phenols functionalized nylon 6 polymers by zinc
acetate catalyst using Mannich reaction. Materials Chemistry and Physics. 2018;216:517-
525.

[23] Im K, Jeon J. Synthesis of Plant Phenol-derived Polymeric Dyes for Direct or


Mordant-based Hair Dyeing. Journal of Visualized Experiments. 2016;(118).

[24] Ajani O, Akinremi O, Ajani A, Edobor-Osoh A, Anake W. Synthesis and


Spectroscopic Study of Naphtholic and Phenolic Azo Dyes. Physical Review & Research
International. 2013;3(1):28-41.

[25] DeBono R, Laitung G. Phenolic household disinfectants — further precautions


required. Burns. 1997;23(2):182-185.

[26] Argese E, Bettiol C, Marchetto D, De Vettori S, Zambon A, Miana P et al. Study on


the toxicity of phenolic and phenoxy herbicides using the submitochondrial particle
assay. Toxicology in Vitro. 2005;19(8):1035-1043.

[27] Michałowicz J, Duda W. Phenols – Sources and Toxicity. Polish Journal of


Environmental Studies. 2007;16(3):347-362.

[28] Rättö M, Ritschkoff A, Viikari L. Enzymatically polymerized phenolic compounds


as wood preservatives. Holzforschung. 2004;58(4):440-445.

[29] Zabka M, Pavela R. Antifungal efficacy of some natural phenolic compounds


against significant pathogenic and toxinogenic filamentous fungi. Chemosphere.
2013;93(6):1051-1056.

[30] Badanthadka M, Mehendale H. Chlorophenols. In: Wexler et al. P, ed. by.


Encyclopedia of Toxicology Volume 1. Amsterdam: Elsevier; 2014.

[31] Igbinosa E, Odjadjare E, Chigor V, Igbinosa I, Emoghene A, Ekhaise F et al.


Toxicological Profile of Chlorophenols and Their Derivatives in the Environment: The
Public Health Perspective. The Scientific World Journal. 2013;2013:1-11.

[32] Ullmann’s Encyclopedia of Industrial Chemistry. 6th ed. New York: Wiley; 2001.

[33] Eikhman R, Shemyakin M, Vozhdaeva V. Anilinokrasochnaya. 1934;4:523.


[34] Tishchenko V, Churbakov A. Journal of applied chemistry of the USSR. 1934;7:764.

[35] Vorozhtzov Jr., N.N. and Oshuev, A.G. 1933. Anilinokrasochnaya Promyshl. 3, 245.

[36] Global Phenol Market Report and Forecast 2021-2026 [Internet].


Expertmarketresearch.com. 2021 [cited 7 September 2021]. Available from:
https://www.expertmarketresearch.com/reports/phenol-market-report

[37] Chaudhry G, Chapalamadugu S. Biodegradation of halogenated organic compounds.


Microbiological Reviews. 1991;55(1):59-79.

[38] Wade L. phenol | Definition, Structure, Uses, & Facts [Internet]. Encyclopedia
Britannica. 2021 [cited 7 September 2021]. Available from:
https://www.britannica.com/science/phenol

[39] [Internet]. Shell.com. 2021 [cited 7 September 2021]. Available from:


https://www.shell.com/business-customers/chemicals/our-
products/phenol/_jcr_content/par/textimage.stream/1447712147800/e2796436311990d66
7d25cd639351276d2ea574d/datasheet-phenol.pdf

[40] Yao L, Jiang Y, Shi J, Tomas-Barberan F, Datta N, Singanusong R et al. Flavonoids


in Food and Their Health Benefits. Plant Foods for Human Nutrition. 2004;59(3):113-
122.

[41] Czaplicka M. Sources and transformations of chlorophenols in the natural


environment. Science of The Total Environment. 2004;322(1-3):21-39.

[42] Mohammadi S, Kargari A, Sanaeepur H, Abbassian K, Najafi A, Mofarrah E.


Phenol removal from industrial wastewaters: a short review. Desalination and Water
Treatment. 2014;53(8):2215-2234.

[43] Li D, Park J, Oh J. Silyl Derivatization of Alkylphenols, Chlorophenols, and


Bisphenol A for Simultaneous GC/MS Determination. Analytical Chemistry.
2001;73(13):3089-3095.

[44] Santana C, Padrón M, Ferrera Z, Rodríguez J. Development of a solid-phase


microextraction method with micellar desorption for the determination of chlorophenols
in water samples. Journal of Chromatography A. 2007;1140(1-2):13-20.

[45] de Morais P, Stoichev T, Basto M, Vasconcelos M. Extraction and preconcentration


techniques for chromatographic determination of chlorophenols in environmental and
food samples. Talanta. 2012;89:1-11.

[46] Jin M, Yang Y. Simultaneous determination of nine trace mono- and di-
chlorophenols in water by ion chromatography atmospheric pressure chemical ionization
mass spectrometry. Analytica Chimica Acta. 2006;566(2):193-199.
[47] Kulkarni S, Kaware D. Review on Research for Removal of Phenol from
Wastewater. International Journal of Scientific and Research Publications. 2013;3(4).

[48] Pérez Ramírez, E.E., de la Luz Asunción, M., Saucedo Rivalcoba, V., Martínez
Hernández, A., Velasco Santos, C. 2021. In: Phenolic Compounds Natural Sources,
Importance and Applications. M. Soto-Hernández, M. Palma-Tenango and R. García-
Mateos (Eds.) InTechOpen, London. 344-372.

96. Villegas L, Mashhadi N, Chen M, Mukherjee D, Taylor K, Biswas N. A Short


Review of Techniques for Phenol Removal from Wastewater. Current Pollution Reports.
2016;2(3):157-167.

97.Busca, G., Berardinelli, S., Resini C., Arrighi, L. 2008. Technologies for the removal
of phenol from fluid streams: a short review of recent developments. J. Hazard. Mater.
160. 265–288.

98.Chemat, F., Boutekedjiret, C. 2015. Extraction//Steam distillation. In: Reference


Module in Chemistry, Molecular Sciences and Chemical Engineering. Elsevier,
Amsterdam. 1-12.

99.Park, H.S., Koduru, J.R., Choo, K.H., Lee, B. 2015. Activated carbons impregnated
with iron oxide nanoparticles for enhanced removal of bisphenol A and natural organic
matter. J. Hazard. Mater. 286, 15–324.

100.Tran, V.S., Ngo, H.H., Guo, W., Zhang, J., Liang, S., Ton‐That C., Zhang, X. 2015.
Typical low cost biosorbents for adsorptive removal of specific organic pollutants from
water. Bioresour. Tech. 182, 353–363.

101.de la Luz‐Asunción, M., Sánchez‐Mendieta, V., Martínez‐Hernández, A.L., Castaño,


V.M., Velasco‐Santos, C. 2015. J. Nanomater. 1–15.

102.Carmona, M., Lucas A., Valverde J., Velasco B., Rodriguez, J. 2006. Chem. Eng. J.
117, 155-160.

103.Zhou, L., Cao, H., Descorme, C., Xie, Y. 2018. Front. Environ. Sci. Eng. 12, 1-20.

104.Ribeiro A R, Nunes O C, Pereira M F, Silva A M. 2015. Environ. Int. 75, 33–51.

105.Andreozzi R, Caprio V, Insola A, Marotta R. 1999. Catal. Today 53, 51–59.

106.Víctor-Ortega, M.D., Ochando-Pulido, J.M. 2016. Chem. Eng. Trans. 47, 253-258.

107.Debellefontaine, H., Chakchouk, M., Foussard, J.N., Tissot. D., Striolo, P. 1996.
Environ. Pollut. 1996, 92, 155–164.
108.Dietrich, M., Rall.T.L., Canney,P.J. 1985. Wet air oxidation of hazardous organics in
wastewater. Environ. Prog. 4, 171–177.

109.Fortuny, A., Ferrer, C., Bengoa, C., Font, J., Fabregat, A. 1995. Catal. Today, 24,
79–83.

110.Maugans, C. B., Akgerman, A. 2003. Catalytic wet oxidation of phenol in a trickle


bed reactor over a Pt/TiO2 catalyst. Water Research 37, 319–328.

111.Suárez-Ojeda, M. Fabregat A., Stüber F., Fortuny A., Carrera J., Font J. 2007. Chem.
Eng. J. 132, 105-111.

112.Monteros A., Lafaye G., Cervantes A., Del Angel G., Barbier Jr. J., Torres, G. 2015.
Catal. Today 258 564-569.

113.Baloyi J., Ntho, T., Moma, J. 2018. Vol. 8, 5197-5211.

114.Serra-Pérez, E., Álvarez-Torrellas, S., Ismael Águeda, V., Delgado, J. A., Ovejero,
G., & García, J. (2019). Applied Surface Science, 473, 726–737.

115.TlaMakatsa, T.J., Baloyi, J., Ntho, T., Masuku, C.M. 2021. Catalytic wet air
oxidation of phenol: Review of the reaction mechanism, kinetics, and CFD modeling,
Crit. Rev. Environ. Sci. Tech. 51, 1891-1923.

116.Eftaxias, A. 2002. Catalytic wet air oxidation of phenol in a trickle bed reactor:
Kinetics and reactor modelling. Dissertation. Department of Chemical Engineering,
Higher Technical School of Chemical Engineering, Rovira i Virgili University, Spain.

117.Saputera W., Putrie A., Esmailpour A., Sasongko D., Suendo V., Mukti R., 2021.
Catalysts 11, 998.

118.Wang, K., Abdala, A., Hilal, N., Khraisheh, M. 2017. Membrane Characterization.
259–306.

119.Sioutopoulos, D.C., Karabelas, A.J. 2012. J. Membr. Sci. 407, 34–46.

120.Dordick, J.S., Marletta, M.A., Klibanov, A.M. 1987. Biotech. Bioeng. 30, 31-36.

121.Kumar, R.V., Basumatary, A.K., Ghoshal, A.K., Pugazhenthi, G. 2014. RSC Adv. 5,
6246–6254.

122.Qadir, D., Mukhtar, H., Keong, L.K. 2016. Sep. Purif. Rev. 46, 62–80.

123.Comninellis, C. H., and Pulgarin, C. 1993. J. Appl.Electrochem., 23(2), 108–112.


124.Simond, O., Schaller, V., Comninellis, C.H. 1997. Elctrochim. Acta 42, 2009–2012.

125.L. Gu, B. Wang, H. Z. Ma and W. P. Kong. 2006. J. Chem. Technol. Biotechnol. 81,
1697-1704.

126. Bashir M.J.K., Isa M.H., Kutty S.R.M., Awang Z.B., Azizs H.A., Mohajeri S.,
Farooqi, I.H. 2009. Waste Manage. 29, 2534-2541.

127.Gomes, L., Freitas, R.G., Malpass, G.R.P., Pereira, E.C., Motheo, A. 2009. J.Appl.
Electrochem. 39, 117-121.

128.Gaudet, J., Tavares, A.C., Trasatti, S., Guay., D. 2005. Chem. Mater. 17, 1570-1579.

129.Zhao, T., Lu, J., Hu, C., Zhu, C., Zhao, J., and Dong, W. 2014. Int. J. Electrochem.
Sci. 9, 2354–2366.

130.Stucki, S., Kötz, R., Carcer, B., and Suter, W. 1991. J. Appl. Electrochem. 21, 99–
104.

131.Polcaro, A. M., Vacca, A., Mascia, M., and Palmas, S. 2005. Electrochim. Acta, 50,
1841–1847.

132.Guido, B., Berardinelli, S., Resini, C., and Arrighi, L. 2008. J. Hazard. Mater. 160,
265–288.

133. Saratale, R., Hwang, K., Song, J., Saratale, G., Kim, D. (2016). J. Environ. Eng.
142, 04015064.

134.Ramírez, E., Asunción, M.D., Rivalcoba, V.S., Hernández, A., & Santos, C.V. 2017.
Removal of phenolic compounds from water by adsorption and photocatalysis. In:
Phenolic Compounds - Natural Sources, Importance and Applications.

135.Makino, N, Mason, H.S. 1973. J. Biol. Chem. 24, 5731-5735.

136.Karam, J. and Nicell, J.A. 1997. J. Chem. Technol. Biotechnol. 2, 141-153.

137.Chiong, T., Lau, S.Y., Khor, E.H., Danquah, M.K. 2014. OA Biotechnology 10, 9-
15.

138.Pera-Titus, M., Garcia-Molina, V., Baños, M.A., Giménez, J., Esplugas, S. 2004.
Appl. Catal. B Environ. 47, 219–256.

139.Dąbrowski, A., Podkościelny, P., Hubicki, Z., Barczak, M. 2005. Chemosphere 58,
1049–1070.
140.Soto, M.L., Moure, A, Dominguez, H., Parajo, J.C. 2011. J Food Eng. 105, 1–27

141.Kuśmierek K. 2016. React. Kinet. Mech. Catal. 119, 19–34.

You might also like