Surface water pollution by pharmaceuticals and an alternative of removal by low-cost adsorbents: A review

Surface water pollution by pharmaceuticals and an alternative of removal by low-cost adsorbents: A review

Accepted Manuscript Surface water pollution by pharmaceuticals and an alternative of Removal by lowcost adsorbents: a review Heloise Beatriz Quesada,...

886KB Sizes 2 Downloads 144 Views

Accepted Manuscript Surface water pollution by pharmaceuticals and an alternative of Removal by lowcost adsorbents: a review

Heloise Beatriz Quesada, Aline Takaoka Alves Baptista, Luís Fernando Cusioli, Daiana Seibert, Charleston de Oliveira Bezerra, Rosângela Bergamasco PII:

S0045-6535(19)30225-5

DOI:

10.1016/j.chemosphere.2019.02.009

Reference:

CHEM 23121

To appear in:

Chemosphere

Received Date:

12 November 2018

Accepted Date:

03 February 2019

Please cite this article as: Heloise Beatriz Quesada, Aline Takaoka Alves Baptista, Luís Fernando Cusioli, Daiana Seibert, Charleston de Oliveira Bezerra, Rosângela Bergamasco, Surface water pollution by pharmaceuticals and an alternative of Removal by low-cost adsorbents: a review, Chemosphere (2019), doi: 10.1016/j.chemosphere.2019.02.009

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

1

1

SURFACE WATER POLLUTION BY PHARMACEUTICALS AND

2

AN

3

ADSORBENTS: A REVIEW.

ALTERNATIVE

OF

REMOVAL

BY

LOW-COST

4 5

Heloise Beatriz Quesadaa, Aline Takaoka Alves Baptistaa, Luís Fernando Cusiolia,

6

Daiana Seiberta, Charleston de Oliveira Bezerraa, Rosângela Bergamascoa

7 8

a

9

900, Parana, Brazil. Tel: 55-44-3011-4782, e-mail: [email protected],

State University of Maringa, Department of Chemical Engineering, Maringa 87020-

10

[email protected],

[email protected],

11

[email protected] and [email protected].

[email protected],

12 13

Abstract

14

Micropollutants, also called emerging contaminants, consist of an extensive group of

15

synthetic and natural substances, including pharmaceuticals, personal care products,

16

steroid hormones, and agrochemicals. Currently, the monitoring of residual

17

pharmaceuticals in the environment has been highlighted due to the fact that many of

18

these substances are found in wastewater treatment plants effluents and surface waters,

19

in concentrations ranging from ng L-1 to μg L-1. Most of these compounds are

20

discharged into the environment continuously through domestic sewage treatment

21

systems. In the present work, it is presented an overview of water pollution by these

22

pollutants, as well as a review of the recent literature about the use of low-cost

23

adsorbents for the removal of the main pharmaceuticals found in surface water, focusing

24

on municipal and agroindustrial wastes as precursors. It was possible to observe several

25

examples of high adsorption capacities of these compounds with such materials,

ACCEPTED MANUSCRIPT

2

26

however other aspects must be considered in order to evaluate the real applicability in

27

water and wastewater treatment, such as competition, recyclability and production cost.

28 29

Keywords: Pharmaceuticals. Water pollution. Adsorption. Low-cost adsorbents.

30 31

1. INTRODUCTION

32 33

Over the last decades, the occurrence of micropollutants in the aquatic

34

environment has become a matter of worldwide concern. Micropollutants, also called

35

emerging contaminants, consist of a vast amount of substances of anthropic or natural

36

origin, including pharmaceuticals and personal care products, steroid hormones and

37

agrochemicals. These substances are commonly present in water resources at low

38

concentrations and each substance has a form and mechanism of action, which not only

39

complicate their detection and analysis but also their removal in drinking water and

40

wastewater treatment plants (Bolong et al., 2009; Luo et al., 2014). Another issue is the

41

lack of maximum permissible concentrations of these compounds, and consequently, no

42

or very few precautions and monitoring actions are taken to ensure that these

43

compounds, specifically micro-range pollutants, are not disposed in surface waters.

44

(Bolong et al., 2009).

45

Emerging pollutants are consumed in large amounts every day, and the main

46

characteristic of this group of contaminants is that they do not need to be persistent in

47

the environment to cause negative impacts; their removal or transformation is

48

compensated by their continuous introduction (Petrović et al., 2003). An important fact

49

is that due to their presence in water resources and consequently in drinking water, they

50

can act as human and aquatic organisms endocrine disruptors (Bolong et al., 2009).

ACCEPTED MANUSCRIPT

3

51

Also, it should be taken into account that micropollutants are not found in water

52

resources individually, therefore this mixture can cause synergistic effects, making it

53

more difficult their detection, quantification, and removal (Luo et al., 2014).

54

A large group of chemicals included in emerging contaminants which is poorly

55

taken into account is pharmaceuticals and active ingredients of personal care products

56

(PPCPs). These are used extensively throughout the world, quantities which are often at

57

the same level of agrochemicals (Daughton and Ternes, 1999).

58

Pharmaceuticals have caused great concern because, after their consumption,

59

traces or metabolites are excreted and reach the water resources, either directly or after

60

inefficient treatment (Kümmerer, 2001). Even though the concentrations of

61

pharmaceuticals residues in surface waters are low, their presence and persistence

62

threaten aquatic and terrestrial life, and their effects should not be ignored. Still, there is

63

great difficulty in estimating long-term effects (Asghar et al., 2018; Zuccato et al.,

64

2008). Considering this fact, many studies have proposed tertiary treatments that

65

effectively remove pharmaceuticals from effluent and drinking water. Some techniques

66

evaluated are nanofiltration and reverse osmosis (Garcia-Ivars et al., 2017; Kamrani et

67

al., 2018; Licona et al., 2018), photocatalysis (Dalrymple et al., 2007), ozonization (He

68

et al., 2016; Wang and Bai, 2017) and adsorption (Álvarez-Torrellas et al., 2017; Nam

69

et al., 2014a).

70

Adsorption is considered one of the best alternatives for the removal of organic

71

pollutants, due to its low cost, simplicity of design and ease of operation. Also, during

72

this process, hazardous products are not formed, a situation that can be verified in other

73

treatments (Ahmaruzzaman, 2008). Basically, this process is the accumulation of a

74

pollutant on the surface of a solid (Ali et al., 2012). Besides the removal efficiency of

75

pharmaceuticals and other contaminants, there are a great number of materials that can

ACCEPTED MANUSCRIPT

4

76

be used as precursors for the adsorbents. The expressive waste generation from

77

agricultural, food and wood industries is an attractive option for the development of the

78

so-called 'low-cost adsorbents' and has encouraged many studies of micropollutants

79

removal by these materials (Reddy et al., 2010).

80

Considering the subjects raised, this review aims to bring the current situation

81

of pharmaceuticals as emerging contaminants of surface water and to detail the

82

adsorption process as an alternative to remove these compounds from water and

83

wastewater.

84 85

2. PHARMACEUTICALS

86 87

Pharmaceuticals are chemicals used to diagnose, treat, change and prevent

88

diseases. The definition is extended to veterinary compounds and can also be applied to

89

illicit drugs (Daughton, 2003; Daughton and Ternes, 1999). A wide variety of human

90

medicines including antibiotics, synthetic hormones, anti-inflammatories, statins, and

91

cytotoxins are produced and consumed, some of them in thousands of tons per year

92

(Boxall, 2004; Metcalfe et al., 2003).

93 94 95 96

Pharmaceuticals differ from other chemical contaminants due to the following characteristics (Putschew and Jekel, 2007): (a) They may be formed by innumerable complex molecules which vary in molecular weight, structure, functionality, and form;

97

(b) They have the capacity to go by cellular membranes and consequently are

98

relatively persistent, once they are not inactivated before reaching the expected

99

therapeutic effect;

ACCEPTED MANUSCRIPT 100 101

5

(c) They are polar molecules with more than one ionizable group and their degree of ionization, among other characteristics, depends on the pH of the medium;

102

(d) They are lipophilic and some moderately soluble in water;

103

(e) Drugs such as erythromycin, naproxen, and sulfamethoxazole may persist

104

in the environment for more than one year; others, such as clofibric acid can persist for

105

several years and become biologically active due to accumulation;

106

(f) After administration, the molecules are absorbed in the human body,

107

distributed and subjected to metabolic reactions that can modify their chemical

108

structure.

109 110

2.1 PHARMACEUTICALS CONSUMPTION AND FATE

111 112

The growth rate of financial expenditure per person of pharmaceutical

113

compounds has declined, but their consumption has steadily increased. Due to the

114

development of generic drugs and the production cost decrease, the price of

115

pharmaceuticals became more accessible, which explains the first statement. In

116

addition, increasing demand for treatment of chronic diseases or aging disorders,

117

coupled with economic forces, such as increasing the benefits of health insurance,

118

stimulate drug procurement by the population (Berndt, 2002; Germer and Sinar, 2010).

119

There is a significant difficulty obtaining information about the global

120

consumption of pharmaceuticals, since the administration and number of compounds

121

vary locally, due to different lifestyles and ease of access. Data from Canada indicate

122

that mainly medications prescribed include acetaminophen, acetylsalicylic acid,

123

ibuprofen, naproxen, and carbamazepine (Boxall, 2004; Metcalfe et al., 2003). In

124

European Union, about 3 thousand different substances are used as human medicines,

ACCEPTED MANUSCRIPT

6

125

among them analgesics, anti-inflammatories, contraceptives, antibiotics, beta-blockers,

126

lipid regulators and neuroactive compounds. In Germany in 2001, some of the most

127

commonly used anti-inflammatory drugs were acetylsalicylic acid, with 836 tonnes,

128

paracetamol, 622 tonnes, and ibuprofen, with 345 tonnes (Fent et al., 2006).

129

In 2012, consumption per capita of antihypertensive drugs was the highest in

130

Germany, Hungary and the Czech Republic (575, 543 and 442 doses per thousand

131

people per day, respectively). The increase in sales of medication for the treatment of

132

diabetes was explained by the increase in obesity-related numbers (OECD, 2013). The

133

number of pharmaceuticals consumed influences the effluent load and, consequently,

134

the residual discarded in surface water.

135

Although pharmaceuticals are usually designed with a single mechanism of

136

action and target, they can also have innumerable effects on non-target receptors. In

137

addition, non-target organisms may have receptors and therefore unexpected effects

138

may result from unintended exposure. (Daughton, 2003; Daughton and Ternes, 1999).

139

Concerning

140

inflammation, neurotoxic responses, gametogenic activity and energy status on C.

141

fluminea, a freshwater clam, after exposure to environmentally realistic concentrations

142

of caffeine, ibuprofen, and carbamazepine. It was found that this non-target organism

143

showed concentration-dependent responses related to the mechanism of action of these

144

compounds.

non-target

organisms,

Aguirre-Martínez

et

al.

(2018)

evaluated

145

Currently, the monitoring of residual pharmaceuticals in the environment has

146

been highlighted due to the fact that many of these substances are found in WWTPs

147

effluents and surface waters, in concentrations ranging from ng L-1 to μg L-1 (Bila and

148

Dezotti, 2003). For example, in the studies cited in this review, it was found

149

concentrations from 0.02 ng L-1 to 9.82 μg L-1.

ACCEPTED MANUSCRIPT 150 151

7

In general, emerging pollutants are transported through the environment by routes demonstrated in Figure 1.

152

153 154

Figure 1 – Several routes of emerging pollutants in the environment

155 156

More specifically in the case of pharmaceuticals, these compounds are

157

discharged into the environment continuously through domestic sewage treatment

158

systems. Figure 2 schematizes the main routes and fates of the pharmaceuticals in the

159

environment, since their consumption.

160

ACCEPTED MANUSCRIPT

8

161 162

Figure 2 – Routes and fates of pharmaceuticals in the environment

163 164

Pharmaceuticals in their parent form or as metabolites reach aquatic

165

environments through different routes. The main one is by its discard as domestic

166

sewage after its consumption, metabolism in the human body and excretion. That way,

167

they reach WWTPs and may undergo additional transformations and chemical

168

reactions, forming other products, sometimes more toxic and persistent. The literature

169

shows, however, that many of these compounds are not biodegraded in conventional

170

treatment, so they are commonly discharged with treated effluent into rivers, lakes, and

171

estuaries. In addition, veterinary products can get in aquatic systems through manure

172

and subsequent outflow and also through direct application in aquaculture, which makes

173

monitoring a challenging action (Farré et al., 2008; Fent et al., 2006; Rivera-Utrilla et

174

al., 2013).

175

In surface and groundwater, the metabolic compounds can be converted back to

176

their parental form by the action of microorganisms. Further, the WWTP effluent is

ACCEPTED MANUSCRIPT

9

177

formed by a complex mixture, which may have synergistic effects. Some of these

178

synergic compounds are more bioactive than their precursor (Daughton and Ternes,

179

1999; Fent et al., 2006; Rivera-Utrilla et al., 2013).

180

In water resources, biodegradation and photodegradation are two processes of

181

natural attenuation of the pharmaceuticals, which varies with the complexity of the

182

compound. Although these processes are essential, their results are unsatisfactory. In

183

addition, constant discarding of the compounds may lead to toxicological effects, which

184

are still uncertain due to unpredictable synergistic effects (Li, 2014).

185

The environmental concern related to the presence of pharmaceuticals in surface

186

and groundwater is related not only to quantity but also to their persistence and

187

detriments on aquatic life, such as toxicity and the potential effect on endocrine

188

functions. As an example, the steroid hormone 17α-ethinylestradiol, used in

189

contraceptive pills, has a production of approximately 200 kg per year in the European

190

Union and a per capita consumption of 0.84-2.59 µg cap -1 d-1, low values compared to

191

other compounds. However, it is extremely potent and persistent at low concentrations

192

and has an effect on the reproduction of fishes in the concentration of 1-4ng L-1 or

193

smaller (Fent et al., 2006; Johnson et al., 2013). One study concluded that this drug, as

194

well as diclofenac, can induce structural rupture in the kidney and intestine of fish and

195

also change the genes linked to metabolism control (Lyssimachou and Arukwe, 2007;

196

Mehinto et al., 2010).

197

It has been a great challenge in assessing the toxicological effects of

198

pharmaceuticals in surface and groundwater due to their complexity and the occurrence

199

of long-term effects. A possible parameter of toxicological risk analysis that is widely

200

used in current papers is the Hazard Quotient (HQ), expressed as the ratio between the

201

measured environmental concentration and the predicted non-environmental effect

ACCEPTED MANUSCRIPT

10

202

concentration (PNEC) of pharmaceuticals residues. Rivera-Jaimes et al. (2018)

203

calculated the HQ for three trophic levels (fish, algae, and daphnia) for the most

204

commonly found pharmaceuticals in surface water of the city of Cuernavaca, Mexico.

205

The authors concluded that atenolol, diclofenac, gemfibrozil, ibuprofen and salicylic

206

acid were more toxic to fish than to daphnia and algae. Algae were more sensitive to

207

indomethacin, sulfamethoxazole, and trimethoprim, while daphnia was mainly affected

208

by acetaminophen, gemfibrozil, carbamazepine, and naproxen. Furthermore, ibuprofen,

209

sulfamethoxazole, diclofenac, and naproxen had the highest ecotoxicological risks in

210

surface water, with HQ varying from 14.8 to 111. Although widely used, this quotient

211

considers the effect of a single compound, not the effects of the mixture of compounds

212

present in the aquatic environment. Mutiyar et al. (2018), obtained values of HQ related

213

to pharmaceuticals residues found in the Yamuna River (India) lower than 1, i.e., the

214

toxicological risks were insignificant. However, the authors added that because of the

215

above consideration, the actual risk may even be higher than the worst scenario

216

considered in their study. They further stated that individual low concentrations of

217

pharmaceuticals are unlikely to show acute toxic effects, but the chronic effects cannot

218

be neglected. In addition to the ecological effects of pharmaceuticals, it is important to

219

emphasize a possible consequence of the presence of antibiotics in surface water. Their

220

constant presence and contact with aquatic microbiota can lead to the development of

221

resistant bacteria and genes and consequently accelerate the development of resistance

222

to antibiotics in pathogenic bacteria, impairing the treatment of human infections

223

(Sapkota et al., 2008; Taylor et al., 2011).

224 225 226

2.2 PHARMACEUTICALS OCCURRENCE IN SURFACE WATER

ACCEPTED MANUSCRIPT

11

227

The presence of pharmaceuticals in surface water bodies is a global problem,

228

and its concentration varies with location and population consumption. Table 1

229

summarizes some of the most detected compounds in countries of different continents.

230 231

Table 1 – Occurrence of pharmaceutical in surface water around the world Min Compound

Max

Mean

(ng L-1) (ng L-1) (ng L-1)

Country

Reference

Analgesics and Anti-inflammatories

Acetaminophen

pKa = 9.38

0

9.61

3.18

China

(Yang et al., 2018)

0

200.00

87.30

France

(Celle-Jeanton et al., 2014)

0

1561.00

445.00

India

(Mutiyar et al., 2018)

354.00

508.00

430.00

Mexico

(Rivera-Jaimes et al., 2018)

-

-

89.60

Poland

(Caban et al., 2015)

0

527.00

-

Portugal

(Paíga et al., 2016)

0

69.15

38.18

Portugal

(Pereira et al., 2017)

South-Africa

(Matongo et al., 2015)

-

Diclofenac

1740.00 1296.00

0.03

0.65

0.20

Sweden

(Lindim et al., 2016)

0

9822.00

209.00

UK

(Burns et al., 2018)

0

15.49

3.95

Malaysia

(Praveena et al., 2018)

258.00

352.00

313.00

Mexico

(Rivera-Jaimes et al., 2018)

-

-

40.00

Poland

(Caban et al., 2015)

0

38.00

-

Portugal

(Paíga et al., 2016)

25.13

51.24

33.56

Portugal

(Pereira et al., 2017)

3462.00

49.00

Spain

(Carmona et al., 2014)

0.02

1.49

-

Sweden

(Lindim et al., 2016)

184.00

248.00

231.00

Mexico

(Rivera-Jaimes et al., 2018)

0

2302.00

662.17

India

(Mutiyar et al., 2018)

-

-

55.40

Poland

(Caban et al., 2015)

0

846.00

317.80

South-Africa

(Matongo et al., 2015)

pKa = 4.15

Ibuprofen

ACCEPTED MANUSCRIPT

12

-

6593.00

830.00

Spain

(Carmona et al., 2014)

0

2.21

-

Sweden

(Lindim et al., 2016)

-

-

21.00

USA

(Ferrer and Thurman, 2012)

0.40

86.90

29.51

Portugal

(Pereira et al., 2017)

0.03

1.27

-

Sweden

(Lindim et al., 2016)

834.00

986.00

911.00

Mexico

(Rivera-Jaimes et al., 2018)

-

-

37.70

Poland

(Caban et al., 2015)

0

260.00

-

Portugal

(Paíga et al., 2016)

-

7189.00

278.00

Spain

(Carmona et al., 2014)

0

0.22

-

Sweden

(Lindim et al., 2016)

-

-

95.00

USA

(Ferrer and Thurman, 2012)

pKa = 4.91 Ketoprofen

pKa = 4.45

Naproxen

pKa = 4.15

Antibiotics Erythromycin

pKa = 8.88 Metronidazole

pKa = 2.38

0

6.46

1.31

Bangladesh

(Hossain et al., 2018)

10.20

183.00

55.02

China

(Yang et al., 2018)

32.89

38.80

35.51

Portugal

(Pereira et al., 2017)

0

240.00

60.00

South-Africa

(Matongo et al., 2015)

0.02

0.70

-

Sweden

(Lindim et al., 2016)

0

263.00

-

UK

(Burns et al., 2018)

-

-

137.00

USA

(Ferrer and Thurman, 2012)

0.05

13.51

2.74

Bangladesh

(Hossain et al., 2018)

0

5.10

0.65

China

Asghar et al. (2018)

ACCEPTED MANUSCRIPT

Sulfamethoxazole

0

7.24

1.39

Bangladesh

(Hossain et al., 2018)

0

2.50

1.22

China

(Asghar et al., 2018)

2.83

20.80

9.79

China

(Yang et al., 2018)

19.26

114.24

56.19

Malaysia

(Praveena et al., 2018)

108.00

502.00

299.00

Mexico

(Rivera-Jaimes et al., 2018)

0

43.00

-

Portugal

(Paíga et al., 2016)

South-Africa

(Matongo et al., 2015)

0 pKa1 = 1.6; pKa2 = 5.7

Trimethoprim

pKa = 7.12

13

5320.00 2172.00

0

0.14

-

Sweden

(Lindim et al., 2016)

0

33.20

-

UK

(Burns et al., 2018)

0

14.73

-

USA

(Bean et al., 2018)

-

-

320.00

USA

(Ferrer and Thurman, 2012)

0

17.20

3.06

Bangladesh

(Hossain et al., 2018)

0

15.70

2.84

China

(Asghar et al., 2018)

0.40

52.10

12.35

China

(Yang et al., 2018)

34.00

74.00

61.00

Mexico

(Rivera-Jaimes et al., 2018)

0

290.00

58.00

South-Africa

(Matongo et al., 2015)

0.02

0.33

0.10

Sweden

(Lindim et al., 2016)

0

76.00

-

UK

(Burns et al., 2018)

0

2.29

-

USA

(Bean et al., 2018)

Antidepressives

Diazepam

-

-

24.30

China

(Wu et al., 2015)

0

305.00

55.50

India

(Mutiyar et al., 2018)

-

12.00

3.00

Spain

(Huerta-Fontela et al., 2011)

-

-

0.40

China

(Wu et al., 2015)

2.01

19.50

-

Portugal

(Paíga et al., 2016)

pKa = 3.4 Fluoxetine

ACCEPTED MANUSCRIPT

pKa = 9.8

14

25.37

25.37

25.37

Portugal

(Pereira et al., 2017)

0

0.01

-

Sweden

(Lindim et al., 2016)

-

-

65.00

USA

(Ferrer and Thurman, 2012)

Antiepileptic

Carbamazepine

pKa = 13.9

0

8.80

1.51

Bangladesh

(Hossain et al., 2018)

0

7.00

1.17

China

(Asghar et al., 2018)

0

3.50

0.44

China

(Yang et al., 2018)

-

-

25.30

China

(Wu et al., 2015)

0

5.80

3.30

France

(Celle-Jeanton et al., 2014)

0

1346.00

412.50

India

(Mutiyar et al., 2018)

8

36.00

19.00

Mexico

(Rivera-Jaimes et al., 2018)

0.02

0.21

-

Sweden

(Lindim et al., 2016)

24.90

214.00

-

Portugal

(Paíga et al., 2016)

0

-

10.90

Portugal

(Pereira et al., 2017)

South-Africa

(Matongo et al., 2015)

130.00

Gabapentin

3240.00 1048.00

-

54.00

13.00

Spain

(Huerta-Fontela et al., 2011)

0.94

9.39

-

USA

(Bean et al., 2018)

-

-

350.00

USA

(Ferrer and Thurman, 2012)

0

195.00

-

UK

(Burns et al., 2018)

0

4.55

-

Sweden

(Lindim et al., 2016)

-

-

54.00

USA

(Ferrer and Thurman, 2012)

pKa1 = 3.68; pK2 = 10.7 Anti-hyperglycemic 0.20

121.40

34.73

China

(Asghar et al., 2018)

0.44

8.40

2.60

Sweden

(Lindim et al., 2016)

45.20

2595.00

677.00

Metformin UK (Burns et al., 2018)

ACCEPTED MANUSCRIPT

15

pKa = 12.4

Beta-blockers

Atenolol

pKa = 9.6

Propanolol

-

900.00

470.00

Spain

(Huerta-Fontela et al., 2011)

4.00

10.00

7.00

Mexico

(Rivera-Jaimes et al., 2018)

-

-

21.70

USA

(Battaglin et al., 2018)

0

13.00

4.90

France

(Celle-Jeanton et al., 2014)

0

100.00

31.12

UK

(Burns et al., 2018)

-

-

166.00

USA

(Ferrer and Thurman, 2012)

0.18

8.47

-

Sweden

(Lindim et al., 2016)

-

270.00

54.00

Spain

(Huerta-Fontela et al., 2011)

0

64.90

11.66

United Kingdom

(Burns et al., 2018)

-

-

53.00

USA

(Ferrer and Thurman, 2012)

pKa = 9.42 Central nervous system (CNS) stimulant

Caffeine

pKa = 14

0

220.00

56.90

China

(Asghar et al., 2018)

18.40

293.00

100.51

China

(Yang et al., 2018)

0

81.00

41.30

France

(Celle-Jeanton et al., 2014)

0

2640.00

977.00

India

(Mutiyar et al., 2018)

16.27

36.00

30.83

Malaysia

(Praveena et al., 2018)

0

332.00

68.60

South-Africa

(Matongo et al., 2015)

-

-

591.77

USA

(Battaglin et al., 2018)

8.05

26.92

-

USA

(Bean et al., 2018)

-

-

220.00

USA

(Ferrer and Thurman, 2012)

Mexico

(Rivera-Jaimes et al., 2018)

Lipid regulators Gemfibrozil

14.00

24.00

20.00

ACCEPTED MANUSCRIPT

16

-

3735.00

77.00

Spain

(Carmona et al., 2014)

-

-

95.00

USA

(Ferrer and Thurman, 2012)

pKa = 4.5

232 233

The most detected drug in surface water considering the investigated studies

234

was carbamazepine. This compound is the antiepileptic most commonly found in

235

surface water (Rivera-Utrilla et al., 2013). Burns et al. (2018) detected this substance in

236

10 of the 12 samples collected in Ouse and Foss rivers in the United Kingdom. The

237

constant occurrence of this compound in surface and groundwater was studied by Clara

238

et al. (2004), which found that it is very persistent and is not subject to biodegradation

239

in conventional WWTP.

240

Besides carbamazepine, Burns et al. (2018) observed a high frequency of traces

241

of gabapentin, metformin, and trimethoprim. The authors found that metformin obtained

242

the second highest concentration (6111 ng L-1) in the effluent from a WWTP, only

243

losing to gabapentin (8541 ng L-1); furthermore, it was the compound with the highest

244

annual mean concentration in surface water of two rivers of the city. Metformin,

245

although poorly evaluated by its presence in surface water (present in only three studies

246

of Table 1), has caused concern. Asghar et al. (2018) considered this compound to be

247

one of the most critical of the 33 drugs identified, due to its high concentration in

248

effluents from Wuhan (China) WWTP, which was higher than 100 ng L-1 and to the

249

high occurrence rate in six rivers of the city. Metformin is an antihyperglycemic used in

250

the treatment of diabetes, and the authors justified their critical presence in the country's

251

high number of diabetics.

252

A great frequency of antibiotics could be observed in surface waters. Asghar et

253

al. (2018) found that this group of pharmaceuticals corresponded to 28% of the total

ACCEPTED MANUSCRIPT

17

254

occurrence. Sulfamethoxazole was the second most identified compound in Table 1 and

255

obtained the highest mean, found by Matongo et al. (2015) in Matsunduzi River in

256

South Africa. This compound obtained a high detection frequency among the antibiotics

257

evaluated in Brahmaputra River (Bangladesh) by Hossain et al. (2018), reaching 70%,

258

and was present in all samples collected in the Apatlaco River (Mexico) by Rivera-

259

Jaimes et al. (2018). Generally, the mechanism of sulfamethoxazole is potentiated with

260

trimethoprim, which explains the significant occurrence of this drug in the literature

261

consulted in this study. The maximum concentration of trimethoprim in Table 1 was

262

found by Matongo et al. (2015) of 290 ng L-1. The authors found an even greater

263

concentration of sediment collected in a reservoir in the city, stating that the residence

264

time allowed sorption of the product.

265

Traces of atenolol in aquatic environments are also discussed in the literature.

266

Lindim et al. (2016) have explained that atenolol is not or almost not metabolized in the

267

human body, so it is excreted in the concentrations similar to those consumed. Of all the

268

evaluated pharmaceuticals, this compound obtained the highest concentration found by

269

the authors. Huerta-Fontela et al. (2011) warned that atenolol, besides present in

270

Llobregat River in Spain (mean of 470 ng L-1), is one of the 5 drugs of the 32 evaluated

271

that persists in the water supply after all stages of a conventional DWTP, which shows a

272

great difficulty of removal of the substance.

273

Considering anti-inflammatories, the importance of acetaminophen (or

274

paracetamol) residues in aquatic environments should be emphasized. In Table 1, this

275

compound received the highest concentration among all cited, present in the study of

276

Burns et al. (2018). As a complement, the authors pointed out that acetaminophen

277

obtained the highest concentrations in the influent of all three WWTPs studied, reaching

278

282 319 ng L-1, a condition that, added to the incomplete removal, explains the presence

ACCEPTED MANUSCRIPT

18

279

of the drug in Ouse River in all 12 months evaluated. Paíga et al. (2016) reported a

280

similar pattern; the greatest mass load of the influent of two WWTPs was related to

281

paracetamol, also causing the greatest effluent load and higher concentrations along Lis

282

River in Portugal.

283

Caffeine, a central nervous system stimulant, is commonly associated with

284

analgesics and anti-inflammatories in medicine formulations. This fact, coupled with its

285

presence in several consumed beverages worldwide, justifies the frequency of traces of

286

caffeine in surface water. Asghar et al. (2018) found that the highest concentration was

287

related to caffeine, among all the other pharmaceuticals. Also, Celle-Jeanton et al.

288

(2014) concluded that caffeine together with acetaminophen obtained the highest

289

concentrations among the compounds evaluated in the Allier River, justifying the fact

290

by that both are part of the most consumed drugs in France, which beats their

291

biodegradability. Silva et al. (2014) observed that some of the highest concentrations

292

were found in urban spaces, a fact that could be justified by leakages from septic tanks.

293

The authors also highlighted that the sample of which the concentration was less than

294

the limit of detection was collected in a totally uninhabited area, consolidating the

295

concept that caffeine can be a marker of domestic wastewater.

296

The results discussed above justify the importance of complementary treatments

297

to the conventional, both in WWTPs and in DWTTs. One of the possibilities would be

298

the use of adsorption techniques, detailed in the next topic.

299 300

3. ADSORPTION PROCESSES

301 302

Adsorption is a phase transfer process that is widely used to remove substances

303

from the fluid phase (gases or liquids) and transfer to the solid phase (adsorbent

304

particle). It can be observed in different environmental compartments as a natural

ACCEPTED MANUSCRIPT

19

305

process. Regarding water or effluents treatments, the interaction occurs between a solid

306

and a pollutant, such as pharmaceuticals molecules. The pollutant is called adsorbate

307

and the solid is adsorbent. This technique has been used for the efficient removal of a

308

wide variety of contaminants (Ali et al., 2012; Worch, 2012).

309

According to Ho et al., (2000) in practice, adsorption is carried out in batch or

310

fixed bed column containing a certain mass of porous adsorbent. Under these

311

conditions, the effects of mass transfer are inevitable. The complete adsorption

312

sequence comprises three steps:

313 314

Step 1 - Film diffusion (external diffusion): transport of the adsorbate present in the

315

solution to the external surface of the adsorbent.

316

Step 2 - Diffusion of pores (intraparticle diffusion): transport of the adsorbate from the

317

surface of the adsorbent into the pores.

318

Step 3 - Surface reaction: fixation of the adsorbate on the surface of the adsorbent pores.

319 320

It is important to note that the third step is very fast, and the total adsorption rate

321

is determined by film and/or intraparticle diffusion. Also, since these two steps act in

322

series, the slower process characterizes adsorption (Worch, 2012).

323

Adsorption has been shown to be an excellent and promising technique due to its

324

numerous advantages, including low cost, accessibility, efficiency and environmental

325

benignity (Ali et al., 2012). Also, compared to the conventional methods of separation,

326

the advantages include chemical and/or biological sludge minimization, potential

327

adsorbent regeneration potential, no requirement for nutrients addition and the

328

possibility of recovering the adsorbed material if it has an economic value added

329

(Reddy et al., 2010).

ACCEPTED MANUSCRIPT

20

330

The possible materials to be used as adsorbents are various, some examples

331

being: activated carbon of mineral, animal or vegetable origin, ion exchange resin,

332

carbon nanotubes, chitosan, fly ashes and organic resins (Zhu et al., 2017). Besides the

333

efficiency, it is important to analyze the development of adsorbents and application cost,

334

as well as regeneration capacity. From this economic perspective, the term 'low-cost

335

adsorbents' was created. A low-cost adsorbent is defined as a material that requires little

336

processing, is abundant in nature or may be a waste material or by-product from an

337

industry activity (Rafatullah et al., 2010). In the case of the last materials, the economy

338

in the acquisition can offset the cost of processing. For that reason, a cost study is

339

important. Examples of low-cost adsorbents are clays, parts of plants, animals or other

340

materials with high carbon content, such as fruit residues, bark, algae, mosses, hair, and

341

keratin (Ahmaruzzaman, 2008; Rafatullah et al., 2010).

342

The selection of precursors should take into account the following factors:

343

material accessibility, hazardousness, carbon and oxygen content, abrasion resistance,

344

thermal stability, pore diameter, and high adsorption and regeneration capacity (Ali et

345

al., 2012). The textural, structural, morphological and chemical characterizations are

346

indicators of these qualities. Very common analyzes are the Point of Zero Charge

347

(PZC), zeta potential, scanning electron microscopy (SEM), nitrogen physisorption,

348

Fourier transform infrared spectroscopy (FTIR) and X-ray diffraction (XRD).

349 350

3.1 ADSORPTION STUDIES

351 352

In order to evaluate the applicability of adsorption to the removal of

353

contaminants such as pharmaceuticals, a complete study should be performed.

354

Generally, it contains several steps, such as the effect of pH and ionic strength, kinetics,

ACCEPTED MANUSCRIPT

21

355

isotherms, thermodynamics, desorption, and regeneration. These steps can be conducted

356

using either batch or column methods (Tran et al., 2017).

357

In general, three of these studies are the main constituents of the adsorption

358

theory, which have interdependence. The equilibrium study is considered to be the basis

359

of all adsorption models and a precondition for kinetics and dynamics. It assumes that

360

the adsorption capacity is a function of the concentration of the solution. However, the

361

kinetic study assumes that both the adsorption capacity and the concentration are

362

functions of the contact time. The dynamics is based on the two studies and assumes

363

that in addition to being a function of time, the concentration and adsorption capacity

364

are a function of space. This is used when the adsorption is conducted in a fixed bed

365

column (Worch, 2012).

366

Within the studies involving adsorption, it is common to evaluate two

367

parameters: adsorption capacity and percentage of removal. The increase in the

368

percentage of removal is usually proportional to the increase of the mass of the

369

adsorbent due to the greater availability of adsorption sites, however, does not

370

positively affect the adsorption capacity. This happens because once the percentage of

371

removal reaches 100%, the available sites are not saturated, causing an erroneous

372

estimation of the adsorption capacity (Fan et al., 2017).

373 374

3.1.1

Adsorption kinetics

375 376

As explained above, the adsorption kinetics is the basis for further studies.

377

According to Ho (2006), several studies on adsorption of pollutants have been

378

developed to have a better understanding of the mechanisms and to obtain the order of

ACCEPTED MANUSCRIPT

22

379

reaction found through kinetics. However, the understanding of the adsorption kinetics

380

is limited by the theoretical complexity of the mechanisms.

381

Simplified, the kinetics explains how fast the reaction occurs and still indicates

382

the factors that affect the rate of reaction (Crini and Badot, 2008). Typically, the

383

adsorption equilibrium is not reached instantaneously, as in the case of porous

384

adsorbents. The mass transfer of the solution into the pores inside the particles has

385

resistances, which determine the time required for the equilibrium (Worch, 2012).

386

The adsorption kinetics can be analyzed by mathematical models. The most used

387

models are pseudo-first-order (Ho and McKay, 1999) and pseudo-second-order (Ho and

388

McKay, 1999; Tran et al., 2017). However, the Elovich (Allen and Scaife, 1966; Crini

389

and Badot, 2008) and Intraparticle Diffusion (Weber and Morris, 1963) models have

390

been widely applied.

391 392

3.1.2

Adsorption equilibrium

393 394

The analysis and design of the adsorption process require a study of the

395

equilibrium, which is the most important in the understanding of the process (Vasanth et

396

al., 2007). The adsorption isotherms are equilibrium equations and they are applied to

397

the adsorption process after sufficient time to reach equilibrium at a constant

398

temperature (Kumar, 2007).

399

Since there are a wide variety of equilibrium isotherms to describe the

400

adsorption process, some mathematical models are commonly used to describe

401

equilibrium interactions in a solid-liquid system, such as Langmuir (Langmuir, 1917),

402

Freundlich (Freundlich, 1906), Sips, Tempkin, and Toth (Günay et al., 2007).

403

ACCEPTED MANUSCRIPT 404

3.1.3

23

Adsorption thermodynamics

405 406

Since motion is an intrinsic property of matter and energy is motion-related, it

407

is possible to understand that transformations, whether physical or chemical, are

408

associated with energetic variations. Thermodynamics is part of the physical sciences

409

that evaluates these variations (Nascimento et al., 2014).

410

Information about the adsorption capacity of a material can be obtained through

411

the thermodynamic parameters, such as Gibbs Free Energy (ΔG °), Enthalpy (ΔH °) and

412

Entropy (ΔS°). These are important thermodynamic parameters for the study of the

413

adsorption mechanisms that can confirm the viability, spontaneity and heat exchange

414

for the adsorption process (Zolgharnein et al., 2011).

415 416 417

4. PHARMACEUTICALS ADSORPTION BY LOW-COST ADSORBENTS – A REVIEW ON RECENT LITERATURE

418 419

Many papers have investigated the potential of low-cost adsorbents for the

420

removal of drugs from aqueous solutions. Table 2 was developed taking into account

421

the variety of precursors, mainly municipal and agroindustrial residues.

422 423 424 425

24

426

Table 2 – Adsorption of pharmaceuticals by low-cost adsorbents Chemical Compound

Precursor

Treatment agent

Remov Kinetics

Isotherm

Thermodynamics

Reference

(mg g-1) al (%)

Bamboo

TH

-

-

58

-

-

-

(Solanki and Boyer,

Coconut shell

AC

-

-

94

-

-

-

2017)

-

98

PSO

-

-

AC Acetaminophen

K2CO3, KOH

Cork

pKa = 9.38

qe, max

(Mestre et al., 2014)

AC

-

-

65

PSO

-

-

Cork bark

IN

-

0.99

65

-

-

-

Grape stalk

IN

-

1.74

60

PFO

Langmuir

-

Yohimbe bark

IN

-

31.00

65

-

-

-

(Villaescusa et al., 2011)

ΔG° < 0 Municipal solid

PFO, AC

KOH

180.00

-

wastes

-

ΔH° > 0

(Sumalinog et al., 2018)

PSO ΔS° > 0

Peach stones Atenolol

AC

K2CO3

0.77

80

PSO

Langmuir

-

AC

K2CO3

500.00

-

PSO

Langmuir

-

AC

KOH

555.60

-

PSO

Langmuir

-

AC

-

250.32

-

Elovich

Sips

-

Apple tree branches

Palm kernel shell

(Cabrita et al., 2010) (Marques et al., 2018)

(To et al., 2017)

25

pKa = 9.6

Carbamazepine

Cork

IN

0.37

88

PSO

Freundlich

-

Fish waste

TH

Sewage sludge

(Dordio et al., 2011)

-

27.50

-

-

Sips

-

TH

-

217.40

-

-

Sips

-

Palm kernel shell

AC

-

335.00

-

Elovich

Sips

-

(To et al., 2017)

Peach stones

AC

H3PO4

40.00

-

-

Sips

-

(Torrellas et al., 2015)

(Nielsen et al., 2015)

ΔG° < 0 pKa = 13.9

Rice straw

IN

-

18.00

-

PSO

Freundlich

ΔH° > 0

(Z. Liu et al., 2013)

ΔS° > 0 AC

Grape stalk

pKa = 14

-

98

PSO

-

-

KOH

Cork Caffeine

K2CO3,

(Mestre et al., 2014)

AC

-

-

85

PSO

-

-

IN

-

89.19

75

-

Sips

-

CH

H3PO4

129.56

85

-

Sips

-

AC

H3PO4

395.00

94

-

Sips

-

AC

H3PO4

260.00

-

-

Sips

-

126.00

-

-

Sips

-

Peach stones AC

H3PO4, HNO3

(Portinho et al., 2017)

(Torrellas et al., 2015)

26

ΔG° < 0 Pineapple plant leaves

AC

H3PO4

151.50

-

PSO

Langmuir

ΔH° < 0

(Beltrame et al., 2018)

ΔS° > 0

Diclofenac

Cocoa pod husk

AC

H2SO4

5.53

93

PSO

Freundlich

-

(De Luna et al., 2017)

Coconut shell

AC

-

-

100

-

Freundlich

-

(Nam et al., 2014a)

Olive- waste cakes

AC

H3PO4

38.00

96

PSO

Langmuir

-

(Baccar et al., 2012)

ΔG° > 0 Pine wood

AC

-

0.33

69

PFO

Langmuir

ΔH° > 0 ΔS° < 0 (Lonappan et al., 2018) ΔG° < 0

pKa = 4.15 Pig manure

AC

-

0.95

99

PSO

Freundlich

ΔH° < 0 ΔS° > 0 ΔG° < 0

Ibuprofen

TH

-

5.00

90

PSO

Langmuir

ΔH° < 0 ΔS° > 0

(Chakraborty et al.,

ΔG° < 0

2018)

Aegle marmelos shell

pKa = 4.91

AC

-

12.66

95

PSO

Langmuir

ΔH° < 0 ΔS° > 0

27

AC

150.00

82

PSO

Langmuir

-

KOH

Cork

Cork

K2CO3,

(Mestre et al., 2014)

AC

-

125.00

70

PSO

Langmuir

-

IN

-

0.32

98

PSO

Freundlich

-

(Dordio et al., 2011)

ΔG° < 0 Mung bean husk

AC

-

59.76

99

PSO

Langmuir

ΔH° < 0

(Mondal et al., 2016)

ΔS° < 0 Olive- waste cakes

AC

H3PO4

8.00

Almond shell

AC

H2O2

344.80

79

PSO

Langmuir

-

(Baccar et al., 2012)

PSO

Langmuir

-

(Zbair et al., 2018)

ΔG° < 0 Sulfamethoxazole

Coffee waste

CH

H2SO4

256.95

-

PSO

Langmuir

ΔH° < 0

(Ahsan et al., 2018)

ΔS° < 0 Fish waste

TH

-

17.72

-

Sips

-

(Nielsen and Bandosz,

Sewage sludge

TH

-

79.02

-

Sips

-

2016)

pKa1 = 1.6; pKa2 = 5.7

ΔG° < 0 (Ariful Ahsan et al., Tea leaves

CH

H2SO4

247.29

-

PSO

Temkim

ΔH° < 0 2018) ΔS° < 0

427

IN – in natura, TH – thermal treatment, CH – chemical treatment, AC – activated carbon, PFO – pseudo-first-order, PSO – pseudo-second-order

ACCEPTED MANUSCRIPT

28

428

In Table 2, three classifications of adsorbents were used: in natura (IN),

429

thermically treated (TH) and activated carbon (AC). These treatments can significantly

430

alter the surface properties and pores structure, which is highly related to the uptake of

431

pharmaceuticals molecules. The adsorbents can be used in natura after a simplified

432

washing pretreatment; chemically treated, after contact with chemical solutions for the

433

removal of undesired organic and inorganic matter contained on the surface; and

434

thermally treated, with furnace heating, to increase the surface area breaking the less

435

stable bonds and consequently releasing the volatile fraction of the precursor material

436

(Akhtar et al., 2007).

437

The chemical treatments are used in biomass to improve its adsorption capacity

438

of contaminants and to change the electrical behavior of the surface. The chemical

439

attack is expected to be able to destroy bonds between functional groups and the

440

adsorbent surface. This phenomenon may result in a pore size increase as well as

441

interactions with other functional groups of the contaminant (Módenes et al., 2017). The

442

chemical agents include organic and mineral acids (HCl, HNO3, H2SO4, acetic acid,

443

citric acid, and formic acid), bases and alkali solutions (NaOH, Na2CO3, Ca(OH)2 and

444

CaCl2), oxidizing (H2O2 and K2MnO4), among other compounds (Abdolali et al., 2014).

445

More advanced treatment generates the activated carbon, prepared in a two-

446

stage operation. The first step is the carbonization of the raw material at temperatures

447

below 800ºC in the absence of oxygen. Activation is then carried out with the use of an

448

oxidizing agent (vapor, carbon dioxide or air) at elevated temperatures and sometimes

449

with a catalyst. Activation may be accompanied by chemical or physical treatment

450

(Aygün et al., 2003; Wigmans, 1989). Lately, new methods of activation are being

451

investigated, such as microwave-assisted pyrolysis (Zbair et al., 2018). This method has

452

advantages compared to the conventional two-stage operation, such as more ease of

ACCEPTED MANUSCRIPT

29

453

control, reduced heating time, energy and gas utilization (Yuen and Hameed, 2009).

454

Although the activation process results in the further development of the surface pores,

455

its high cost may become an economic issue, which leads to a greater interest of

456

simplified and efficient treatments of the adsorbents.

457

As expected, it is observed in Table 2 that the higher adsorption capacities are

458

found in activated carbon, followed by chemically or thermally treated adsorbents and,

459

finally, in natura materials. Some works compared different treatments, which allows

460

verifying the influence of each of them on the structure, morphology, and surface loads,

461

characteristics that directly affect the adsorption of drugs. These papers will be

462

highlighted in the discussion of this topic.

463

Portinho et al. (2017), compared three types of adsorbents derived from grape

464

stalk (in natura, chemically treated with phosphoric acid and activated carbon) in the

465

adsorption of caffeine. They observed the above pattern and obtained the maximum qe

466

values of 89.194, 129.56 and 395 mg g-1, respectively. As expected, activated carbon

467

(AC) has a higher surface area and volume of micropores (1009.86 m² g-1 and 0.568 cm³

468

g-1) compared to untreated biomass (6.23 m² g-1 and 0.003 cm³ g-1) and chemically

469

treated (4.21 m² g-1 and 0.002 cm3 g-1). This effect promoted a greater number of sites

470

for adsorption, which implies higher values of qe. Furthermore, the authors emphasized

471

the importance of phosphoric acid addition in the two treatments, which added

472

functional groups to the surface and favored the adsorbate-adsorbent interactions. In this

473

case, the chemical reagent added oxygen groups, which resulted in hydrophilicity to the

474

grape stalk and facilitated the adsorption of polar molecules, such as caffeine.

475

Chakraborty et al. (2018) compared the adsorption capacity of ibuprofen by

476

two adsorbents derived from Aegle marmelos barks: thermally treated and activated

477

with steam. After activation, it was verified that the specific area and pore volume

ACCEPTED MANUSCRIPT

30

478

increased from 4.4 to 308 m² g-1 and from 0.184 to 0.384 cm³ g-1, respectively. This

479

increase not only resulted in the greater removal of ibuprofen but decreased the time

480

required to reach the equilibrium.

481

Regarding the treatments and chemical activations, it is possible to observe the

482

significant occurrence of phosphoric acid (H3PO4) in Table 2. Phosphoric acid is

483

considered to be the most environmentally-friendly reagent when compared to others

484

acids, more corrosive and hazardous, such as nitric acid (HNO3) and sulfuric acid

485

(H2SO4). These also can damage the structure of the adsorbent (Rajapaksha et al.,

486

2016). Benaddi et al. (1998) explained that phosphoric acid promotes the

487

depolymerization of cellulose, dehydration of biopolymers, formation of aromatic rings

488

and elimination of phosphate groups, allowing the increase of the surface area of the

489

adsorbents, besides the addition of functional groups.

490

Potassium hydroxide (KOH) and potassium carbonate (K2CO3) were also

491

frequent reagents in Table 2. Marques et al. (2018) compared the chemical activations

492

by both compounds in apple branches for the adsorption of atenolol. In general, the use

493

of KOH resulted in larger specific areas (maximum of 2472 m² g-1) compared to K2CO3

494

(maximum of 1963 m² g-1). Through scanning electron microscopy, it was observed that

495

the activation with K2CO3 promoted the maintenance of some of the morphological

496

characteristics of the precursor. The authors reported that the activation process with

497

KOH caused more significant destruction of the structure of the AC particles, promoting

498

a more homogeneous particle size distribution of the material. Considering the critical

499

size of the atenolol molecule (0.7 nm), its retention occurs in supermicropores. This fact

500

justified the higher adsorption capacity of adsorbents treated with KOH, which resulted

501

in a higher volume of supermicropores (maximum of 1.19 cm³ g-1).

ACCEPTED MANUSCRIPT

31

502

When it comes to the removal of pharmaceutical compounds from contaminated

503

water, it is important to understand their protonation behavior in aqueous solution. The

504

effect of the pH of the solution on the adsorption processes is important to understand

505

the interaction between the adsorbate and adsorbent since the variation of the pH can

506

promote changes in the superficial charges of adsorbents and influence the protonation

507

of functional groups present in the contaminants (Beltrame et al., 2018). Thus, the pH

508

can affect both the adsorbent surface properties and the nature of the adsorbate, causing

509

different types of mechanisms during the adsorption process (Baccar et al., 2012), and

510

several can be observed in the papers contained in Table 2.

511

The pH of the Point of Zero Charge (pHPZC) of an adsorbent depends on the

512

chemical and electronic properties of the functional groups and is a good indicator for

513

the adsorption process (Song et al., 2010). For a given particle and according to its

514

amphoteric character, it is known that the surface is neutral at pH = pHPZC, negatively

515

charged at pHs higher than pHPZC and positively charged at pHs below pHPZC (Baccar et

516

al., 2012). Regarding pharmaceuticals, these may be weak acids or bases, and their

517

protonation will be based on their pKa values. An acid has the tendency to donate

518

protons, and its dissociation is represented below (Geffertová et al., 2017):

519 HA ⇌H + + A -

520 521 522

Weak acids at pH < pKa are predominantly in their protonated forms (HA),

523

while at pH > pKa they are predominantly in their deprotonated forms (A-). On the

524

contrary, weak bases have the tendency to receive protons, and their ionization occurs

525

as follows (Pardue et al., 2004):

526

ACCEPTED MANUSCRIPT

32

HB + ⇌H + + B

527 528 529

At pH < pKa, weak bases are predominantly in their protonated forms (HB+),

530

while at pH > pKa they are predominantly deprotonated (HB). Briefly, ionizable

531

molecules such as pharmaceuticals may interact with the adsorbent through electrostatic

532

attraction or repulsion, and this interaction varies according to their pKa values (Huerta-

533

Fontela et al., 2011)

534

Exemplifying electrostatic interactions, diclofenac is considered a weak acid

535

since its pKa is around 4.15 (Nam et al., 2014b). According to Nielsen et al. (2015), for

536

hydrophobic compounds, such as diclofenac, adsorption can be largely affected by pH

537

changes, therefore, electrostatic and specific interactions (based on surface polarity,

538

functional groups, organic and inorganic components of the adsorbent) between the

539

contaminant and the surface of the adsorbent particle have an impact on adsorption.

540

Lonappan et al. (2018) verified in their studies that the isoelectric point of pig manure

541

biochar was 2.15 and the adsorption of diclofenac was pH dependent. The maximum

542

removal efficiency of this compound was 99.6% observed at pH 2 (Table 2) and this

543

percentage decreased to 88.8% at pH 12.5. This was justified by the changes in the

544

surface charge and, consequently, the negative surface of the biosorbent repelled by the

545

diclofenac anion.

546

Baccar et al. (2012) in their studies of the biosorbent derived from olive residues

547

(pHPCZ = 5.03) also observed that high pH reduced the uptake of diclofenac and

548

ibuprofen, and this effect was more noticeable when the pH became alkaline. Ibuprofen

549

has pKa = 4.91 and diclofenac pKa = 4.15, both acidic drugs are essentially neutral

550

molecules at pH below the pKa value, however, they acquire a negative charge when

551

the pH is above the pKa value due to the dissociation of molecules of this compound.

ACCEPTED MANUSCRIPT

33

552

Thus, at pH > pHPZC (pH = 8.61), the surface of the adsorbent is negatively charged and

553

a higher proportion of the ionized form of the pharmaceuticals is also charged (pH >

554

pKa), leading to an electrostatic repulsion between adsorbate anions and the surface of

555

the adsorbent and consequently, the removal of pharmaceuticals has decreased.

556

However, although many papers justify the pharmaceuticals adsorption by

557

electrostatic interactions, this is not a rule. Portinho et al. (2017) verified that the best

558

caffeine adsorption capacity was obtained at pH 2. Caffeine is a weak base with pKa =

559

14. This implies that at acid pHs the drug is in its protonated form. When analyzing the

560

adsorbents derived from grape stalks by their surface charges, it was observed that the

561

pHPCZ was around 7.5, which shows that at pH 2 the surface of the particles would be

562

positively charged. Therefore, an electrostatic repulsion would occur, which would

563

adversely affect the adsorption process. The authors then justified that better removals

564

at acid pHs were due to non-electrostatic effects, such as, for example, hydrogen bonds

565

between caffeine and grape stalks. Hydrogen bonds occur preferentially when the

566

surface of the adsorbent is positively charged (Sotelo et al., 2012).

567

In short, Moreno-Castilla (2004) cited three possible mechanisms of adsorption

568

of organic compounds, which can be implied to pharmaceuticals: the π–π dispersion

569

interaction, the hydrogen bonding formation, and the electron donor-acceptor complex

570

formation mechanism. These are related to the chemical properties of the adsorbent

571

surface, that can be verified, for example, through the FTIR spectra analysis. In the case

572

of lignocellulosic materials, some functional groups, such as carboxyl, hydroxyl and

573

amides may be related to the process (Dai et al., 2018).

574

Villaescusa et al. (2011) found likely binding sites for acetaminophen on grape

575

stalks surface through the FTIR analysis, such as hydroxyl groups (peak around 3350

576

cm−1), lignin aromatic bonds (1620 cm-1), guayacil unit (1034, 1159 and 1263 cm-1) and

ACCEPTED MANUSCRIPT

34

577

syringil unit (1105 and 1315 cm-1). After adsorption, there was a shift related to the

578

aromatic rings, indicating the possible adsorption site, and then the authors considered

579

that hydrogen bonding and π-stacking interactions with lignin provided the main

580

mechanism for the adsorption of paracetamol in the grape stalk.

581

Zbair et al. (2018) observed by the FTIR spectra of almond shells activated

582

carbon that some peaks were shifted and the intensity changed when loaded with

583

sulfamethoxazole. The authors considered the respective functional groups as the main

584

adsorption sites. The shifted peaks were from 3397 to 3370 cm-1 (Hydroxyl groups),

585

from 1684 to 1596 cm-1 (C=C vibration of aromatic rings), from 1181 to 1051 cm-1 (C-

586

OH) and 640 to 518 cm-1 (C-H deformation). Taking these changes into account, the

587

possible mechanisms were considered to be due to Van der Waals forces or weak

588

electrostatic interaction.

589

Nielsen et al. (2015) found some indications of the adsorption mechanism by

590

analyzing proton binding curves. This analysis showed that their adsorbents developed

591

from sewage sludge and fish waste exhibited basic nature (pHPZC were between 9 and

592

10). After the adsorption of carbamazepine, there was a decrease in basicity on the

593

sludge carbon. Considering the basic nature of the pharmaceutical molecule, this fact

594

could imply that acid-base interactions were not the main adsorption mechanism. A

595

different situation occurred for the fish-derived carbon, and the increase of basicity was

596

related to the dispersive interactions of the aromatic rings of carbamazepine and the

597

exposed amine groups. Also, it was observed that in the adsorbent treated with lower

598

temperatures, there were more changes in the acidity after the adsorption, suggesting

599

chemical interactions, such as the complexation of the pharmaceutical on metal cations.

600

Concerning this aspect, the authors added that higher temperatures lead to higher levels

ACCEPTED MANUSCRIPT

35

601

of aromatization, attracting the aromatic rings of carbamazepine with stronger

602

dispersive forces.

603

Discussing the kinetic models, it is possible to verify the good fit of the pseudo-

604

second-order model in Table 2. As previously mentioned, together with the pseudo-

605

first-order model, this equation is widely applied to kinetic adsorption data. However,

606

due to restricted conditions and application generally, in the first minutes of reaction,

607

the pseudo-first-order model was applied to a few processes.

608

Lonappan et al. (2018) adjusted the linearized pseudo-first, pseudo-second and

609

Elovich models for the kinetic data of diclofenac adsorption on activated carbon of pine

610

wood. Although the pseudo-second-order model obtained a correlation coefficient (R²)

611

of 0.99, the calculated qe value (400 μg g-1) differed significantly from the experimental

612

one (331 μg g-1), a situation that did not occur with the pseudo-first-order-model (qe, calc

613

= 339 µg g-1). Sumalinog et al. (2018) adjusted the linear forms of the pseudo-first and

614

pseudo-second-order equations in kinetic data of paracetamol adsorption in activated

615

carbons derived from municipal solid waste controlled at three different temperatures.

616

In the three curves, the R² coefficients of the two models were satisfactory, varying

617

from 0.986 to 0.999. In addition, the calculated qe values were close. For these reasons,

618

the authors concluded that both models applied to adsorption.

619

In relation to the equilibrium study, the most prominent model in Table 2 was

620

Langmuir, followed by Sips, Freundlich and, finally, Temkin. The occurrence of the

621

Sips model is justified because this is a simplification of Langmuir at high

622

concentrations of adsorbate. To et al. (2017) explained the good fit of both models to

623

the equilibrium data of atenolol and carbamazepine adsorption in palm kernel shell. The

624

fit was analyzed by the sum of the error squares, and this parameter was 0.000001 for

625

Sips, 0.4 for Langmuir and 262.7 for Freundlich. According to the authors, the good fit

ACCEPTED MANUSCRIPT

36

626

of the Langmuir model followed by Sips completed the process understanding,

627

confirming that the pharmaceuticals were adsorbed in a monolayer on the adsorbent

628

surface. Furthermore, the Sips model, which resulted in nF values of 0.81 and 0.37 for

629

carbamazepine and atenolol, respectively, showed that activated carbon has some

630

heterogeneity, which is not as intense as the good fit of Freundlich model would imply.

631

About the effect of temperature, analyzed by thermodynamic parameters, the

632

works cited in Table 2 did not present a pattern. Some showed that the increase in

633

temperature favored the adsorption capacity of pharmaceuticals (Beltrame et al., 2018;

634

Z. Liu et al., 2013; Lonappan et al., 2018; Sumalinog et al., 2018) and others concluded

635

that this effect disfavored the process (Ahsan et al., 2018; Ariful Ahsan et al., 2018;

636

Chakraborty et al., 2018; Lonappan et al., 2018; Mondal et al., 2016). These situations

637

can be justified either by the Gibbs Free Energy (ΔGº) or the Enthalpy values (ΔHº).

638

The Gibbs Free Energy (ΔGº) is related to spontaneity, fundamental for the

639

viability of the process. Sumalinog et al. (2018) observed that the values of this

640

parameter suffered a decrease of 36% with the temperature increase from 303 to 323K.

641

This situation demonstrated the increase of the spontaneity generated by the temperature

642

increase, also favoring the adsorption capacity. The same was also pointed out by

643

Beltrame et al. (2018) and Liu et al. (2013). In contrast, some authors noticed that the

644

higher values of ΔGº occurred with higher temperatures, i.e., there was a decrease in

645

spontaneity, a fact that accompanied the decrease in the adsorption capacity (Ahsan et

646

al., 2018; Ariful Ahsan et al., 2018; Chakraborty et al., 2018; Mondal et al., 2016).

647

It is noteworthy that only one study obtained a positive value for this parameter,

648

indicating a non-spontaneous adsorption. Lonappan et al. (2018) found positive values

649

for the 5 temperatures analyzed in the study of adsorption of diclofenac in pine wood.

650

The authors justified this fact by the energy barrier that could have impeded the process

ACCEPTED MANUSCRIPT

37

651

at the pH used (6.5), since electrostatic forces played an important role in the

652

adsorption, highly pH dependent.

653

Some studies justified the effect of temperature by the enthalpy values. Positive

654

values of ΔH° indicate endothermic reactions, i.e., they absorb heat from the medium

655

and negative values indicate exothermic reactions, i.e., they release heat. Some authors

656

have therefore related the beneficial effect of temperature increase to positive enthalpy

657

values (H. Liu et al., 2013; Sumalinog et al., 2018) and the beneficial effect of

658

decreasing the temperature to negative values of enthalpy (Ahsan et al., 2018; Ariful

659

Ahsan et al., 2018; Chakraborty et al., 2018; Mondal et al., 2016).

660

Another point to be considered in adsorption of pharmaceuticals is the problem

661

of competition. In real water matrices, these compounds are never alone, and the

662

presence of salts and other compounds should be analyzed because they can enhance or

663

inhibit the adsorption capacity (Fernández et al., 2014; Silva et al., 2018). Therefore,

664

there are papers that studied competitive adsorption, by adding two or more

665

pharmaceuticals in aqueous solution or different types of salts. For example, Fernández

666

et al. (2014) studied the competitive effects of sulfamethoxazole and sulfamethazine and

667

observed a reduction of the adsorption capacity when compared to the values related to

668

each compound separately. In the test using separated solutions of 50 mg L-1, the

669

adsorption capacity of sulfamethoxazole and sulfamethazine was 54.2 and 40.1 mg g-1,

670

respectively. When combined the two solutions, the adsorption capacity was reduced to

671

37.2 and 19.0 mg g-1, respectively. In all tests of adsorption competition, it was verified

672

a higher qe for sulfamethoxazole, showing that the adsorbent had more affinity for that

673

compound.

674

Chakraborty et al. (2018) evaluated the competitive effects on the adsorption of

675

ibuprofen and diclofenac and observed that the removal of the first pharmaceutical was

ACCEPTED MANUSCRIPT

38

676

decreased in the presence of the other. This situation was explained by the difference in

677

the molecular structures of the compounds since the molecule of ibuprofen is smaller

678

than that of diclofenac, and the affinity for this compound was higher due to

679

electrostatic interactions.

680

Sotelo et al. (2012) studied competitive adsorption of diclofenac in two

681

aqueous matrices: ultrapure water and a wastewater treatment plant effluent. The

682

authors explained that in this effluent there was a complex mixture of humic acids,

683

inorganic compounds, carbohydrates, and proteins. All these molecules competed with

684

diclofenac for binding sites on the activated carbon surface, and their availability

685

decreased. The experimental data confirmed this statement since the qe decreased from

686

329 mg g-1 to 184 mg g-1 when used the wastewater treatment plant effluent.

687

Despite all the adsorption studies, to evaluate the application of low-cost

688

adsorbents, regeneration must be considered, due to sustainability and economic

689

concerns. Only a few works cited in Table 2 studied this topic. Ariful Ahsan et al.,

690

(2018) performed successive sulfamethoxazole adsorptions, regenerating their modified

691

adsorbent from tea leaves with ethanol. After three cycles, the adsorption capacity was

692

reduced from approximately 90 to 82 mg g-1, and the authors concluded that the

693

material could be used repeatedly for removal of this compound, being an indicator of

694

applicability. Chakraborty et al. (2018) performed four cycles of ibuprofen adsorption,

695

using methanol. Their low-cost adsorbents were regenerated, achieving more than 74%

696

of desorption after the fourth cycle and the qe remained at least 61% of the initial value.

697

Zbair et al. (2018) evaluated five cycles, using ethanol for desorption. The

698

efficiency of sulfamethoxazole removal of their almond shell carbon decreased only

699

7%, and about 89% of the material was recovered. The authors also performed a

700

nitrogen physisorption analysis, observing that the surface specific area and pore

ACCEPTED MANUSCRIPT

39

701

volume decreased from 1274 to 1134 m2 g-1 and from 1.67 to 1.43 m3 g-1, respectively.

702

This data confirmed the deposition of the pharmaceutical molecules on the adsorbent

703

surface.

704

After the cycles or the exhaustion, the adsorbents must be disposed of. The

705

disposal of the loaded adsorbents is fundamental for their safety assessment application

706

(Simeonidis et al., 2017). The spent activated carbon is often incinerated, but there are

707

some alternatives of other uses. Some adsorbents can be used as a component of

708

building materials, such as cement and bricks (Zhou et al., 2019). Another alternative is

709

to use these as fuel in the boilers/incinerators, depending on its heating value, or for the

710

development of fuel-briquettes (Kushwaha et al., 2010).

711

Lastly, since low-cost adsorbents are being discussed, some aspects of costs

712

should be considered. Again, only a few papers cited performed this kind of analysis.

713

Chakraborty et al. (2018) and Zbair et al. (2018) considered the acquisition, size

714

reduction, washing, drying, carbonization, steam activation, reagents and electricity

715

costs for the analysis. It is noteworthy that, in the case of the two works, there were no

716

costs in the acquisition of the precursor materials, since they were residues (wood apple

717

and almond shell). As mentioned, the economy in the acquisition can offset the cost of

718

processing, and there are simplified and complex treatments. Chakraborty et al. (2018)

719

compared two adsorbents, one after simple carbonization and the other after vapor

720

activation. Obviously, activation increased the cost of processing but the adsorption

721

capacity more than doubled. In this case, it must be balanced the efficiency with

722

economic aspects. The cost of the steam activated adsorbent was estimated at $ 3.6/kg, a

723

low value compared to the activated carbon from Zbair et al. (2018) ($27.8/kg).

724

However, the processing used by the last authors were more complex, because of the

725

reagents (H2O2) and nitrogen atmosphere. Also, the adsorption capacity was superior

ACCEPTED MANUSCRIPT

40

726

(344.8 compared to 12.6 mg g-1). For these reasons, costs vs. efficiency should be very

727

carefully analyzed when it comes to applicability.

728

Analyzing Table 2 in general and the discussion about the recyclability and

729

costs, it can be verified the variety of precursors of adsorbent materials for the removal

730

of pharmaceuticals from contaminated water, as well as the variety of treatments in

731

order to improve the adsorption capacity. Furthermore, it was possible to verify the

732

difference in behavior of the adsorbents and the contaminants, which implies the need

733

for complete studies of the process for the possible application in WWTPs and DWTPs.

734 735

CONCLUSION

736 737

Pharmaceuticals are complex molecules that can be persistent in the

738

environment and resistant to conventional treatments of wastewater and drinking water.

739

It is apparent that pharmaceutical compounds are being found in surface water at low

740

and moderate concentrations. However, even in low concentrations, little is known

741

about its long-term effects, considering its influence on aquatic organisms and human

742

health. Recent researches have indicated the toxicological effect of drugs on the organic

743

functions of aquatic organisms in minimal concentrations and some have already been

744

classified as endocrine disrupters.

745

Given the environmental impacts presented, the importance of a complementary

746

treatment to the conventional one which removes efficiently these compounds becomes

747

clear. Adsorption is an alternative to the reduced cost that has been shown to be efficient

748

in the removal of various organic and inorganic compounds and has been investigated

749

for the removal of several pharmaceuticals. The possibility of precursors of adsorbents

750

is wide, and municipal, agricultural and industrial wastes have gained interest and have

ACCEPTED MANUSCRIPT

41

751

been applied as so-called 'low-cost adsorbents'. In recent articles, it was clear the

752

efficiency of these adsorbents, considering the high adsorption capacities and

753

percentage of removal. However, this does not necessarily mean that their application in

754

WWTPs and DWTPs is feasible.

755

The first problem is the reduced number of studies analyzing the removal of

756

pharmaceuticals in real effluents; most studies use solutions prepared in laboratories

757

with ultrapure water. However, in real matrices, these contaminants are never alone and

758

there is competition between different organic and inorganic compounds, which can

759

significantly decrease the adsorption and removal capacity, as observed in some papers.

760

Another factor to consider is the recyclability. There is some difficulty in finding this

761

topic in the recent literature related to the adsorption of pharmaceuticals, which raises

762

doubts about the resistance and durability of the materials studied. In industry, these

763

features are the key to the interest of application. Moreover, for the real interest in so-

764

called ‘low-cost adsorbents’, is the estimation of production investment, a fundamental

765

point that is little discussed, which consequently interfere in the searching for these

766

materials for this purpose. Therefore, considering these issues, it becomes necessary a

767

careful evaluation of the feasibility of transferring the process from laboratory scale to a

768

pilot scale, which is complex and requires complete studies.

769 770

Acknowledgments

771

The authors thank the National Council for Scientific and Technological

772

Development - CNPq and Higher Education Personnel Improvement Coordination e

773

CAPES for financial support.

774 775

References

ACCEPTED MANUSCRIPT

42

776

Abdolali, A., Guo, W.S., Ngo, H.H., Chen, S.S., Nguyen, N.C., Tung, K.L., 2014.

777

Typical lignocellulosic wastes and by-products for biosorption process in water

778

and wastewater treatment: A critical review. Bioresour. Technol. 160, 57–66.

779

https://doi.org/10.1016/j.biortech.2013.12.037

780

Aguirre-Martínez, G. V., André, C., Gagné, F., Martín-Díaz, L.M., 2018. The effects of

781

human drugs in Corbicula fluminea. Assessment of neurotoxicity, inflammation,

782

gametogenic activity, and energy status. Ecotoxicol. Environ. Saf. 148, 652–663.

783

https://doi.org/10.1016/j.ecoenv.2017.09.042

784

Ahmaruzzaman, M., 2008. Adsorption of phenolic compounds on low-cost adsorbents:

785

A

review.

Adv.

Colloid

786

https://doi.org/10.1016/j.cis.2008.07.002

Interface

Sci.

143,

48–67.

787

Ahsan, M.A., Islam, M.T., Hernandez, C., Kim, H., Lin, Y., Curry, M.L., Gardea-

788

Torresdey, J., Noveron, J.C., 2018. Adsorptive Removal of Sulfamethoxazole and

789

Bisphenol A from Contaminated Water using Functionalized Carbonaceous

790

Material Derived from Tea Leaves. J. Environ. Chem. Eng. 6, 4215–4225.

791

https://doi.org/10.1016/j.jece.2018.06.022

792

Akhtar, M., Moosa Hasany, S., Bhanger, M.I., Iqbal, S., 2007. Sorption potential of

793

Moringa oleifera pods for the removal of organic pollutants from aqueous

794

solutions. J. Hazard. Mater. https://doi.org/10.1016/j.jhazmat.2006.07.016

795

Ali, I., Asim, M., Khan, T.A., 2012. Low cost adsorbents for the removal of organic

796

pollutants

797

https://doi.org/10.1016/j.jenvman.2012.08.028

798 799 800

from

wastewater.

J.

Environ.

Manage.

113,

170–183.

Allen, J.A., Scaife, P.H., 1966. The Elovich Equation and Chemisorption Kinetics. Aust. J. Chem. https://doi.org/10.1071/CH9662015c Álvarez-Torrellas, S., Peres, J.A., Gil-Álvarez, V., Ovejero, G., García, J., 2017.

ACCEPTED MANUSCRIPT

43

801

Effective adsorption of non-biodegradable pharmaceuticals from hospital

802

wastewater with different carbon materials. Chem. Eng. J. 320, 319–329.

803

https://doi.org/10.1016/j.cej.2017.03.077

804

Ariful Ahsan, M., Islam, M.T., Imam, M.A., Hyder, A.G., Jabbari, V., Dominguez, N.,

805

Noveron, J.C., 2018. Biosorption of Bisphenol A and Sulfamethoxazole from

806

Water using Sulfonated Coffee Waste: Isotherm, Kinetic and Thermodynamic

807

Studies. J. Environ. Chem. Eng. https://doi.org/10.1016/j.jece.2018.10.004

808

Asghar, M.A., Zhu, Q., Sun, S., Peng, Y., Shuai, Q., 2018. Suspect screening and target

809

quantification of human pharmaceutical residues in the surface water of Wuhan,

810

China, using UHPLC-Q-Orbitrap HRMS. Sci. Total Environ. 635, 828–837.

811

https://doi.org/10.1016/j.scitotenv.2018.04.179

812

Aygün, A., Yenisoy-Karakaş, S., Duman, I., 2003. Production of granular activated

813

carbon from fruit stones and nutshells and evaluation of their physical, chemical

814

and

815

https://doi.org/10.1016/j.micromeso.2003.08.028

adsorption

properties.

Microporous

Mesoporous

Mater.

816

Baccar, R., Sarrà, M., Bouzid, J., Feki, M., Blánquez, P., 2012. Removal of

817

pharmaceutical compounds by activated carbon prepared from agricultural by-

818

product.

819

https://doi.org/10.1016/j.cej.2012.09.099

Chem.

Eng.

J.

211–212,

310–317.

820

Battaglin, W.A., Bradley, P.M., Iwanowicz, L., Journey, C.A., Walsh, H.L., Blazer,

821

V.S., 2018. Pharmaceuticals, hormones, pesticides, and other bioactive

822

contaminants in water, sediment, and tissue from Rocky Mountain National Park,

823

2012–2013.

824

https://doi.org/10.1016/j.scitotenv.2018.06.150

825

Sci.

Total

Environ.

643,

651–673.

Bean, T.G., Rattner, B.A., Lazarus, R.S., Day, D.D., Burket, S.R., Brooks, B.W.,

ACCEPTED MANUSCRIPT

44

826

Haddad, S.P., Bowerman, W.W., 2018. Pharmaceuticals in water, fish and osprey

827

nestlings in Delaware River and Bay. Environ. Pollut. 232, 533–545.

828

https://doi.org/10.1016/j.envpol.2017.09.083

829

Beltrame, K.K., Cazetta, A.L., de Souza, P.S.C., Spessato, L., Silva, T.L., Almeida,

830

V.C., 2018. Adsorption of caffeine on mesoporous activated carbon fibers prepared

831

from

832

https://doi.org/10.1016/j.ecoenv.2017.08.034

pineapple

plant

leaves.

Ecotoxicol.

Environ.

Saf.

147,

64–71.

833

Benaddi, H., Legras, D., Rouzaud, J.N., Beguin, F., 1998. Influence of the atmosphere

834

in the chemical activation of wood by phosphoric acid. Carbon N. Y.

835

https://doi.org/10.1016/S0008-6223(98)80123-1

836

Berndt, E.R., 2002. Pharmaceuticals in U . S . Health Care : Determinants of Quantity

837

and Price The U . S . pharmaceutical industry has again become the focus of

838

consider ? J. Econ. Perspect. 16, 45–66.

839 840

Bila, D.M., Dezotti, M., 2003. Fármacos no meio ambiente. Quim. Nova 26, 523–530. https://doi.org/10.1590/S0100-40422003000400015

841

Bolong, N., Ismail, A.F., Salim, M.R., Matsuura, T., 2009. A review of the effects of

842

emerging contaminants in wastewater and options for their removal. Desalination

843

239, 229–246. https://doi.org/10.1016/j.desal.2008.03.020

844

Boxall, A.B. a, 2004. The environmental side effects. EMBO Rep. 5, 1110–1116.

845

Burns, E.E., Carter, L.J., Kolpin, D.W., Thomas-Oates, J., Boxall, A.B.A., 2018.

846

Temporal and spatial variation in pharmaceutical concentrations in an urban river

847

system. Water Res. 137, 72–85. https://doi.org/10.1016/j.watres.2018.02.066

848

Caban, M., Lis, E., Kumirska, J., Stepnowski, P., 2015. Determination of

849

pharmaceutical residues in drinking water in Poland using a new SPE-GC-

850

MS(SIM) method based on Speedisk extraction disks and DIMETRIS

ACCEPTED MANUSCRIPT 851

derivatization.

Sci.

Total

Environ.

852

https://doi.org/10.1016/j.scitotenv.2015.08.076

45 538,

402–411.

853

Cabrita, I., Ruiz, B., Mestre, A.S., Fonseca, I.M., Carvalho, A.P., Ania, C.O., 2010.

854

Removal of an analgesic using activated carbons prepared from urban and

855

industrial

856

https://doi.org/10.1016/j.cej.2010.07.058

residues.

Chem.

Eng.

J.

163,

249–255.

857

Carmona, E., Andreu, V., Picó, Y., 2014. Occurrence of acidic pharmaceuticals and

858

personal care products in Turia River Basin: From waste to drinking water. Sci.

859

Total Environ. 484, 53–63. https://doi.org/10.1016/j.scitotenv.2014.02.085

860

Celle-Jeanton, H., Schemberg, D., Mohammed, N., Huneau, F., Bertrand, G., Lavastre,

861

V., Le Coustumer, P., 2014. Evaluation of pharmaceuticals in surface water:

862

Reliability

863

https://doi.org/10.1016/j.envint.2014.06.015

of

PECs

compared

to

MECs.

Environ.

Int.

73,

10–21.

864

Chakraborty, P., Banerjee, S., Kumar, S., Sadhukhan, S., Halder, G., 2018. Elucidation

865

of ibuprofen uptake capability of raw and steam activated biochar of Aegle

866

marmelos shell : Isotherm , kinetics , thermodynamics and cost estimation. Process

867

Saf. Environ. Prot. 118, 10–23. https://doi.org/10.1016/j.psep.2018.06.015

868

Clara, M., Strenn, B., Kreuzinger, N., 2004. Carbamazepine as a possible anthropogenic

869

marker in the aquatic environment: Investigations on the behaviour of

870

Carbamazepine in wastewater treatment and during groundwater infiltration. Water

871

Res. 38, 947–954. https://doi.org/10.1016/j.watres.2003.10.058

872

Crini, G., Badot, P.M., 2008. Application of chitosan, a natural aminopolysaccharide,

873

for dye removal from aqueous solutions by adsorption processes using batch

874

studies: A review of recent literature. Prog. Polym. Sci. 33, 399–447.

875

https://doi.org/10.1016/j.progpolymsci.2007.11.001

ACCEPTED MANUSCRIPT

46

876

Dai, Y., Sun, Q., Wang, W., Lu, L., Liu, M., Li, J., Yang, S., Sun, Y., Zhang, K., Xu, J.,

877

Zheng, W., Hu, Z., Yang, Y., Gao, Y., Chen, Y., Zhang, X., Gao, F., Zhang, Y.,

878

2018. Utilizations of agricultural waste as adsorbent for the removal of

879

contaminants:

880

https://doi.org/10.1016/j.chemosphere.2018.06.179

A

review.

Chemosphere

211,

235–253.

881

Dalrymple, O.K., Yeh, D.H., Trotz, M.A., 2007. Removing pharmaceuticals and

882

endocrine-disrupting compounds from wastewater by photocatalysis. J. Chem.

883

Technol. Biotechnol. https://doi.org/10.1002/jctb.1657

884

Daughton, C.G., 2003. Cradle-to-cradle stewardship of drugs for minimizing their

885

environmental disposition while promoting human health. I. Rational for and

886

avenues toward a green pharmacy. Environ. Health Perspect. 111, 757–774.

887

https://doi.org/10.1289/ehp.5947

888

Daughton, C.G., Ternes, T.A., 1999. Pharmaceuticals and Personal Care Products in the

889

Environment: Agents of Subtle Change? Click here for the ERRATA for Table 2.

890

Environ. Heal. Perspect. @BULLET. https://doi.org/10.2307/3434573

891

De Luna, M.D.G., Murniati, Budianta, W., Rivera, K.K.P., Arazo, R.O., 2017. Removal

892

of sodium diclofenac from aqueous solution by adsorbents derived from cocoa pod

893

husks.

894

https://doi.org/10.1016/j.jece.2017.02.018

J.

Environ.

Chem.

Eng.

5,

1465–1474.

895

Dordio, A. V., Gonçalves, P., Texeira, D., Candeias, A.J., Castanheiro, J.E., Pinto, A.P.,

896

Carvalho, A.J.P., 2011. Pharmaceuticals sorption behaviour in granulated cork for

897

the selection of a support matrix for a constructed wetlands system. Int. J. Environ.

898

Anal. Chem. 91, 615–631. https://doi.org/10.1080/03067319.2010.510605

899

Fan, S., Wang, Y., Wang, Z., Tang, J., Tang, J., Li, X., 2017. Removal of methylene

900

blue from aqueous solution by sewage sludge-derived biochar: Adsorption

ACCEPTED MANUSCRIPT

47

901

kinetics, equilibrium, thermodynamics and mechanism. J. Environ. Chem. Eng.

902

https://doi.org/10.1016/j.jece.2016.12.019

903

Farré, M. la, Pérez, S., Kantiani, L., Barceló, D., 2008. Fate and toxicity of emerging

904

pollutants, their metabolites and transformation products in the aquatic

905

environment.

906

https://doi.org/10.1016/j.trac.2008.09.010

907 908

TrAC

-

Trends

Anal.

Chem.

27,

991–1007.

Fent, K., Weston, A.A., Caminada, D., 2006. Ecotoxicology of human pharmaceuticals. Aquat. Toxicol. 76, 122–159. https://doi.org/10.1016/j.aquatox.2005.09.009

909

Fernández, A.M.L., Rendueles, M., Díaz, M., 2014. Competitive Retention of

910

Sulfamethoxazole (SMX) and Sulfamethazine (SMZ) from Synthetic Solutions in a

911

Strong Anionic Ion Exchange Resin. Solvent Extr. Ion Exch. 32, 763–781.

912

https://doi.org/10.1080/07366299.2014.940235

913

Ferrer, I., Thurman, E.M., 2012. Analysis of 100 pharmaceuticals and their degradates

914

in water samples by liquid chromatography/quadrupole time-of-flight mass

915

spectrometry.

916

https://doi.org/10.1016/j.chroma.2012.03.059

917 918

J.

Chromatogr.

A

1259,

148–157.

Freundlich, H.M.F., 1906. Over the Adsorption in Solution. J. Phys. Chem. https://doi.org/10.1016/S0166-4328(00)00267-9

919

Garcia-Ivars, J., Martella, L., Massella, M., Carbonell-Alcaina, C., Alcaina-Miranda,

920

M.I., Iborra-Clar, M.I., 2017. Nanofiltration as tertiary treatment method for

921

removing trace pharmaceutically active compounds in wastewater from wastewater

922

treatment plants. Water Res. https://doi.org/10.1016/j.watres.2017.08.070

923

Geffertová, D., Ali, S.T., Šolínová, V., Krečmerová, M., Holý, A., Havlas, Z., Kašička,

924

V., 2017. Investigation of the acid-base and electromigration properties of

925

5‐azacytosine derivatives using capillary electrophoresis and density functional

ACCEPTED MANUSCRIPT 926

theory

calculations.

J.

Chromatogr.

927

https://doi.org/10.1016/j.chroma.2016.11.058

A

48 1479,

185–193.

928

Germer, J., Sinar, E., 2010. Pharmaceutical consumption and residuals potentially

929

relevant to nutrient cycling in Greater Accra, Ghana. J. Agric. Rural Dev. Trop.

930

Subtrop. 111, 41–53.

931

Günay, A., Arslankaya, E., Tosun, I., 2007. Lead removal from aqueous solution by

932

natural and pretreated clinoptilolite: Adsorption equilibrium and kinetics. J.

933

Hazard. Mater. 146, 362–371. https://doi.org/10.1016/j.jhazmat.2006.12.034

934

He, Y., Sutton, N.B., Rijnaarts, H.H.H., Langenhoff, A.A.M., 2016. Degradation of

935

pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under

936

simulated

937

https://doi.org/10.1016/j.apcatb.2015.09.015

938 939 940 941

solar

irradiation.

Appl.

Catal.

B

Environ.

Ho, Y.S., 2006. Review of second-order models for adsorption systems. J. Hazard. Mater. https://doi.org/10.1016/j.jhazmat.2005.12.043 Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem. 34, 451–465. https://doi.org/10.1016/S0032-9592(98)00112-5

942

Ho, Y.S., McKay, G., 1998. A Comparison of Chemisorption Kinetic Models Applied

943

to Pollutant Removal on Various Sorbents. Process Saf. Environ. Prot. 76, 332–

944

340. https://doi.org/10.1205/095758298529696

945 946

Ho, Y.S., Ng, J.C.Y., McKay, G., 2000. Kinetics of pollutant sorption by biosorbents: Review. Sep. Purif. Methods. https://doi.org/10.1081/SPM-100100009

947

Hossain, A., Nakamichi, S., Habibullah-Al-Mamun, M., Tani, K., Masunaga, S.,

948

Matsuda, H., 2018. Occurrence and ecological risk of pharmaceuticals in river

949

surface

950

https://doi.org/10.1016/j.envres.2018.04.030

water

of

Bangladesh.

Environ.

Res.

165,

258–266.

ACCEPTED MANUSCRIPT

49

951

Huerta-Fontela, M., Galceran, M.T., Ventura, F., 2011. Occurrence and removal of

952

pharmaceuticals and hormones through drinking water treatment. Water Res. 45,

953

1432–1442. https://doi.org/10.1016/j.watres.2010.10.036

954

Johnson, A.C., Dumont, E., Williams, R.J., Oldenkamp, R., Cisowska, I., Sumpter, J.P.,

955

2013. Do concentrations of ethinylestradiol, estradiol, and diclofenac in European

956

rivers exceed proposed EU environmental quality standards? Environ. Sci.

957

Technol. 47, 12297–12304. https://doi.org/10.1021/es4030035

958

Kamrani, M., Akbari, A., Yunessnia lehi, A., 2018. Chitosan-modified acrylic

959

nanofiltration membrane for efficient removal of pharmaceutical compounds. J.

960

Environ. Chem. Eng. 6, 583–587. https://doi.org/10.1016/j.jece.2017.12.044

961

Kumar, K.V., 2007. Optimum sorption isotherm by linear and non-linear methods for

962

malachite

green

onto

lemon

963

https://doi.org/10.1016/j.dyepig.2006.03.026

peel.

Dye.

Pigment.

964

Kümmerer, K., 2001. Drugs in the environment: Emission of drugs, diagnostic aids and

965

disinfectants into wastewater by hospitals in relation to other sources - A review.

966

Chemosphere 45, 957–969. https://doi.org/10.1016/S0045-6535(01)00144-8

967

Kushwaha, J.P., Srivastava, V.C., Mall, I.D., 2010. Treatment of dairy wastewater by

968

commercial activated carbon and bagasse fly ash: Parametric, kinetic and

969

equilibrium modelling, disposal studies. Bioresour. Technol. 101, 3474–3483.

970

https://doi.org/10.1016/j.biortech.2010.01.002

971 972 973

Langmuir, I., 1917. the Constitution and Fundamental Properties of Solids and Liquids. J. Am. Chem. Soc. https://doi.org/10.1021/ja02254a006 Li, W.C., 2014. Occurrence, sources, and fate of pharmaceuticals in aquatic

974

environment

and

soil.

Environ.

975

https://doi.org/10.1016/j.envpol.2014.01.015

Pollut.

187,

193–201.

ACCEPTED MANUSCRIPT

50

976

Licona, K.P.M., Geaquinto, L.R. d. O., Nicolini, J. V., Figueiredo, N.G., Chiapetta,

977

S.C., Habert, A.C., Yokoyama, L., 2018. Assessing potential of nanofiltration and

978

reverse osmosis for removal of toxic pharmaceuticals from water. J. Water Process

979

Eng. 25, 195–204. https://doi.org/10.1016/j.jwpe.2018.08.002

980

Lindim, C., van Gils, J., Georgieva, D., Mekenyan, O., Cousins, I.T., 2016. Evaluation

981

of human pharmaceutical emissions and concentrations in Swedish river basins.

982

Sci. Total Environ. 572, 508–519. https://doi.org/10.1016/j.scitotenv.2016.08.074

983

Liu, H., Ning, W., Cheng, P., Zhang, J., Wang, Y., Zhang, C., 2013. Evaluation of

984

animal hairs-based activated carbon for sorption of norfloxacin and acetaminophen

985

by comparing with cattail fiber-based activated carbon. J. Anal. Appl. Pyrolysis

986

101, 156–165. https://doi.org/10.1016/j.jaap.2013.01.016

987

Liu, Z., Zhou, X., Chen, X., Dai, C., Zhang, J., Zhang, Y., 2013. Biosorption of

988

clofibric acid and carbamazepine in aqueous solution by agricultural waste rice

989

straw. J. Environ. Sci. (China) 25, 2384–2395. https://doi.org/10.1016/S1001-

990

0742(12)60324-6

991

Lonappan, L., Rouissi, T., Kaur, S., Verma, M., Surampalli, R.Y., 2018. Bioresource

992

Technology An insight into the adsorption of diclofenac on di ff erent biochars :

993

Mechanisms , surface chemistry , and thermodynamics 249, 386–394.

994

https://doi.org/10.1016/j.biortech.2017.10.039

995

Luo, Y., Guo, W., Ngo, H.H., Nghiem, L.D., Hai, F.I., Zhang, J., Liang, S., Wang, X.C.,

996

2014. A review on the occurrence of micropollutants in the aquatic environment

997

and their fate and removal during wastewater treatment. Sci. Total Environ. 473–

998

474, 619–641. https://doi.org/10.1016/j.scitotenv.2013.12.065

999

Lyssimachou, A., Arukwe, A., 2007. Alteration of brain and interrenal StAR protein,

1000

P450scc, and Cyp11β mRNA levels in Atlantic salmon after nominal waterborne

ACCEPTED MANUSCRIPT

51

1001

exposure to the synthetic pharmaceutical estrogen ethynylestradiol. J. Toxicol.

1002

Environ. Heal. - Part A Curr. Issues. https://doi.org/10.1080/10937400600882905

1003

Marques, S.C.R., Mestre, A.S., Machuqueiro, M., Gotvajn, A.Ž., Marinšek, M.,

1004

Carvalho, A.P., 2018. Apple tree branches derived activated carbons for the

1005

removal

1006

https://doi.org/10.1016/j.cej.2018.01.076

of

β-blocker

atenolol.

Chem.

Eng.

J.

345,

669–678.

1007

Matongo, S., Birungi, G., Moodley, B., Ndungu, P., 2015. Pharmaceutical residues in

1008

water and sediment of Msunduzi River, KwaZulu-Natal, South Africa.

1009

Chemosphere 134, 133–140. https://doi.org/10.1016/j.chemosphere.2015.03.093

1010

Mehinto, A.C., Hill, E.M., Tyler, C.R., 2010. Uptake and biological effects of

1011

environmentally relevant concentrations of the nonsteroidal anti-inflammatory

1012

pharmaceutical diclofenac in rainbow trout (oncorhynchus mykiss). Environ. Sci.

1013

Technol. https://doi.org/10.1021/es903702m

1014

Mestre, A.S., Pires, R.A., Aroso, I., Fernandes, E.M., Pinto, M.L., Reis, R.L., Andrade,

1015

M.A., Pires, J., Silva, S.P., Carvalho, A.P., 2014. Activated carbons prepared from

1016

industrial pre-treated cork: Sustainable adsorbents for pharmaceutical compounds

1017

removal. Chem. Eng. J. 253, 408–417. https://doi.org/10.1016/j.cej.2014.05.051

1018

Metcalfe, C.D., Miao, X.S., Koenig, B.G., Struger, J., 2003. Distribution of acidic and

1019

neutral drugs in surface waters near sewage treatment plants in the lower Great

1020

Lakes, Canada. Environ. Toxicol. Chem. https://doi.org/10.1897/02-627

1021

M�denes, A.N., de Oliveira, A.P., Espinoza-Qui�ones, F.R., Trigueros, D.E.G.,

1022

Kroumov, A.D., Borba, C.E., Hinterholz, C.L., Bergamasco, R., 2017. Potential of

1023

Salvinia auriculata biomass as biosorbent of the Cr(III): directed chemical

1024

treatment, modeling and sorption mechanism study. Environ. Technol. (United

1025

Kingdom) 38, 1474–1488. https://doi.org/10.1080/09593330.2016.1234002

ACCEPTED MANUSCRIPT

52

1026

Mondal, S., Bobde, K., Aikat, K., Halder, G., 2016. Biosorptive uptake of ibuprofen by

1027

steam activated biochar derived from mung bean husk: Equilibrium, kinetics,

1028

thermodynamics, modeling and eco-toxicological studies. J. Environ. Manage.

1029

182, 581–594. https://doi.org/10.1016/j.jenvman.2016.08.018

1030

Moreno-Castilla, C., 2004. Adsorption of organic molecules from aqueous solutions on

1031

carbon

materials.

Carbon

N.

1032

https://doi.org/10.1016/j.carbon.2003.09.022

Y.

42,

83–94.

1033

Mutiyar, P.K., Gupta, S.K., Mittal, A.K., 2018. Fate of pharmaceutical active

1034

compounds (PhACs) from River Yamuna, India: An ecotoxicological risk

1035

assessment

1036

https://doi.org/10.1016/j.ecoenv.2017.12.041

approach.

Ecotoxicol.

Environ.

Saf.

150,

297–304.

1037

Nam, S.W., Choi, D.J., Kim, S.K., Her, N., Zoh, K.D., 2014a. Adsorption

1038

characteristics of selected hydrophilic and hydrophobic micropollutants in water

1039

using

1040

https://doi.org/10.1016/j.jhazmat.2014.01.037

activated

carbon.

J.

Hazard.

Mater.

270,

144–152.

1041

Nam, S.W., Jo, B. Il, Yoon, Y., Zoh, K.D., 2014b. Occurrence and removal of selected

1042

micropollutants in a water treatment plant. Chemosphere 95, 156–165.

1043

https://doi.org/10.1016/j.chemosphere.2013.08.055

1044

Nascimento, R.F., Lima, A.C.A., Vidal, C.B., Melo, D.Q., Raulino, G.S.C., 2014.

1045

Adsorção: Aspectos teóricos e aplicações ambientais, Imprensa Universitária da

1046

Universidade Federal do Ceará (UFC). https://doi.org/10.13140/RG.2.1.4340.1041

1047

Nielsen, L., Bandosz, T.J., 2016. Analysis of sulfamethoxazole and trimethoprim

1048

adsorption on sewage sludge and fish waste derived adsorbents. Microporous

1049

Mesoporous Mater. 220, 58–72. https://doi.org/10.1016/j.micromeso.2015.08.025

1050

Nielsen, L., Zhang, P., Bandosz, T.J., 2015. Adsorption of carbamazepine on

ACCEPTED MANUSCRIPT

53

1051

sludge/fish waste derived adsorbents: Effect of surface chemistry and texture.

1052

Chem. Eng. J. 267, 170–181. https://doi.org/10.1016/j.cej.2014.12.113

1053

OECD, 2013. Health at a Glance - OECD Indicators.

1054

Paíga, P., Santos, L.H.M.L.M., Ramos, S., Jorge, S., Silva, J.G., Delerue-Matos, C.,

1055

2016. Presence of pharmaceuticals in the Lis river (Portugal): Sources, fate and

1056

seasonal

1057

https://doi.org/10.1016/j.scitotenv.2016.08.089

variation.

Sci.

Total

Environ.

573,

164–177.

1058

Pardue, H.L., Odeh, I.N., Tesfai, T.M., 2004. Unified Approximations: A New

1059

Approach for Monoprotic Weak Acid-Base Equilibria. J. Chem. Educ. 81, 1367.

1060

https://doi.org/10.1021/ed081p1367

1061

Pereira, A.M.P.T., Silva, L.J.G., Laranjeiro, C.S.M., Meisel, L.M., Lino, C.M., Pena,

1062

A., 2017. Human pharmaceuticals in Portuguese rivers: The impact of water

1063

scarcity in the environmental risk. Sci. Total Environ. 609, 1182–1191.

1064

https://doi.org/10.1016/j.scitotenv.2017.07.200

1065

Petrović, M., Gonzalez, S., Barceló, D., 2003. Analysis and removal of emerging

1066

contaminants in wastewater and drinking water. TrAC - Trends Anal. Chem. 22,

1067

685–696. https://doi.org/10.1016/S0165-9936(03)01105-1

1068

Portinho, R., Zanella, O., Féris, L.A., 2017. Grape stalk application for caffeine removal

1069

through

adsorption.

J.

Environ.

1070

https://doi.org/10.1016/j.jenvman.2017.07.033

Manage.

202,

178–187.

1071

Praveena, S.M., Shaifuddin, S.N.M., Sukiman, S., Nasir, F.A.M., Hanafi, Z.,

1072

Kamarudin, N., Ismail, T.H.T., Aris, A.Z., 2018. Pharmaceuticals residues in

1073

selected tropical surface water bodies from Selangor (Malaysia): Occurrence and

1074

potential

1075

https://doi.org/10.1016/j.scitotenv.2018.06.058

risk

assessments.

Sci.

Total

Environ.

642,

230–240.

ACCEPTED MANUSCRIPT 1076

Putschew, A., Jekel, M., 2007. Analysis, Fate and Removal of Pharmaceuticals in the

1077

Water

1078

https://doi.org/10.1016/S0166-526X(07)50013-9

1079

54

Cycle,

Comprehensive

Analytical

Chemistry.

Rafatullah, M., Sulaiman, O., Hashim, R., Ahmad, A., 2010. Adsorption of methylene

1080

blue

on

low-cost

adsorbents:

A

1081

https://doi.org/10.1016/j.jhazmat.2009.12.047

review.

J.

Hazard.

Mater.

1082

Rajapaksha, A.U., Chen, S.S., Tsang, D.C.W., Zhang, M., Vithanage, M., Mandal, S.,

1083

Gao, B., Bolan, N.S., Ok, Y.S., 2016. Engineered/designer biochar for contaminant

1084

removal/immobilization from soil and water: Potential and implication of biochar

1085

modification.

1086

https://doi.org/10.1016/j.chemosphere.2016.01.043

Chemosphere

148,

276–291.

1087

Reddy, D.H.K., Seshaiah, K., Reddy, A.V.R., Rao, M.M., Wang, M.C., 2010.

1088

Biosorption of Pb2+ from aqueous solutions by Moringa oleifera bark: Equilibrium

1089

and

1090

https://doi.org/10.1016/j.jhazmat.2009.09.128

kinetic

studies.

J.

Hazard.

Mater.

1091

Rivera-Jaimes, J.A., Postigo, C., Melgoza-Alemán, R.M., Aceña, J., Barceló, D., López

1092

de Alda, M., 2018. Study of pharmaceuticals in surface and wastewater from

1093

Cuernavaca, Morelos, Mexico: Occurrence and environmental risk assessment.

1094

Sci.

1095

https://doi.org/10.1016/j.scitotenv.2017.09.134

Total

Environ.

613–614,

1263–1274.

1096

Rivera-Utrilla, J., Sánchez-Polo, M., Ferro-García, M.Á., Prados-Joya, G., Ocampo-

1097

Pérez, R., 2013. Pharmaceuticals as emerging contaminants and their removal from

1098

water.

1099

https://doi.org/10.1016/j.chemosphere.2013.07.059

1100

A

review.

Chemosphere

93,

1268–1287.

Sapkota, A., Sapkota, A.R., Kucharski, M., Burke, J., McKenzie, S., Walker, P.,

ACCEPTED MANUSCRIPT

55

1101

Lawrence, R., 2008. Aquaculture practices and potential human health risks:

1102

Current

1103

https://doi.org/10.1016/j.envint.2008.04.009

knowledge

and

future

priorities.

Environ.

Int.

1104

Silva, C.P., Jaria, G., Otero, M., Esteves, V.I., Calisto, V., 2018. Waste-based

1105

alternative adsorbents for the remediation of pharmaceutical contaminated waters:

1106

Has a step forward already been taken? Bioresour. Technol. 250, 888–901.

1107

https://doi.org/10.1016/j.biortech.2017.11.102

1108

Silva, C.P., Lima, D.L.D., Schneider, R.J., Otero, M., Esteves, V.I., 2014. Evaluation of

1109

the anthropogenic input of caffeine in surface waters of the north and center of

1110

portugal

1111

https://doi.org/10.1016/j.scitotenv.2014.01.120

by

ELISA.

Sci.

Total

Environ.

479–480,

227–232.

1112

Simeonidis, K., Martinez-Boubeta, C., Rivera-Gil, P., Ashraf, S., Samaras, T.,

1113

Angelakeris, M., Tresintsi, S., Mitrakas, M., Parak, W.J., Monty, C., Balcells, L.,

1114

2017. Regeneration of arsenic spent adsorbents by Fe/MgO nanoparticles. J. Chem.

1115

Technol. Biotechnol. 92, 1876–1883. https://doi.org/10.1002/jctb.5187

1116 1117

Solanki, A., Boyer, T.H., 2017. Environmental Science Water Research & Technology 553–565. https://doi.org/10.1039/c6ew00224b

1118

Song, X.L., Zhang, M.W., Zhang, Y., Huang, S.T., Geng, B.Y., Meng, R. Bin, Yang,

1119

Y.Z., Zhong, Y.S., Liu, H.Y., 2010. Surface modification of coconut-based

1120

activated carbon by SDS and its effects on Pb2+adsorption. J. Cent. South Univ.

1121

20, 1156–1160. https://doi.org/10.1007/s11771-013-1598-4

1122

Sotelo, J.L., Rodríguez, A.R., Mateos, M.M., Hernández, S.D., Torrellas, S.A.,

1123

Rodríguez, J.G., 2012. Adsorption of pharmaceutical compounds and an endocrine

1124

disruptor from aqueous solutions by carbon materials. J. Environ. Sci. Heal. - Part

1125

B

Pestic.

Food

Contam.

Agric.

Wastes

47,

640–652.

ACCEPTED MANUSCRIPT 1126

56

https://doi.org/10.1080/03601234.2012.668462

1127

Sumalinog, D.A.G., Capareda, S.C., de Luna, M.D.G., 2018. Evaluation of the

1128

effectiveness and mechanisms of acetaminophen and methylene blue dye

1129

adsorption on activated biochar derived from municipal solid wastes. J. Environ.

1130

Manage. 210, 255–262. https://doi.org/10.1016/j.jenvman.2018.01.010

1131

Taylor, N.G.H., Verner-Jeffreys, D.W., Baker-Austin, C., 2011. Aquatic systems:

1132

Maintaining, mixing and mobilising antimicrobial resistance? Trends Ecol. Evol.

1133

https://doi.org/10.1016/j.tree.2011.03.004

1134

To, M., Hadi, P., Hui, C., Sze, C., Lin, K., Mckay, G., 2017. Mechanistic study of

1135

atenolol , acebutolol and carbamazepine adsorption on waste biomass derived

1136

activated

1137

https://doi.org/10.1016/j.molliq.2017.05.037

carbon.

J.

Mol.

Liq.

241,

386–398.

1138

Torrellas, S.Á., García Lovera, R., Escalona, N., Sepúlveda, C., Sotelo, J.L., García, J.,

1139

2015. Chemical-activated carbons from peach stones for the adsorption of

1140

emerging contaminants in aqueous solutions. Chem. Eng. J. 279, 788–798.

1141

https://doi.org/10.1016/j.cej.2015.05.104

1142

Tran, H.N., You, S.J., Hosseini-Bandegharaei, A., Chao, H.P., 2017. Mistakes and

1143

inconsistencies regarding adsorption of contaminants from aqueous solutions: A

1144

critical

1145

https://doi.org/10.1016/j.watres.2017.04.014

review.

Water

Res.

120,

88–116.

1146

Vasanth Kumar, K., Sivanesan, S., 2007. Isotherms for Malachite Green onto rubber

1147

wood (Hevea brasiliensis) sawdust: Comparison of linear and non-linear methods.

1148

Dye. Pigment. https://doi.org/10.1016/j.dyepig.2005.07.021

1149

Villaescusa, I., Fiol, N., Poch, J., Bianchi, A., Bazzicalupi, C., 2011. Mechanism of

1150

paracetamol removal by vegetable wastes: The contribution of Π-Π interactions,

ACCEPTED MANUSCRIPT

57

1151

hydrogen bonding and hydrophobic effect. Desalination 270, 135–142.

1152

https://doi.org/10.1016/j.desal.2010.11.037

1153

Wang, J., Bai, Z., 2017. Fe-based catalysts for heterogeneous catalytic ozonation of

1154

emerging

1155

https://doi.org/10.1016/j.cej.2016.11.118

1156 1157 1158 1159 1160 1161

contaminants

in

water

and

wastewater.

Chem.

Eng.

J.

Weber, W.J., Morris, J.C., 1963. Kinetics of Adsorption on Carbon from Solution. J. Sanit. Eng. Div. https://doi.org/10.1080/002689796173345 Wigmans, T., 1989. Industrial aspects of production and use of activated carbons. Carbon N. Y. https://doi.org/10.1016/0008-6223(89)90152-8 Worch, E., 2012. Adsorption Technology in Water Treatment, Walter de Gruyter. https://doi.org/10.1515/9783110240238

1162

Wu, M., Xiang, J., Que, C., Chen, F., Xu, G., 2015. Occurrence and fate of psychiatric

1163

pharmaceuticals in the urban water system of Shanghai, China. Chemosphere 138,

1164

486–493. https://doi.org/10.1016/j.chemosphere.2015.07.002

1165

Yang, Y.Y., Zhao, J.L., Liu, Y.S., Liu, W.R., Zhang, Q.Q., Yao, L., Hu, L.X., Zhang,

1166

J.N., Jiang, Y.X., Ying, G.G., 2018. Pharmaceuticals and personal care products

1167

(PPCPs) and artificial sweeteners (ASs) in surface and ground waters and their

1168

application as indication of wastewater contamination. Sci. Total Environ. 616–

1169

617, 816–823. https://doi.org/10.1016/j.scitotenv.2017.10.241

1170

Yuen, F.K., Hameed, B.H., 2009. Recent developments in the preparation and

1171

regeneration of activated carbons by microwaves. Adv. Colloid Interface Sci. 149,

1172

19–27. https://doi.org/10.1016/j.cis.2008.12.005

1173

Zbair, M., Ait, H., Anfar, Z., 2018. Porous carbon by microwave assisted pyrolysis : An

1174

effective and low-cost adsorbent for sulfamethoxazole adsorption and optimization

1175

using

response

surface

methodology.

J.

Clean.

Prod.

202,

571–581.

ACCEPTED MANUSCRIPT 1176

58

https://doi.org/10.1016/j.jclepro.2018.08.155

1177

Zhou, L., Zhou, H., Hu, Y., Yan, S., Yang, J., 2019. Adsorption removal of cationic

1178

dyes from aqueous solutions using ceramic adsorbents prepared from industrial

1179

waste

1180

https://doi.org/10.1016/j.jenvman.2019.01.009

coal

gangue.

J.

Environ.

Manage.

234,

245–252.

1181

Zhu, S., Liu, Y. guo, Liu, S. bo, Zeng, G. ming, Jiang, L. hua, Tan, X. fei, Zhou, L.,

1182

Zeng, W., Li, T. ting, Yang, C. ping, 2017. Adsorption of emerging contaminant

1183

metformin

1184

https://doi.org/10.1016/j.chemosphere.2017.03.071

using

graphene

oxide.

Chemosphere.

1185

Zolgharnein, J., Shahmoradi, A., Ghasemi, J., 2011. Pesticides Removal Using

1186

Conventional and Low-Cost Adsorbents: A Review. Clean - Soil, Air, Water.

1187

https://doi.org/10.1002/clen.201000306

1188

Zuccato, E., Castiglioni, S., Bagnati, R., Chiabrando, C., Grassi, P., Fanelli, R., 2008.

1189

Illicit drugs, a novel group of environmental contaminants. Water Res.

1190

https://doi.org/10.1016/j.watres.2007.09.010

1191

ACCEPTED MANUSCRIPT Highlights

    

Pharmaceuticals have been found in surface water in μg L-1 to ng L-1. Pharmaceuticals can be persistent in the environment. Adsorption is effective in removing organic and inorganic pollutants from water. Adsorption has been investigated for the removal of several pharmaceuticals. Municipal and agro-industrial wastes can be used as precursors of adsorbents.