Antiscalants in RO Membran Escaling Control

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Water Research 183 (2020) 115985

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Review

Antiscalants in RO membrane scaling control*


Wei Yu, Di Song, Wei Chen, Hu Yang*
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing, 210023, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Reverse osmosis (RO) plays an important role in freshwater production. Mineral scaling is an inevitable
Received 2 January 2020 problem in the RO desalination process. Various methods, including the pretreatment of feed water, the
Received in revised form optimization of operational processes, the development of novel membrane materials, and the addition
4 May 2020
of antiscalants, have been developed to mitigate scale formation in RO systems. Among these methods,
Accepted 24 May 2020
Available online 18 June 2020
the addition of antiscalants is a relatively cost-effective and convenient technique for membrane scaling
control. In the current work, various kinds of antiscalants, scale inhibition mechanisms, and their ap-
plications to RO membrane scaling control are reviewed. Weakness of existing antiscalants and challenge
Keywords:
Antiscalants
arising from their practical applications, such as membrane fouling caused by antiscalants, increased
Membrane scaling control bacterial growth, dosing control, and the disposal of resultant concentrates, are also presented. To
Reverse osmosis effectively alleviate scaling on RO membrane by using antiscalants, the development of novel, high-
Scale inhibition mechanisms performance, and environment-friendly antiscalants on the basis of an in-depth study of the inhibi-
Applications in RO tion mechanisms and well-established structureeactivity relationships is urgently necessary. The opti-
Prospects on antiscalants mization of antiscalants and their combinations with other pretreatments in practical RO operations are
essential in efficient scaling control.
© 2020 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Scale inhibition mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1. Scale formation and growth kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.1. Scale formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.2. Scale growth kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2. Normal inhibition mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.1. Metal ion-based scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.2. Silica-based scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3. Antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1. Phosphorus antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2. Synthetic polymeric antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3. Environment-friendly antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4. Evaluation of inhibition performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.1. Operation methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.2. Influencing factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.2.1. Dose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.2.2. pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.2.3. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.2.4. Coexisting substance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

*
Supported by the National Natural Science Foundation of China (grant nos.
51778279 and 51978325).
* Corresponding author.
E-mail address: [email protected] (H. Yang).

https://doi.org/10.1016/j.watres.2020.115985
0043-1354/© 2020 Elsevier Ltd. All rights reserved.
2 W. Yu et al. / Water Research 183 (2020) 115985

Abbreviations MW Molecular weight


NF Nanofiltration
AA Acrylic acid OM Organic matter
ANeAA Acrylonitrile-acrylic acid PA Polyacrylate
AOP Advanced oxidation process PAA Polyacrylic acid
ATMP Amino trimethylene phosphonic acid PALAM Polyallylamine hydrochloride
BSA Bovine serum albumin PAMALAM Poly(acrylamide-co-diallyldimethylammonium
CATIN Cationic inulin chloride)
CMC Carboxymethyl cellulose PAMAM Polyaminoamide
CMI Carboxymethyl inulin PASP Poly(aspartic acid)
CMS Carboxymethyl starch PBTCA 2-Phosphonobutane-1,2,4-tricarboxylic acid
DS Degree of substitution PEI Polyethyleneimine
EDTA Ethylenediaminetetraacetate PESA Polyepoxysuccinic acid
EPS Extracellular polymeric substance RO Reverse osmosis
GFH Granular ferric hydroxide SA Sodium alginate
HA Humic acid SHMP Sodium hexametaphosphate
LSI Langelier saturation index St-g-PAA Starch-graft-polyacrylic acid
MA Maleic acid s-EPS Soluble extracellular polymeric substance
MF Microfiltration

5. Applications of antiscalants in RO desalination systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14


5.1. Phosphorus antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.2. Synthetic polymeric antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.3. Environment-friendly antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
6. Existing problems and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6.1. Existing problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6.1.1. Enhanced biofouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6.1.2. Dosing control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
6.1.3. Concentrate disposal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
6.2. Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.2.1. Optimizing the application of antiscalants in RO process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.2.2. Environment friendliness of antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.2.3. Fabrication of high-performance antiscalants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Declaration of competing interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1. Introduction fouling, organic fouling, inorganic fouling, and biofouling (Andrews


et al., 2008; Fritzmann et al., 2007; Goh et al., 2018; Guo et al., 2012;
Reverse osmosis (RO), illustrated in Fig. 1, is one of the most Jiang et al., 2017; Malaeb and Ayoub, 2011; Matin et al., 2019).
widely used membrane-based desalination processes to alleviate Inorganic fouling, which generally refers to mineral scaling, is a
the water shortage due to its low energy consumption and high challenging problem due to the high amount of total dissolved
efficiency (Fritzmann et al., 2007). However, membrane fouling solids of feed water and the nearly complete salt rejection
poses an obstacle to the application of RO desalination processes. (Andrews et al., 2008). When salt concentration increases and ex-
Membrane fouling can be divided into four types: particulate ceeds the solubility of sparingly soluble salts, scale formation from

Fig. 1. Conventional RO desalination process.


W. Yu et al. / Water Research 183 (2020) 115985 3

bulk and surface crystallization easily occurs, causing membrane

Henthorne and Boysen, 2015; Jamaly et al., 2014;

Relatively high cost and not commercializable yet. (Goh et al., 2018; Kang and Cao, 2012; Tong et al.,

(Darton, 2000; Fritzmann et al., 2007; Greenlee


blocking and the deterioration of membrane performance. Com-

Membrane damage caused by high pressure and (Antony et al., 2011; Goh et al., 2018; Greenlee
(Anis et al., 2019; Badruzzaman et al., 2019;
mon mineral scaling includes metal ion-based scales, which are

frequent cleaning; increase energy consumption; et al., 2009a; Henthorne and Boysen, 2015;
mostly alkaline earth metal-based scales, and silica-based scales

Shenvi et al., 2015; Tong et al., 2019)


(Matin et al., 2019; Oh et al., 2009; Tong et al., 2019). To mitigate
the risk of membrane scaling, many methods, including feed

et al., 2009a; Tong et al., 2019)


water pretreatment, the optimization of operational processes,
2019; Zhang et al., 2016b) the development of novel membrane materials, and the addition
of antiscalants, have been developed in accordance with the
characteristics of RO operating systems (Fig. 1) and different feed
Tong et al., 2019)

water (Anis et al., 2019; Antony et al., 2011; Badruzzaman et al.,


2019; Goh et al., 2018; Greenlee et al., 2009a; Henthorne and
References

Boysen, 2015; Jamaly et al., 2014; Kang and Cao, 2012; Tong
et al., 2019). Table 1 summarizes the details of these methods
for mitigating membrane scaling, including their advantages and
disadvantages.
Requirement of additional equipment; corrosion

As indicated in Table 1, pretreatment processes, which typically


Increased biofouling; phosphorus emissions;

consist of softening, pH adjustment, and the use of larger pore size


membrane, are necessary to reduce scale formation (Anis et al.,
2019; Badruzzaman et al., 2019; Darton, 2000; Greenlee et al.,
reduced permeate production, etc.

2009a; Henthorne and Boysen, 2015; Jamaly et al., 2014; Kelle


Zeiher et al., 2003; Tong et al., 2019). Lime softening and ion ex-
difficult concentrate disposal.

change softening can remove scale-forming substances, such as


Ca2þ and Mg2þ, and decrease scaling potential in the subsequent
membrane procedure (Anis et al., 2019; Jamaly et al., 2014; Tong
et al., 2019). Adjusting the pH of feed water is effective for miti-
Disadvantages

problem, etc.

gating the formation of certain types of sparingly soluble salts,


because the solubility of carbonate minerals and silica scale is
significantly related to pH (Bush et al., 2018; Milne et al., 2014;
Tong et al., 2019). Pretreatment using larger pore size membrane
before the RO unit, including microfiltration (MF), ultrafiltration,
of operational spacer, applied pressure, hydrodynamic conditions, the setting of environment; remove existing
reduce the scaling propensity.

and nanofiltration (NF), is also an effective method for alleviating


Reduce the amounts of scale-

Inhibit scale formation by chelation, dispersion, crystal distortion Achieve high water recovery;
The optimization Feed water characteristics, crossflow velocity, the design of feed Improve membrane surface
forming ions in feed water;

simple operation; high cost


performance of membrane.

scale formation by separation of contaminants such as colloids,


Improve the antifouling

some organic matters (OMs), multivalent ions and part of mono-


valent ions (Anis et al., 2019; Badruzzaman et al., 2019; Jamaly
et al., 2014; Tong et al., 2019). These pretreatment methods play
performance.

important roles in the stable operation of a RO unit, but subse-


Advantages

quent problems caused by these methods, such as the wastes


deposits.

produced by lime softening, increased costs due to additional


equipment, and aggravated corrosion under low pH, cannot be
disregarded (Al-Rammah, 2000; Anis et al., 2019; Gill, 1999;
Hasson et al., 1998; Shahid et al., 2018). The overall design of a RO
Softening, pH adjustment, and the use of larger pore size

system exerts critical effects on its stable operation. Many aspects,


including feed water characteristics (e.g., pH, temperature,
composition), the selection and optimization of operational pa-
rameters (e.g., applied pressure, hydrodynamic conditions), target
recovery, and membrane cleaning, should be considered (Goh
Summary of various methods for control of membrane scaling.

The development Surface modification, physical blending.

et al., 2018; Greenlee et al., 2009a; Henthorne and Boysen, 2015;


membrane cleaning or target recovery.

Liu et al., 2019; Matin et al., 2019; Qasim et al., 2019; Tong et al.,
2019). The design and preparation of novel antifouling mem-
branes have been popular issues because membrane fouling is
closely related to membrane surface properties (Goh et al., 2018;
Tong et al., 2019; Zhang et al., 2016b). Tong et al. (2019) summa-
and threshold effect.

rized existing pieces of evidence that prove the importance of


membrane surface properties in regulating mineral scale forma-
tion during membrane desalination and discussed the challenges
membrane.

experienced in the design and fabrication of novel scaling-


Details

resistant membranes. Besides, the addition of antiscalants is one


of the most commonly applied methods for inhibiting scaling in
the RO process, especially in the situation that a high water re-
pretreatment

The addition of

covery is set to reduce concentrate volume but cause an enhanced


membrane

antiscalant
processes

materials
Feed water

of novel

scaling potential, due to its high efficiency, low cost, and easy
Methods

operation (Darton, 2000; Rahardianto et al., 2007; Tong et al.,


Table 1

2019). Antiscalants comprise a class of water treatment agents


that can inhibit scale formation through one or more effective
4 W. Yu et al. / Water Research 183 (2020) 115985

mechanisms, such as chelation, dispersion, crystal distortion, and


threshold (Darton, 2000; Li et al., 2011). Many reports have
confirmed that the use of antiscalants in RO processes can effec-
tively mitigate scale formation and assure the long-term stable
operation of RO systems (Al-Shammiri et al., 2000; Darton, 2000;
Fritzmann et al., 2007; Greenlee et al., 2009a; Tong et al., 2019).
Fig. 2. The process of crystal nucleation and growth (Darton, 2000).
Thus, a comprehensive understanding of the effects and mecha-
nisms of antiscalants in RO systems will be extremely helpful in
improving antiscalants’ performance and availability. chemical components (Matin et al., 2019; Tong et al., 2019). Metal
In fact, research on the applications and mechanisms of anti- ion-based scale, including calcium carbonate, magnesium carbon-
scalants to RO systems remains unsystematic, although the addi- ate, calcium sulfate, and barium sulfate, is generally formed via
tion of antiscalants to RO systems has been confirmed and classified crystallization. The formation of scales, i.e., the growth of crystal,
as one of the popular pretreatment methods in many studies (Anis occurs as described in Fig. 2. (1) When the salt concentration is over
et al., 2019; Greenlee et al., 2009a; Henthorne and Boysen, 2015; the solubility of sparingly soluble salts, the ions start colliding to
Jamaly et al., 2014; Qasim et al., 2019; Tong et al., 2019). However, form ion pairs and cluster as micronuclei. (2) Some of the micro-
most previous reviews focused on the development of novel anti- nuclei become nucleation centers, and then ion clusters begin
scalants and their traditional applications to the boiler water and aligning in an “orderly” manner to form stable nuclei. (3) Crystals
cooling water industries (Chaussemier et al., 2015; Hasson et al., grow gradually from nuclei, and this process is irreversible (Al-
2011). Thus, this work presented exhaustive knowledge about the Roomi and Hussain, 2016; Darton, 2000; Matin et al., 2019).
applications and mechanisms of antiscalants to RO systems. The Silica-based scale can be classified into silica scaling and metal
effects of various internal and external factors on the scale forma- silicates (Ning, 2002; Topçu et al., 2017). The formation of silica-
tion and the antiscalants’ performance were summarized. Finally, based scale is a complex process resulting from silica polymeriza-
the problems caused by existing antiscalants and the limitations tion and the accumulation of amorphous colloidal silica, which
and prospects of antiscalants in practical RO applications were depends on various conditions including pH, temperature, ionic
discussed in detail. Scheme 1 describes the overview of various strength, the presence of multivalent ions, and the silica concen-
antiscalants and their applications in RO membrane scaling control. tration (Bush et al., 2018; Milne et al., 2014). Silica polymerization is
considered to occur via dehydration and formation of SieOeSi
anhydride bonds: n Si(OH)4 / (OH)3SieOeSi(OH)3 dimer /
2. Scale inhibition mechanisms oligomers / colloidal polymers/ infinite extensions (SiO2)n, as
shown in Fig. 3 (Ketsetzi et al., 2008; Milne et al., 2014; Ning, 2002;
2.1. Scale formation and growth kinetics Weng, 1995; Zhang et al., 2012). The existence of multivalent cat-
ions in a silica-containing solution is one of main causes to decrease
2.1.1. Scale formation the solubility of silica and easily form metal silicates due to the
Mineral scaling can be approximately categorized into metal interactions between multivalent cations and silicate anions,
ion- and silica-based scales based on the characteristics of their

Scheme 1. Schematic overview of various antiscalants in RO membrane scaling control.


W. Yu et al. / Water Research 183 (2020) 115985 5

Fig. 3. The process of silica polymerization (Milne et al., 2014; Zhang et al., 2012).

expressed by
" #
bg3 Vm 2 f ðqÞ
J ¼ A,exp  (1)
k3 T 3 ðln SÞ2

where A is a frequency factor; b is a geometric (shape) factor of 16p/


3 for the spherical nucleus; and f(q) is a correction factor, which is
equal to 1.0 for homogeneous nucleation and to <1.0 for hetero-
geneous nucleation (Hasson et al., 2003). Several researchers
(Abdel-Aal et al., 2015; El-Shall et al., 2005) have used f(q) ¼ 0.01 to
estimate the heterogeneous nucleation rate. Vm is the molar volume
Fig. 4. Scale formation mechanisms in RO system (Matin et al., 2019; Oh et al., 2009; of the phase forming, and g is the surface energy (J/m2). S is the
Shirazi et al., 2010). supersaturation ratio, k is the Boltzmann constant, and T is the
absolute temperature.
The induction time tind is inversely proportional to J (Hasson
resulting in scale formation even at concentrations below silica et al., 2001, 2003; Liu et al., 2019), and the relation between tind
saturation (Milne et al., 2014; Ning, 2002; Weng, 1995). and S can be written as
In RO systems, the widely accepted membrane scaling can be
classified into bulk (homogeneous) crystallization and surface bg3 Vm 2 f ðqÞ
(heterogeneous) crystallization, as shown in Fig. 4 (Goh et al., 2018; ln tind ¼ B þ (2)
k3 T 3 ðln SÞ2
Matin et al., 2019; Oh et al., 2009; Qasim et al., 2019; Shirazi et al.,
2010). On the one hand, solutions are concentrated several times in
where B is a constant. Thus, surface energy can be determined
accordance with water recovery. When salt concentration increases
experimentally by measuring the slope of the straight line that
and exceeds the solubility of sparingly soluble salts, scales are
correlates lntind with 1/(lnS)2.
formed in the solutions, namely, the products of bulk crystalliza-
For surface crystallization, surface crystallization can be greatly
tion, and deposited onto membrane surface, causing membrane
enhanced by concentration polarization on the membrane surface
blocking. On the other hand, scales can directly form and grow on
in addition to supersaturation level (Goh et al., 2018; Matin et al.,
membrane surface under enhanced concentration polarization and
2019; Semiat et al., 2001; Shirazi et al., 2010). Concentration po-
salt supersaturation (Matin et al., 2019). The prevailing scale for-
larization refers to the phenomenon in which a boundary layer
mation in RO process is influenced by several factors, such as su-
exists near the membrane surface, within which solute concen-
persaturation level, feed water characteristics, and hydrodynamics.
tration is higher than bulk solute concentration, due to the
Generally, bulk crystallization is more obvious at high supersatu-
convective flow from the bulk feed toward the membrane surface
ration level and low crossflow velocity (Matin et al., 2019). Surface
(Qasim et al., 2019). The concentration polarization level (CP) can be
crystallization is closely related not only to supersaturation level
expressed by (Hasson et al., 2001, 2003; Matin et al., 2019;
but also to concentration polarization on the membrane surface
Sutzkover et al., 2000):
(Goh et al., 2018; Matin et al., 2019; Semiat et al., 2001; Shirazi et al.,
2010). Silica scaling in RO systems is similar to the metal ion-based Cm Jv

scale formation in the two crystallization processes mentioned CP ¼ ¼ ekm (3)


Cb
earlier, including deposits of amorphous colloidal silica and direct
scaling on membrane surface. where Cm and Cb are the salt concentrations at the membrane
surface and bulk solution, respectively. Jv is the permeate flux, and
km is the mass transfer coefficient. The evaluation of CP requires
2.1.2. Scale growth kinetics
estimating the mass transfer coefficient (km). Sutzkover et al.
Scale crystal growth kinetics, which is based on the online
(2000) reported a simple technique to determine km and CP in a
observation and monitoring of the entire crystallization processes,
RO system. km is given by
is an extremely useful method for investigating the detailed scaling
process (Agarwal and Berglund, 2004). The measurements of
ðJv Þsalt
changes in bulk fluid properties, such as turbidity, conductivity, ion km ¼ ( " #) (4)
ðJ Þ
concentration, and pH, are typically applied to obtain the infor- ln pbDPpp , 1  ðJvvÞ salt
mation of induction time (tind) and crystal growth rate (R) (Abdel- H O2

Aal et al., 2015; Fischer et al., 2011; Lee et al., 2015; Olderøy et al.,
2009; Rabizadeh et al., 2017). On the basis of classic nucleation where pb and pp are the osmotic pressure of the saline feed and the
theory (Boerlage et al., 2002; Dydo et al., 2004; Hasson et al., 2003; permeate, respectively; (Jv)H2O and (Jv)salt are the permeate flux of
Manoli and Dalas, 2001; Mullin, 2001), the nucleation rate (J) can be the salt-free water and the saline solution, respectively; and DP is
6 W. Yu et al. / Water Research 183 (2020) 115985

the transmembrane pressure. typically influence bulk crystallization of metal ion-base scale by
Cm can be estimated using Eq. (3). Then, the supersaturation interfering with one or more of the crystallization stages, because
ratio in the membrane system (Sw) that prevails on the membrane these chemical agents with various functional groups, such as
surface for a type of sparingly soluble salt AxBy can be calculated phosphonic, carboxylic, and sulfonic groups, can interact with
using the following equation: scale-forming substances and inhibit scale formation through one
or more of the widely accepted inhibition mechanisms, which
½CA m y ½CB m x ðCP,½CA b Þy ðCP,½CB b Þx CP xþy ½CA b y ½CB b x include chelation, dispersion, lattice distortion, and threshold ef-
Sw ¼ ¼ ¼
Ksp Ksp Ksp fects, as summarized in Table 2 (Al-Roomi and Hussain, 2016;
(5) Amjad, 2014; Darton, 2000; Dobberschütz et al., 2018; Li et al.,
2011; Rahman, 2013).
where [CA]m and [CB]m refer to membrane surface concentration, Chelation refers to the complexation of antiscalants with scale-
[CA]b and [CB]b refer to bulk concentration, and Ksp is the solubility forming cations via a stoichiometric reaction, which can increase
product. In summary, the supersaturation level for a certain salt on the solubility of sparingly soluble salts (Al-Roomi and Hussain,
a membrane surface can be obtained in accordance with CP based 2016; Darton, 2000). Dispersion effect is that antiscalants are
on Eqs. (3)e(5). High CP can lead to a high solute concentration in adsorbed onto crystals, generally providing them with surface
the boundary layer causing increased osmotic pressure and su- charge, and causing mineral crystals to repel one another through
persaturation level, which enhance membrane surface crystalliza- steric interactions, keeping them dispersed in the solution (Al-
tion and increase the risk of scaling and reducing permeate flux Roomi and Hussain, 2016; Darton, 2000; Li et al., 2011). For
(Matin et al., 2019; Qasim et al., 2019). In general, increased applied enveloping crystals and providing them with surface charge, anti-
pressure and water recovery can lead to severe concentration po- scalants with a relatively high molecular weight (MW)
larization; a high crossflow velocity can move formed particles (>20,000 Da) are suitable as effective dispersants (Darton, 2000).
away from the membrane surface and reduce concentration po- Antiscalants can disrupt the crystallization process by interacting
larization (Goh et al., 2018; Matin et al., 2019; Qasim et al., 2019; with mineral nuclei, inhibiting and/or distorting crystal growth (Al-
Shirazi et al., 2010). Roomi and Hussain, 2016; Darton, 2000; Li et al., 2011; Liu et al.,
The surface energy for a process of surface nucleation on 2019). The formed crystal morphologies can be changed due to
membrane surface can be also estimated using tind and the obtained the adsorption of antiscalants, called lattice distortion. Influencing
Sw based on Eq. (2). When Eqs. (2) and (5) are combined, the the clustering process and preventing the precipitation of sparingly
following equation that reflects membrane surface crystallization soluble salts via sub-stoichiometric amounts of antiscalants is
based on CP can be obtained as considered threshold effect (Darton, 2000; Ketrane et al., 2009).
The inhibition and distortion of scale formation in the crystal
bg3 Vm 2 f ðqÞ growth process are caused by the adsorption of antiscalants onto
ln tind ¼ B þ  2 (6) the active sites of crystal surfaces. Exploring the adsorption
CP xþy ½CA b y ½CB b x
k3 T 3 ln Ksp behavior of antiscalants on crystal surfaces can provide in-depth
understanding of scale inhibition mechanisms. The coverage (q)
From Eqs. (2) and (6), determining the onset of membrane of antiscalants on the crystal surface can changed the crystal
scaling and tind is crucial for investigating the membrane scaling growth rate in the absence of antiscalants (R0) to Ri, as shown in the
following equation (Akin et al., 2008; Kırbog €
a and Oner, 2009;
process. Hasson et al. (2001,2003) extracted tind data from plots of
permeate flux versus time, which were successfully correlated with Nielsen et al., 2012):
the supersaturation level based on nucleation theory. The mem-
brane scaling can be assessed by tind and obtained crystallization Ri ¼ R0 ð1  qÞ (7)
surface energy. However, difficulties and uncertainties remain in The adsorption behavior can be investigated by fitting the
determining tind on the membrane surface. Heterogeneous nucle- experimental data that correspond to different antiscalant con-
ation on membrane surface is clearly favored, which causes surface centrations with a suitable adsorption isothermal model; accord-
crystallization even at low supersaturation levels; hence, incipient ingly, the adsorption constant can be obtained and compared (Akin
membrane scaling may not cause an evident change in the €
a and Oner,
et al., 2008; Kırbog 2009; Lin and Singer, 2005; Nielsen
measured bulk fluid properties or permeate flux (Karabelas et al.,
et al., 2012). Langmuir-type model has been widely used to inter-
2020). van de Lisdonk et al. (2001) presented a supersaturation
pret the decrease in growth rates of some sparingly soluble salts in
prediction model aiming at monitoring the early stage of scaling,
the presence of antiscalants, and the coverage q for this adsorption
which can calculate the supersaturation ratio of sparingly soluble
model is described by the following equation (Akin et al., 2008;
salts on membrane surface. The scaling monitor was proved to be €
Kırbog a and Oner, 2009; Nielsen et al., 2012):
able to detect scaling before it occurred in the last stage of the pilot
plant. Besides, real-time membrane observation is a sensitive
KCi
method for identifying incipient scaling but with limitations in q¼ (8)
1 þ KCi
spatial resolution (Karabelas et al., 2020). Reliable techniques for
monitoring incipient scaling are important for the quantitative
where K is the affinity constant, which is the ratio of the rate
assessment of antiscalant inhibitory effectiveness in RO systems,
constants for adsorption and desorption (kads/kdes); and Ci is the
and further effort is required to develop useful monitoring
total equilibrium concentration of an antiscalant. From Eqs. (7) and
methods.
(8), the relationship between Ri and Ci can be given as

R0 1
2.2. Normal inhibition mechanisms ¼1 þ (9)
R0  Ri KCi
2.2.1. Metal ion-based scale The affinity constant K can be obtained from the slope of the
Different mineral scales have various features and follow linear relationship between R0/(R0Ri) and 1/Ci, as a measure of the
different inhibition mechanisms. In general, antiscalants can adsorption affinity of antiscalants on the crystal surface. Akin et al.
W. Yu et al. / Water Research 183 (2020) 115985 7

Table 2
Scale inhibition mechanisms (Al-Roomi and Hussain, 2016; Darton, 2000; Li et al., 2011; Rahman, 2013).

Scale inhibition Description Illustration Typical antiscalants


mechanism

Chelation Chelate with cationic ions to increase the solubility of sparingly soluble salts. Phosphorus–containing antiscalants,
polymeric antiscalants

Dispersion Generally, provide mineral crystal surface charge and keep them dispersed in Polymeric antiscalants
solution by electrostatic repulsion.

Crystal lattice Interfere with normal crystal growth to distort the crystallization process. Phosphorus–containing antiscalants,
distortion polymeric antiscalants

Threshold effect Disrupt the clustering and ordering processes with sub-stoichiometric Phosphorus–containing antiscalants,
amounts of inhibitor. polymeric antiscalants

(2008) studied the effect of carboxymethyl inulin (CMI) on the metal silicates, the required antiscalants and the corresponding
crystal growth of calcium oxalate. The results indicated that tind inhibition mechanisms are quite different. For metal silicates,
increased and Ri decreased with increasing carboxyl content on the similar to the aforementioned popular metal ion-based scale, the
polymer backbone. The adsorption of polymeric antiscalants on the interactions of antiscalants with multivalent cations should be the
crystal can be fitted in a Langmuir-type isothermal model; the primary inhibition mechanisms. The antiscalants that contain
obtained affinity constant was high for the CMI with a high anionic groups may increase the solubility of metal silicates by
carboxylation degree, which may reflect its strong adsorption onto interacting with multivalent cations (Topçu et al., 2017), thus pre-
the crystal surface (Akin et al., 2008). venting the formation of metal silicates.
Molecular dynamics simulations have also been conducted to For silica scaling, the inhibition mechanisms mostly include
investigate the interaction between the crystal surface and addi- interfering with the silica polymerization process and dispersing
tives (Chen et al., 2013; Ji et al., 2017; Shen et al., 2018; Shi et al., colloidal silica particles (Weng, 1995; Zhang et al., 2011). The sur-
2012, 2013, 2017; Zhang et al., 2016a). The binding energy (Eb) of face charges of various silica species are negative within a wide pH
antiscalants on the crystal surface can be computed using the region due to the deprotonation of surface silanol groups (Schindler
following equation: €berg, 1996; Weng, 1995; Zhang et al., 2011), con-
et al., 1976; Sjo
  ventional antiscalants that contain anionic groups thus exhibit
Eb ¼  Eads ¼  Esystem  Esurface  Eantiscalant (10) nearly no inhibition efficiency against silica scale (Goh et al., 2018;
Neofotistou and Demadis, 2004b; Zhang et al., 2011). It is expected
where Esystem is the total energy of the antiscalantesurface system, that positively charged antiscalants can interact with negatively
Esurface is the energy of the crystal surface, and Eantiscalant is the en- charged silica via electrostatic attraction, thereby not only inter-
ergy of the free antiscalant molecule. The larger the binding energy, fering with the polymerization of silica monomers or soluble silica
the more favorable the interactions between the crystal surface and oligomers, but also dispersing colloidal silica particles and pre-
the antiscalant (Shi et al., 2017; Zhang et al., 2016a). This method venting their agglomeration, as shown in Fig. 5 (Topçu et al., 2017;
can provide a theoretical basis for the interaction between anti- Zhang et al., 2011).
scalant and crystal surface at the molecular level.
In addition to bulk crystallization, the addition of antiscalants 3. Antiscalants
may inhibit surface crystallization by changing the surface energy
of heterogeneous nucleation and decreasing the nucleation rate, 3.1. Phosphorus antiscalants
which can be reflected by Eq. (6) (Hasson et al., 2001, 2003).
However, antiscalants can also influence surface crystallization via Antiscalants can be approximately classified into phosphorus
adsorption onto membrane surface and alteration of membrane and phosphorus-free antiscalants (Al-Roomi and Hussain, 2016;
surface properties, sometimes causing aggravated membrane Hasson et al., 2011; He et al., 2009; Ketrane et al., 2009; Li et al.,
scaling, which would be discussed in detail later (Ang et al., 2016; 2011). Phosphorus antiscalants include phosphates and phospho-
Benecke et al., 2018). nates (Nowack, 2003; Rashchi and Finch, 2000). The two structures
are different. Phosphates are defined as compounds that contain
PeO linkage, with the orthophosphate ion (PO3- 4 ) as the basic
2.2.2. Silica-based scale building unit (Rashchi and Finch, 2000), such as sodium tripoly-
Given the considerable difference between silica scaling and phosphate and sodium hexametaphosphate (SHMP), as shown in
8 W. Yu et al. / Water Research 183 (2020) 115985

Fig. 5. The available interactions between silica and cationic antiscalants (Zhang et al., 2011).

Fig. 6. Some typical phosphate antiscalants: (a) phosphate, (b) tripolyphosphate, and (c) hexametaphosphate.

Fig. 7. Some typical phosphonate antiscalants: (a) phosphonate, (b) 1-hydroxyethylidene-1,1-diphosphonic acid, (c) amino trimethylene phosphonic acid (ATMP), and (d) 2-
phosphonobutane-1,2,4-tricarboxylic acid (PBTCA).

Fig. 6. Phosphonates are compounds containing one or more C- Finch, 2000), phosphates easily hydrolyze into orthophosphate
PO(OH)2 groups, which are more stable than phosphates due to the ions, which may form precipitates with calcium ions and increase
existence of stable covalent CeP bond (Nowack, 2003). The struc- the potential of biofouling (Al-Roomi and Hussain, 2015; Andrews
tures of some typical phosphonates are presented in Fig. 7. Phos- et al., 2008; Darton, 2000; Ghani and Al-Deffeeri, 2010; Li et al.,
phates and phosphonates can be both used as complexing agents in 2017). Phosphonates thus have a wider range of applications
many applications due to their excellent complexing ability. because of their structural stability.
Although phosphate was found to prevent scale formation or In short, as the most popular inhibitors, phosphorus inhibitors
dissolve precipitates by forming relatively stable soluble complexes including phosphates and phosphonates account for a large share
with metal ions as early as the end of the 19th century (Rashchi and in the antiscalants’ market considering their good cost-effective
W. Yu et al. / Water Research 183 (2020) 115985 9

based scale, some cationic polymers have been used in inhibiting


silica scale formation (Chen et al., 2017; Demadis and Neofotistou,
2007; Demadis and Stathoulopoulou, 2006; Neofotistou and
Demadis, 2004a, Neofotistou and Demadis, 2004b; Zhang et al.,
2011). Anionic antiscalants have no effect on silica scale forma-
tion due to the special properties of silica scale that differ from
those of metal ion-based scale (Neofotistou and Demadis, 2004b).
By contrast, some cationic polymers are effective at preventing
silica scale formation (Ketsetzi et al., 2008; Neofotistou and
Demadis, 2004a, Neofotistou and Demadis, 2004b). Several
studies have reported the inhibition effects of some cationic poly-
mers, such as polyaminoamide (PAMAM) dendrimers with eNH2
termini (Neofotistou and Demadis, 2004a, Neofotistou and
Demadis, 2004b), adipic acid/amine-terminated polyether D230/
diethylenetriamine copolymer (Zhang et al., 2011), synthesized
cationic inulin (CATIN) (Ketsetzi et al., 2008), polyethyleneimine
(PEI), polyallylamine hydrochloride (PALAM), and poly(acrylamide-
co-diallyldimethylammonium chloride) (PAMALAM) (Demadis and
Stathoulopoulou, 2006), on silica scale. These studies have shown
that inhibition performance is related to polymeric structure and
cationic charge density (Demadis and Stathoulopoulou, 2006;
Fig. 8. Some typical anionic monomers used in the synthesis of polymeric anti- Ketsetzi et al., 2008). Positively charged polymers can associate
scalants: (a) acrylic acid (AA), (b) maleic acid (MA), and (c) 2-acrylamido-2- with negatively charged silica monomers or soluble silica oligo-
methylpropane sulfonic acid. mers, and thus, interfere with the silica polymerization process and
prevent the formation of colloid silica particles (Zhang et al., 2011).
Different cationic polymeric antiscalants, such as PAMAM, PEI,
performance (Gryta, 2012; Lin and Singer, 2005). With increasing
PALAM, and PAMALAM, with various cationic groups, i.e., weak
focus on environmental issues, phosphorus emissions that will lead
cationic groups (primary and secondary amines, eNHþ 3 and
to heavy eutrophication and increased biofouling have been
eRNHþ 2 ) and strong cationic groups (quaternary ammonium,
receiving considerable attention, and thus, their utilization is
eNRþ 3 ), typically exhibit different inhibition performance (Demadis
gradually restricted (Al-Roomi and Hussain, 2015).
and Stathoulopoulou, 2006; Ketsetzi et al., 2008; Neofotistou and
Demadis, 2004b), although all of them can interact with nega-
3.2. Synthetic polymeric antiscalants tively charged silica through electrostatic effects. Strong cationic
polymers that contained quaternary ammonium salt groups,
The development of low-phosphorus or phosphorus-free scale PALAM and PAMAMLAM, demonstrated relatively poor perfor-
inhibitors has become a major issue to reduce phosphorus emis- mance against silica compared with PEI which contains amine
sions from the use of antiscalants. Phosphorus antiscalants have groups, accordingly, strong cationic groups may lead to decreased
been partly replaced with synthetic polymeric scale inhibitors inhibition performance, which may be due to the distinct in-
without phosphorus. For common metal ion-based scale, the car- teractions with silica from different cationic groups causing various
boxylic group on the polymeric backbone is an effective functional effects (Demadis and Stathoulopoulou, 2006). Besides, cationic
group for inhibiting scale formation (Yang et al., 2017; Zhang et al., polymers with strong cationic groups are typically ready to
2016c). The class of polycarboxylic acid inhibitors synthesized aggregate silica colloids through their evident coagulation effects
through the polymerization of monomers that contain carboxylic apart from their antiscaling effect (Dao et al., 2016; Yang et al.,
groups can interact with scale-forming substances via carboxylic 2016). However, the specific relationship between silica and anti-
groups and efficiently prevent scale formation (Al-Hamzah et al., scalants with different molecular structures is not yet well under-
2014; Kavitha et al., 2011; Senthilmurugan et al., 2010; Tang et al., stood. Thus, further research on the interactions between
2008). In addition to carboxylic groups, inhibitors that contain antiscalants and silica-based scaling is required to gain an in-depth
multiple types of functional groups exhibit better inhibition per- understanding of the inhibition mechanisms for silica-based scale
formance due to synergistic effects compared with those with a and to develop effective antiscalants against silica-based scaling.
single functional group, such as hydroxyl, sulfonic, amino, and In addition to ionic antiscalants, water-soluble nonionic poly-
ether groups (Chen et al., 2015a; Dietzsch et al., 2013; Shakkthivel mers, such as polyacrylamide (Lee et al., 2015; Wang et al., 2006; Yu
and Vasudevan, 2006; Yang et al., 2017; Zhang et al., 2016c). The et al., 2006) and poly(vinyl alcohol) (Lakshminarayanan et al.,
introduction of multiple types of functional groups into polymeric 2003), also have evident effects on CaCO3 crystallization. Crystal
antiscalants via the copolymerization of various monomers in- morphologies can be influenced by the interference of polymers
creases water solubility and improves the dispersion effect of that results from the crystal distortion effect (Lakshminarayanan
polymers, enhancing the inhibition performance of antiscalants et al., 2003; Lee et al., 2015; Wang et al., 2006; Yu et al., 2006).
(Dietzsch et al., 2013; Gao et al., 2015; Li et al., 2017; Shakkthivel Apart from the electric nature of functional groups, the content of
and Vasudevan, 2006; Yang et al., 2017). Some typical anionic functional groups on polymeric antiscalants plays an important role
monomers used in the synthesis of antiscalants are presented in in their scale inhibition performance (Chen et al., 2015a). A high
Fig. 8. These synthetic polymeric antiscalants can effectively chelate content of ionic groups typically results in high inhibition perfor-
with metal ions and disperse microcrystal particles in solutions due mance (Wang et al., 2017b; Yu et al., 2019). However, a suitable
to the existence of hydrophilic anionic groups, and thus, show content of ionic groups is sometimes required, and a high charge
excellent scale inhibition efficiency (Shakkthivel and Vasudevan, density is occasionally detrimental to inhibition behavior because it
2006; Yang et al., 2017). may enhance other effects, such as the flocculation effect (Wang
Apart from anionic polymeric antiscalants against metal ion- et al., 2017b; Yu et al., 2018). Chen et al., 2015a used serine and
10 W. Yu et al. / Water Research 183 (2020) 115985

polysuccinimide to prepare modified poly(aspartic acid) (Ser-PASP) (Chaussemier et al., 2015; Hasson et al., 2011). At present, the
and found Ser-PASP obtained at a molar ratio of 0.8 exhibited better most promising alternatives to conventional phosphorus or
inhibition performance against CaCO3 than those of Ser-PASP at nonbiodegradable synthetic polymeric inhibitors are PASP, poly-
high or low molar ratio and poly(aspartic acid) (PASP), which may epoxysuccinic acid (PESA), and their derivatives (Chaussemier et al.,
be attributed to the synergistic effect induced at suitable contents 2015; Hasson et al., 2011). Their structures are presented in Fig. 9.
and ratio of the introduced hydroxyl and carboxylic groups. PASP and PESA have good chelating ability and dispersibility,
Moreover, polymer-based materials exhibit the structural char- making them as general purpose antiscalants, which have been
acteristics of multiplicity, which considerably differs from those of applied in various fields, such as industrial circulating water, boiler
small molecular materials (Elias, 1984; Flory, 1953). Apart from water, oil field water, and RO desalination systems (Chaussemier
different contents and types of chemical compositions on the local et al., 2015; Chen et al., 2015a; Hasson et al., 2011; Liu et al.,
chains mentioned earlier, the structural morphology, molecular 2012a; Xu et al., 2012; Zhang et al., 2016c). To improve the
size, and conformation of polymers due to their distinct long-chain comprehensive inhibition performance and application range of
structural features are particularly important determinants of final PASP and PESA, considerable research has focused on introducing
performance, including the inhibition performance of polymeric new functional groups onto them through various chemical
antiscalants. For example, MW, which is the most basic and modification processes (Chaussemier et al., 2015; Chen et al.,
important structural parameter of polymers and indicates macro- 2015a; Gao et al., 2015; Pramanik et al., 2017; Zhang et al., 2016c,
molecular size, exerts considerable influence on scale inhibition 2017).
performance (Darton, 2000; Matin et al., 2019; Shen et al., 2018; Apart from PASP and PESA, some natural polymers that contain
Wang et al., 2017b). A relatively low MW is generally favorable for abundant functional groups with good chelating and dispersion
suppressing crystallization, and a relatively high MW is required to effects, such as starch, cellulose, lignin, inulin, tannin, and collagen,
act as dispersants (Darton, 2000; Matin et al., 2019). Polyacrylic acid are also considered sources of potential environment-friendly
(PAA), which is the most common synthetic polymeric antiscalant, antiscalants due to their advantages of abundant source, low cost,
with different MWs has been found to exhibit varying inhibition and biodegradability (Boels and Witkamp, 2011; Chauhan et al.,
behavior, such as acting as a threshold agent within an MW range of 2012; Chaussemier et al., 2015; Guo et al., 2013; Hasson et al.,
1000e3000 Da, exhibiting lattice distortion effect at an MW of 2011; Qiang et al., 2013; Wang et al., 2017b; Yu et al., 2018, 2019;
5000e10,000 Da, and working as a dispersion agent at an MW of Zhang et al., 2016a). The applications of these natural polymers and
20,000e40,000 Da (Darton, 2000). However, polymeric anti- their derivatives to scale inhibition have been reported, and studies
scalants with excessively high MW or under overdose typically have indicated that suitable chemical modifications can consider-
result in poor inhibition performance, which may be attributed to ably improve their inhibition performance (Boels and Witkamp,
their bridging flocculation effects (Ali et al., 2015; Shen et al., 2018; 2011; Chauhan et al., 2012; Goncharuk et al., 2012; Guo et al.,
Wang et al., 2017b; Yu et al., 2018). For polymeric antiscalants with 2013; Hasson et al., 2011; Qiang et al., 2013; Wang et al., 2017b;
multiple types of functional groups, particularly for copolymers, Yu et al., 2018, 2019; Zhang et al., 2013). Fig. 10 presents the
their distributions on the polymeric backbone, i.e., the sequence structures of several modified natural polymers that can be used as
structure, are evidently influential in scale inhibition, in addition to inhibitors.
the contents and types of functional groups (Wang et al., 2016b; Yu Some OMs secreted by bacteria are also considered potential
et al., 2018). environment-friendly bio-antiscalants because they contain large
Thus, the molecular structure of polymeric antiscalants should amounts of effective functional groups (Li et al., 2019). Soluble
be focused on and considered a key factor for improving the design extracellular polymeric substances (s-EPS), mostly including pro-
of novel effective antiscalants. In contrast with the hydrolysis of teins, polysaccharides, and humic-like substances, extracted from
phosphorus-based inhibitors, these synthetic polymers are more Bacillus cereus (B. cereus) were evaluated as dual bio-functional
resistant to high operating temperatures (Hasson et al., 2011). This corrosion and scale-inhibiting agents (Li et al., 2019); they
also means that they are nonbiodegradable, which may have a demonstrated good anti-scale and anti-corrosion performance
profound environmental influence after their disposal (Hasson through their chelating, adsorption, and biomineralization abilities.
et al., 2011; Matin et al., 2019). With the advantages of biodegradability, strong biosorption affin-
ity, and multiple fixation sites, B. cereus s-EPS may offer a green,
3.3. Environment-friendly antiscalants sustainable, and economic strategy for developing environment-
friendly antiscalants. However, these biomaterials are also consid-
Considering environmental impact, the development of scale ered biofouling, reducing membrane flux because of their easy
inhibitors has moved toward environment-friendly antiscalants adsorption onto membrane surface (Matin et al., 2011). Thus, bio-
that are phosphorus-free, biodegradable, and cost-effective antiscalants can be effectively utilized in non-membrane systems,

Fig. 9. Molecular structures of (a) PASP and (b) PESA.


W. Yu et al. / Water Research 183 (2020) 115985 11

Fig. 10. Some modified natural polymeric antiscalants: (a) carboxymethyl inulin (CMI), (b) carboxymethyl cellulose (CMC), and (c) starch-graft-polyacrylic acid (St-g-PAA).

such as circulating cooling water, but should be carefully controlled et al., 2015). With regard to evaluating scale inhibition perfor-
in membrane systems. mance in dynamic RO systems, a laboratory-scale RO unit is typi-
Similar to synthetic polymeric antiscalants, the molecular cally used to simulate an actual RO process; the scale inhibition
structures of environment-friendly antiscalants, not only the con- performance can be determined by comparing the changes in
tents and types of functional groups but also the distributions of permeate flux with and without antiscalant addition; besides,
these functional groups, long-chain morphologies, and MW, which direct visual membrane surface monitoring by using an electron
are associated with their specific interactions with various scale- microscopy is also a useful method to evaluate scale surface
forming substances, should be focused on and studied in depth to coverage and surface number density that can reflect the effec-
understand their inhibition mechanisms and to design and fabri- tiveness of antiscalants in RO processes (Bartman et al., 2011;
cate novel, high-performance, and environment-friendly Hasson et al., 2001; Lyster et al., 2010; Rahardianto et al., 2008; Shih
antiscalants. et al., 2005; Shmulevsky et al., 2017; Thompson et al., 2017;
Uchymiak et al., 2008). Compared with those of static methods, the
4. Evaluation of inhibition performance operational conditions of RO simulation tests can be closer to those
of actual RO systems by adjusting operational parameters,
4.1. Operation methods including pressure, crossflow velocity, and feed spacer. The evalu-
ation results from dynamic RO simulation tests are certainly more
Proper and accurate evaluation of inhibition performance can instructive.
provide useful guidance such as the selection of antiscalants and
their approximate doses for the effective utilization of antiscalants 4.2. Influencing factors
in practical applications. The detailed operation methods for the
evaluation of scale inhibition performance can be divided into two Many environmental factors can affect the properties of feed
major categories: static and dynamic methods. water and antiscalants, scale formation, and RO operation, and
Static methods include static jar test (GB/T 16632-2008; Yang thus, influence the final inhibition efficiency (Al-Amoudi, 2010;
et al., 2017; Yu et al., 2018; Zhang et al., 2017), bubble method Antony et al., 2011; Karabelas et al., 2020; Milne et al., 2014). The
(HG/T 2024e2009), turbidity method (Abdel-Aal et al., 2015; effects of several popular external factors, such as dose, pH, tem-
Rabizadeh et al., 2017), conductivity method (Goh et al., 2018), and perature, and different coexisting substances, on the scale forma-
constant composition precipitation method (Lakshtanov et al., tion and inhibition performance of antiscalants are discussed as
2011). The most common static method is the static jar test due follows.
to easy operation. The apparent inhibition efficiency can be esti-
mated directly by the change of solute concentration; besides, the 4.2.1. Dose
changes in scale crystal morphology and the degrees of crystal Dose exerts considerable effect on inhibition performance. On
distortion determined by electron microscopy and X-ray diffraction the one hand, an inadequate dose of antiscalants may be insuffi-
can verify the effects of antiscalants also (Kim et al., 2005; Lee et al., cient to interact with scale-forming substances and cause weak
2015; Liu et al., 2012b). However, its limitation is that different inhibition efficiency (Le Gouellec and Elimelech, 2002b). On the
situations and conditions exist between the static jar test and other hand, an overdose of antiscalants may cause detrimental ef-
practical applications, making the obtained results not fully appli- fects, favoring scale formation by enhancing other effects, such as
cable to and inconsistent with many practical conditions (Boerlage bridging flocculation effect or the intermolecular interaction of
et al., 2002; Hasson et al., 1996, 2001). Considerable difference from antiscalants, in addition to the antiscaling effect (Wang et al., 2016b,
practical application conditions also exists in other static methods; Wang et al., 2017b; Yang et al., 2010b, Yang et al., 2010c). Wang
the evaluation results from these methods may be unsuitable for et al., 2016b investigated the influence of the concentration of
practical applications, but they can be used as pre-evaluation ap- alkyl ethoxy carboxylates [C12H25O(CH2CH2O)nCH2COONa] on the
proaches (Hasson et al., 2001). inhibition efficiency against CaCO3 scale. The results indicated that
Dynamic methods can provide credible results for the practical inhibition efficiency decreased slightly with the further increase of
applicability of antiscalants by simulating the conditions for prac- dose after maximum efficiency was reached, which may be due to
tical applications (Al-Roomi and Hussain, 2015, 2016; Al-Roomi the reduced adsorption of antiscalants on CaCO3 crystals caused by
12 W. Yu et al. / Water Research 183 (2020) 115985

the enhanced interaction of hydrophobic chainechain with an in- protonation to deprotonation when pH gradually increases, and the
crease in ethylene oxide content in water when overdosing (Wang interactions between antiscalants and scale-forming substances
et al., 2016b). A similar phenomenon was observed in a polymeric accordingly change, influencing inhibition performance (Ang et al.,
antiscalant, namely, carboxymethyl starch (CMS), against CaCO3, 2016; Ruiz-Agudo et al., 2016; Yang et al., 2010b, Yang et al., 2010c).
which was ascribed to the flocculation effect of the polymeric in- Rahardianto et al. (2008) compared the effect of initial pH
hibitor and enhanced intermolecular hydrogen bonding at exces- (pH0 ¼ 6.4 and pH0 ¼ 7.9), considering most feed water is under
sive CMS dose (Wang et al., 2017b). When inhibiting silica scaling, near neutral conditions, on antiscalant effectiveness in RO desalting
overdosed antiscalants also cause a detrimental effect (Demadis with high gypsum scaling propensity, but no significant variations
and Stathoulopoulou, 2006; Ketsetzi et al., 2008). Ketsetzi et al. were found. It was because the selected antiscalant (Permatreat PC-
(2008) used CATIN to inhibit silica scale, in which inhibitor dose 504) has been already at its fully ionized state at pH above 6.0
over the optimal dose may reduce inhibition activity; this result (Rahardianto et al., 2008). Ruiz-Agudo et al. (2016) reported that
may be due to the aggregation and flocculation effects of the pol- the binding capacity of a commercial copolymer antiscalant (maleic
ycationic inhibitor on the negatively charged colloidal silica parti- acid/allyl sulfonic acid copolymer with phosphonate groups) to
cles. Demadis and Stathoulopoulou (2006) observed the formation complex Ba2þ ion increased with increasing pH; this result was
of “fluffy” precipitates when the dose of PEI or PALAM was higher probably due to increased deprotonation of the acidic functional
than the optimal dose, and correspondingly, decreased inhibition groups of the additive at pH 10 relative to pH 6. Yang et al., 2010b
performance. These cationic polymers are associated with nega- investigated the pH effect of feed water with Ca2þ ions on bovine
tively charged colloidal silica particles, and thus, insoluble serum albumin (BSA) fouling by using PASP; fouling was mitigated
silicaepolymer composites are formed (Demadis and at pH 7.0 compared with that at pH 4.9 (the isoelectric point of BSA)
Stathoulopoulou, 2006). in the absence and presence of PASP due to the stronger electro-
Moreover, excessive antiscalants in RO systems may be readily static repulsion between BSA and BSA-adsorbed RO membrane at
adsorbed onto the membrane surface and increase membrane pH 7.0 that led to less accumulation of BSA on the membrane sur-
fouling caused by themselves (Ang et al., 2016; Gryta, 2012; Qin face. In the presence of PASP, BSAeCa complexes will be uneasily
et al., 2005; Yu et al., 2019). Qin et al. (2005) reported that rela- aggregated and accumulated on the membrane surface but stabi-
tive fluxes decreased with the increase of dosing Flocon 260 (a lized by PASP in water via the possible formation of water-soluble
polycarboxylic acid-type antiscalant) due to the additional organic BSAeCaePASP complexes at high pH (Yang et al., 2010b).
fouling of this antiscalant. A similar phenomenon was observed in
the study of Yu et al. (2019) on the inhibition performance of car- 4.2.3. Temperature
boxymethyl cellulose (CMC) against CaSO4. The normalized flux Temperature exhibits evident changes with the seasons, which
slightly decreased after reaching the optimal dose of 2.0 mg/L, can also affect scale formation by influencing crystal scale growth,
which may be due to the increased adsorption of CMC onto the mineral solubility, mass transfer in RO system, biofouling, and the
membrane surface as additional organic foulants (Yu et al., 2019). In interactions between antiscalants and foulants, eventually affecting
summary, considering inhibition performance and economic is- the utilization of antiscalants (Abdul Azis et al., 2001; Agashichev,
sues, optimal dose should be obtained and controlled carefully in 2005; Jawor and Hoek, 2009; Shakkthivel et al., 2004; Yang et al.,
accordance with the actual situation. 2010a). Thus, temperature effects should be considered in prac-
tical situations.
4.2.2. pH From Eq. (1), increasing temperature will increase the nucle-
Solution pH is an important environmental factor that affects ation rate when the supersaturation ratio S is kept constant
inhibition performance because solution pH can change the solu- (Boerlage et al., 2002). However, solubility can also be influenced by
bility of many types of mineral scaling and the charge properties of temperature change, affecting the supersaturation ratio (Bott,
most ionic antiscalants in water (Prihasto et al., 2009; Qin et al., 1997). The nucleation rate is critically dependent on the com-
2005; Tong et al., 2019). For example, the solubility of common bined effects of the changes in temperature and S (Boerlage et al.,
carbonate minerals strongly depends on pH. As pH decreases, car- 2002). For common inorganic scales, e.g., CaCO3 and CaSO4, the
bonate ions transform into bicarbonate ions, and the solubility of nucleation rate is typically improved with an increase in temper-
bicarbonate minerals increases accordingly (Prihasto et al., 2009; ature, accelerating scale formation (Her et al., 2000; Jawor and
Tong et al., 2019). Qin et al. (2005) explored the effect of pH on the Hoek, 2009; Matin et al., 2019; Shakkthivel et al., 2004), due to
scaling control of RO processes for raw feed from a local nickel- the minimal effects of temperature on its solubility and supersat-
plating company, which contained a high concentration of iron uration ratio (Zhang, 2011). Shakkthivel et al. (2004) conducted and
with a high risk of iron scaling; they found that a low feed pH compared scale inhibition tests of acrylonitrileeacrylic acid
resulted in a low fouling tendency of RO membrane, which can be (ANeAA) and acrylonitrileemethacrylic acid copolymers against
attributed to the scaling elimination of iron at low pH. By contrast, CaSO4 and CaCO3 scales over a temperature range of 50  Ce80  C.
the solubility of amorphous SiO2 is relatively constant at near With an increase in temperature, the inhibition efficiency of the
neutral or acidic pH but increases evidently above pH 8e9 (Milne two copolymers decreased due to the high scaling tendency at
et al., 2014; Tong et al., 2019). Bush et al. (2018) investigated the elevated temperatures. Especially for ANeAA, it acted partially as a
application of pH adjustment pretreatment to prevent silica scaling coagulant at high temperatures to form precipitates with calcium
in the membrane distillation process, and the results showed that ions and demonstrated worsened inhibition performance
alkalization and acidification are effective pretreatment strategies (Shakkthivel et al., 2004).
for preventing silica scaling if pH can be adjusted higher than 10 or Although the nucleation and crystal growth rates typically in-
lower than 5. The increasing solubility under alkaline condition is crease in RO systems, water and solute permeability and mass
due to the formation of ionic species H3SiO 2
4 and H2SiO4 , and the transfer are usually enhanced as temperature increases, which can
silica polymerization rate under acidic condition is lower compared simultaneously reduce energy consumption, solute rejection, and
with neutral and slightly alkaline solutions, reducing the formation CP. Agashichev (2005) found CP resistance decreased and trans-
of insoluble silica (Bush et al., 2018; Milne et al., 2014; Tong et al., membrane flux increased with an increase in temperature using a
2019). proposed model. Jawor and Hoek (2009) studied the effects of feed
Antiscalants with ionic functional groups can convert from water temperature on flux, rejection, and inorganic fouling by
W. Yu et al. / Water Research 183 (2020) 115985 13

gypsum scale formation for simulated brackish water. The afore- Moreover, different feed water characteristics and molecular
mentioned opposite effects of a high-temperature operation can structural features of OMs (e.g., MW) also play important roles in
enable high water recovery and low energy consumption but are the effects of coexisting OMs on mineral scaling in RO systems
accompanied by increased scaling risk (Jawor and Hoek, 2009). (Karabelas et al., 2020; Nielsen et al., 2012). Identifying the complex
Temperature also plays a significant role in microbial actions interactions involved is crucial for improving the understanding of
and biofouling. A relatively high temperature is typically beneficial scaling control in RO operation systems.
for bacterial growth and biofouling formation (Sa nchez, 2018; Yang The effects of existing multiple inorganic ions should also be
et al., 2010a). Abdul Azis et al. (2001) found a high density of considered, given that the existence of these ions may influence
biofouling organisms in an intake bay during a summer experi- scale formation or the inhibition effect of antiscalants (Rahardianto
ment; such high density was presumed to occur due to water et al., 2008; Waly et al., 2012; Wang et al., 2017a). Rahardianto et al.
quality changes in the bay, seasonally and interannually. In addition (2008) suggested that gypsum crystal growth in the presence of
to biofouling itself, inorganic scaling can also be affected directly or bicarbonate and the formation of CaCO3 scale in the gypsum-
indirectly by microbial action due to temperature change (Abdul dominant membrane system were retarded. They also reported
Azis et al., 2001; Yang et al., 2010a). Yang et al., 2010a investi- that bicarbonate and antiscalants may synergistically inhibit gyp-
gated the seasonal characteristic of seawater RO membrane fouling sum scale formation in a RO test due to their adsorption onto
and showed that biofouling and inorganic fouling were serious in gypsum crystals; however, although CaCO3 crystallization was
summer. The cell multiplication was much more abundant and apparently suppressed in the presence of gypsum, CaCO3 crystals
extracellular polymeric substances (EPS) feature intensity was may nucleate and grow at a slow rate and become observable after
stronger in summer; moreover, abundant EPS during summer were a longer period than allowed in the authors’ study (Rahardianto
believed to be the reason for the evident increase in the precipi- et al., 2008). Yu et al. (2019) found flux decline was slightly
tation of silica and calcium due to the strong capabilities of EPS to reduced after the addition of either CO2 3
3 or PO4 in a gypsum-
adsorb these coexisting ions (Yang et al., 2010a). dominant RO process without CMC as antiscalant; this result is
Thus, all the aforementioned effects of temperature must be consistent with Rahardianto et al. (2008)’s report. In the presence
considered, and the use of antiscalants must be adjusted of CMC, the addition of coexisting ions presented nearly no influ-
appropriately. ence on membrane fluxes compared with that without the addition
of coexisting ions due to the strong inhibition effect of CMC (Yu
4.2.4. Coexisting substance et al., 2019). Existing studies have shown that coexisting inor-
In practice, various pollutants coexist in treated water. The ganic anions may exert minimal effects on the inhibition perfor-
scaling propensity of sparingly soluble salts may be influenced by mance of antiscalants but do influence the scaling process
the existence of other pollutants, such as OMs and inorganic ions, (Rahardianto et al., 2008; Waly et al., 2012; Yu et al., 2019).
which further affect the utilization of antiscalants (Matin et al., In addition to inorganic anions, metallic ion impurities, notably
2019). OMs are ubiquitous in all natural water sources, with pro- zinc ion, can also influence scale precipitation, which can hinder
teins, polysaccharides, and humics as the three major types of CaCO3 precipitation and change crystal morphology (Aziz and
widespread organic foulants (Wang et al., 2016a). BSA, sodium Kasongo, 2019; Bystrianský et al., 2016; Ghizellaoui and Euvrard,
alginate (SA), and humic acid (HA) are generally used as model 2008; Lisitsin et al., 2005). Aziz and Kasongo (2019) compared
organic foulants to investigate the effects of OMs on scaling, as the inhibition performance of a commercial antiscalant (Vitec 3000,
representative of proteins, polysaccharides, and humics, respec- containing phosphonic acid H2O3Pþ) and zinc ions; the fluxes after
tively (Benecke et al., 2018; Wang et al., 2016a). Wang et al., 2016a a period of nearly 5 months were 32.78 L/(m2$h) and 30.80 L/
found that the three pregenerated organic fouling layers, BSA, HA, (m2$h) for commercial and zinc ion antiscalants, which were higher
and SA, all reduced the membrane flux compared with virgin than the untreated flux, i.e., 25.56 L/(m2$h). Lisitsin et al. (2005)
membrane in a gypsum-dominant NF process; HA and SA still acted reported a zinc ion concentration of 2.0 mg/L was able to achieve
as nuclei for crystallization, leading to higher surface crystallization an evident suppression effect on the bulk precipitation of CaCO3
of gypsum scaling and more serious flux decline than BSA-fouled and on membrane scaling. Waly et al. (2012) found that in addition
membrane (Wang et al., 2016a). Benecke et al. (2018) reported to zinc ions, magnesium ions significantly increased the induction
similar results that scaling from gypsum surface crystallization was time of CaCO3 crystallization, which may be due to the effects of
enhanced in the presence of SA and HA. However, the antiscaling magnesium ions on nucleation and crystal growth through
effects of various OMs in membrane processes were also observed complexation and growth site blocking. In contrast with the inhi-
(Karabelas et al., 2017; Le Gouellec and Elimelech, 2002a; Lee et al., bition effects of zinc ions and magnesium ions mentioned earlier,
2009). Several studies have reported that the presence of HA Bystrianský et al. (2016) reported the presence of iron ions accel-
slowed down RO/NF membrane flux decline caused by mineral erated crystallization of calcium sulfate and led to steeper flux
scaling, most likely via the complexation effect and adsorption on decline in static jar tests and RO tests, because of the formation of a
active sites of the crystal surface (Le Gouellec and Elimelech, 2002a; solid scalant layer with iron oxides. Metal ions with high scaling
Lee et al., 2009). Karabelas et al. (2017) investigated the effects of SA potential, such as iron ions, should be noted, because they will
during the early stage of membrane scaling, and its inhibition effect evidently contribute to aggravated membrane fouling.
was proven in terms of the reduced deposit density on membranes, In summary, the effects of various coexisting substances on
crystallization rate, and crystal shape modification. SA was sup- scaling have no unified law to follow based on these fragmentary
posed to inhibit the crystallization process likely due to calcium reports; the presence of different impurities may have varying in-
binding with carboxylates of SA (Karabelas et al., 2017). The con- fluences on scale formation by accelerating or inhibiting crystalli-
centrations of OMs in these studies with positive effects on scale zation processes, which highly depends on the specific
inhibition (Karabelas et al., 2017; Le Gouellec and Elimelech, 2002a; characteristics of different systems due to complicated physico-
Lee et al., 2009) were below or equal to 10.0 mg/L, and those in the chemical interactions involved among dominant scale-forming
aforementioned opposite reports (Benecke et al., 2018; Wang et al., substances, antiscalants, membranes, and coexisting substances
2016a) (100.0 mg/L or 3.0e12.0 mg carbon/L) were higher than in solutions (Karabelas et al., 2020; Nielsen et al., 2012). In addition,
10.0 mg/L. This condition may partly explain the difference in the these interactions are strongly influenced by the environmental
apparent contradictory effects of OMs on mineral scaling. factors mentioned earlier, affecting the inhibition performance of
14 W. Yu et al. / Water Research 183 (2020) 115985

antiscalants and the scaling behaviors in actual operations. scale under different operation conditions and membrane types.
Furthermore, comprehensive research on the influence of different Inhibition performance increased with an increase in antiscalant
environmental factors should be conducted to optimize the oper- dose, the membrane area covered by scale decreased, and the flux
ation conditions and understand the inhibition mechanisms for decline caused by scaling was eliminated (Rahardianto et al., 2006).
guiding the utilization of antiscalants in practical applications. In their other study (Shih et al., 2005), similar conclusions were
obtained that the amounts of crystals on the membrane surface
5. Applications of antiscalants in RO desalination systems decreased as antiscalant dose increased, and the size and
morphology of surface crystals were changed significantly. Yang
Various antiscalants have been developed and applied in water et al., 2010b tested the inhibition efficiency of an organophos-
treatment industries, such as boiler water, cooling water, thermal phorus compound (LB-0100) in the control of BSA fouling using a
desalination and membrane industries, to mitigate scaling bench-scale crossflow RO system. The results indicated that BSA
(Chaussemier et al., 2015; Darton, 2000; Hasson et al., 2011; Liu fouling can be considerably reduced by properly dosing anti-
et al., 2019; Nowack, 2003). Specific knowledge regarding the scalants LB-0100. Overdosing causes aggravated RO membrane
performance of different antiscalants in RO processes is beneficial fouling, which may result from the reduced water solubility of the
for understanding their mechanisms and optimizing their appli- formed antiscalant complexes at high antiscalant concentration
cations. However, most of reports and literatures related to those (Yang et al., 2010b).
commercial antiscalants were mainly focused on their application In comparison with previous pH control in scale prevention, the
performance not antiscalants themselves, and their detailed in- application of phosphorus antiscalants evidently reduces and even
formation, including the composition, purity and concentrations, eliminates the use of acid and alleviates the subsequent corrosion
are not provided by manufacturers due to the commercial values, (Bonne  et al., 2000). Bonne
 et al. (2000) used a commercial orga-
making it difficult to thoroughly clarify the role of those commer- nophosphonate antiscalant (Permatreat 191) in a RO pilot plant to
cial antiscalants. Here, the reported performance of various anti- reduce the demand for acid. When a small amount of acid was used
scalants applied to different RO processes is primarily summarized with an antiscalant dose of 1.4 mg/L for the safe control of the
in Table 3 and elucidated as follows. desired Langelier saturation index (LSI) due to feed water fluctua-
tions, recoveries up to 87% can be obtained without scaling, and no
5.1. Phosphorus antiscalants membrane cleaning is necessary within a period of 1 year (Bonne 
et al., 2000). However, Qin et al. (2005) found that the addition of
The fact that phosphates can prevent the precipitation of metal Permatreat 191 even enhanced membrane fouling by using spent
ion-based minerals by forming a soluble complex with metal ions rinse water from a nickel-plating operation after pH adjusted to
or/and distortion of crystal lattice has been discovered for more 4.2e6.8, and another typical polycarboxylic acid-type antiscalant
than 180 years (Rashchi and Finch, 2000). Phosphorus antiscalants, (Flocon 260) did not exhibit an inhibition effect on membrane
including phosphates and phosphonates, can inhibit scale forma- fouling also, which may be due to the complex ingredients of this
tion under low concentrations and have been widely applied in actual wastewater.
various fields (Nowack, 2003; Rashchi and Finch, 2000). They are In addition, some commercial antiscalants consist of several
the most popular antiscalants due to their high effectiveness and active ingredients for high inhibition performance due to the syn-
are the major commercial antiscalants applied to RO systems (Al- ergistic effect, such as combining phosphorus antiscalants with
Rammah, 2000; Bonne  et al., 2000; Rahardianto et al., 2006; synthetic polymeric antiscalants (Ang et al., 2016; He et al., 2009;
Semiat et al., 2003; Shih et al., 2005). SHMP is a commonly applied Pramanik et al., 2017; Semiat et al., 2003). Semiat et al. (2003)
phosphate due to its complexation ability and low cost (Butt et al., assessed the effectiveness of four commercial antiscalants against
1997; Hasson et al., 1998; Rashchi and Finch, 2000). Hasson et al. silica scale on the basis of changes in permeability decline and
(1998) conducted preliminary RO tests to compare the degree of membrane surface supersaturation level; the four antiscalants
scale suppression against CaCO3 scale provided by Calgon (SHMP) were two different blends of phosphonates and polymers (Per-
and Flocon 100 (acrylic polyelectrolyte, 35% solution). Flux decline matreat 510 and Hypersperse SI 300UL), a multi-polymer (Acumer
data indicated that 2.0 mg/L of Calgon can effectively prevent 5000), and an anionic polyelectrolyte (Aquafeed EX-105). All four
permeability decline. For Flocon 100, a dose of 3.3 mg/L cannot antiscalants increased the threshold limit of the scaling onset from
completely suppress scale formation, but a dose of 10.0 mg/L is a membrane surface supersaturation level of 1.7e1.9 in the absence
necessary for full-scale suppression over a test period of approxi- of antiscalants to a level of 2.1e2.3 in the presence of antiscalants.
mately 9 h (Hasson et al., 1998). Permatreat 510 exhibited the best inhibition performance among
However, phosphates occasionally easily hydrolyze into ortho- the four antiscalants due to certain synergistic and complementary
phosphate ions, which may lead to the formation of calcium effects of low-molecular and macromolecular inhibitors (Semiat
phosphate and increase the potential of biofouling; thus, the use of et al., 2003). Butt et al. (1997) used a synthetic brine to evaluate
phosphonates is more common to avoid the inactivity of the efficacy of three inhibitors, namely, SHMP, inhibitor-A (acrylic
phosphate-based antiscalants because of their hydrolyzation (Al- polymer), and inhibitor-B [polyacrylate (PA) and phosphonate].
Roomi and Hussain, 2015; Andrews et al., 2008; Butt et al., 1997; When treated with the inhibitors, the scale-forming constituents
Li et al., 2017; Nowack, 2003). Al-Rammah (2000) conducted (Ca2þ, Sr2þ, and total alkalinity) in solution can remain unchanged
evaluation trials of several acid-free antiscalants with phospho- for 91 h (SHMP), 50 h (inhibitor-A), and 168 h (inhibitor-B);
nates to replace the combination of sulfuric acid and SHMP in a inhibitor-B, a composite antiscalant, achieved the best inhibition
brackish water RO plant. The results showed that a phosphino- performance because of the synergistic effect (Butt et al., 1997).
carboxylate-based antiscalant (M-chemical) completely inhibited Field and laboratory results obtained by Drak et al. (2000) indicated
membrane scaling and provided superior performance to the that all four tested antiscalants enabled to prevent scaling at a
combination of SHMP and sulfuric acid due to its higher stability water recovery level below 88% and LSI levels exceeding 2.0,
and superior dispersion of silt, clay, and metal oxides (Al-Rammah, showing good inhibition performance; however, three composite
2000). Rahardianto et al. (2006) used a commercial formulation, antiscalants, all consisting of phosphonate and PAA, still presented
Vitec 2000, with a phosphino-carboxylic acid polymer, as the active different inhibition effects, indicating the importance of formu-
ingredient and tested its inhibition performance against gypsum lating different ingredients in composite antiscalants.
W. Yu et al. / Water Research 183 (2020) 115985 15

Table 3
Antiscaling performance of various antiscalants in RO process.

Antiscalants Active ingredients Scale-forming Dose (mg/ Performance Mechanism Reference


substance L)

Calgon Sodium hexametaphosphate Ca2þ, HCO


3 1.2, 2.0, 2.0 mg/L could effectively inhibit scale Retard nucleation and crystal Hasson et al.
3.6, 5.5 formation. growth processes. (1998)
Flocon 100 Polyacrylic acid 2.0, 3.3, 10.0 mg/L ensured full scale suppression over
10.0 9 h.
C-chemical Organic phosphonate and Ca2þ, Mg2þ, Brown particulate deposits. Al-Rammah
acrylic polymer HCO 2-
3 , SO4 (2000)
M-chemical Phosphino-carboxylate Clean membrane. Superior dispersion.
H-chemical Phosphonate Brownish black deposits.
Vitec 2000 A phosphino- carboxylic acid Ca2þ, Mg2þ, 0.5e5.0 As dose 3.0 mg/L, membrane surface appeared Retardation of gypsum surface Rahardianto
polymer SO2-
4 to be free of gypsum crystals. scale. et al. (2006)
Vitec 2000 A phosphino- carboxylic acid Ca2þ, Mg2þ, 0.5e5.0 Flux decline was gradually decreased as Crystal distortion. Shih et al.
polymer SO2-
4 antiscalant dose increase (From 30.1% without (2005)
antiscalant to 3.5% with 3.0 mg/L antiscalant).
Permatreat Organophosphonate Ca2þ, Ba2þ, 1.4 Recovery up to 87%, no scaling (in combination Bonne et al.
191 SO2-
4 with addition of H2SO4). (2000)
Permatreat Organophosphonate Ca2þ, Mg2þ, Enhanced fouling. As additional organic fouling. Qin et al.
191 Fe, SO2-
4 (2005)
Flocon 260 Polycarboxylic acid Not inhibit fouling of the membrane.
Permatreat Blend of polymers and Ca2þ, Mg2þ, From a membrane surface supersaturation level Increase of the threshold limit for Semiat et al.
510 phosphonates SiO2 of 1.7e1.9 on the membrane surface in the the onset of scaling. (2003)
Hypersperse Blend of polymers and 3.0e15.0 absence of an antiscalant, to the level of 2.1e2.3,
SI 300 UL phosphonates in the presence of the antiscalant.
Acumer Multi-polymer
5000
Aquafeed Anionic polyelectrolyte
EX-105
Antiscalant Phosphonate Ca2þ, Mg2þ, 4.2 All four tested antiscalants enabled a water Suppress scale formation. Drak et al.
A1 HCO 2-
3 , SO4 , recovery level of at least 88% at LSI levels (2000)
Antiscalant Phosphonate þ polyacrylic SiO2-
3 4.0 exceeding 2.0.
B1 acid
Antiscalant Phosphonate þ polyacrylic 4.5
C1 acid
Antiscalant Phosphonate þ polyacrylic 3.9
D1 acid
SHMP Sodium hexametaphosphate Ca2þ, Mg2þ, 1.0 Keep the scale-forming constituents in solution Chelation and threshold effect. Butt et al.
Sr2þ, HCO
3, for 91h. (1997)
Inhibitor-A Acrylic polymer SO2-
4 , Keep the scale-forming constituents in solution Chelation and dispersion effect.
for 50 h.
Inhibitor-B Polyacrylate þ phosphonate Keep the scale-forming constituents in solution Chelation, dispersion and
for 168h. threshold effect.
Antiscalant Polyacrylate (MW ¼ 1600 Da) Pre-treated 2.0e4.0 Scaling was frequently found during operation Inappropriate dose; dispersive Boerlage
A2 River Rhine with antiscalant A2 but scale layer became less property. et al. (2002)
water cohesive.
Antiscalant Polyacrylate (MW ¼ 3500 Da) containing 2.0 No scaling was detected during operation with Dispersive ability.
B2 Ba2þ, SO2-
4 antiscalant B2 but enhanced biofouling
occurred.
Polycarboxylic acid-based Ca2þ, SO2-
4 0.25, 1.0, Increased antiscalants concentrations prolonged Crystal distortion. Shmulevsky
polymer 2.0 the induction period and led to rougher formed et al. (2017)
crystals.
Flocon 100 Polyacrylic acid Tap water 5.0, 10.0 Sludge-like deposits formed on membrane. Crystal distortion. Tzotzi et al.
Flocon 135 containing Slightly scaled membrane. Crystal distortion. (2007)
Permatreat Ca2þ, Mg2þ, The flux curve is almost constant, sludge-like Crystal distortion.
391 HCO 2-
3 , SO4 , deposits on membrane.
Hypersperse SiO2 The flux curve is almost constant.
AF200
Goodrite Polyacrylic acid/sulfonic acid/ The flux curve is almost constant. Delay bulk precipitation.
K798 sulfonated styrene
Goodrite Polyacrylic acid The flux curve shows a relatively sharp decline, Crystal distortion.
K702 distorted calcite crystals with rough rounded
edges on membrane.
PASP BSA, Ca2þ 2.0e50.0 The inhibition efficiency increased with PASP Stable complex BSAeCaePASP Yang et al.,
concentration from 2.0 to 10.0 mg/L, and via Ca2þ bridging. 2010b
reached 96% at 10.0 mg/L PASP.
LB-0100 An organophosphorus 5.0, 50.0 65% inhibition efficiency with 5.0 mg/L dose; Increase water solubility of the
compound aggravated fouling with 50.0 mg/L dose. formed antiscalant complexes.
PASP HA, Ca2þ 2.0e50.0 The inhibition efficiency increased with PASP Stable complex HAeCaePASP via Yang et al.,
concentration from 2.0 to 10.0 mg/L, and Ca2þ bridging. 2010c
reached 91% efficiency at 10.0 mg/L PASP.
Na2-EDTA Disodium 336.21 The inhibition efficiency for Na2-EDTA was 54%. Remove free and HA-complexed
ethylenediaminetetraacetate (1.0 mmol/ calcium ions.
L)
PASP 1.0
(continued on next page)
16 W. Yu et al. / Water Research 183 (2020) 115985

Table 3 (continued )

Antiscalants Active ingredients Scale-forming Dose (mg/ Performance Mechanism Reference


substance L)

PASP-SEA- Containing carboxylic acid Ca2þ, Mg2þ, The antiscalants led to reduced scale formation, Threshold inhibition; crystal Pramanik
ASP groups and sulfonic acid SO2-
4 , SiO2 with antiscaling effect in the order of PASP-SEA- dispersion; crystal modification. et al. (2017)
groups ASP > PASP > commercial antiscalant.
A commercial phosphonic acid and
antiscalant polycarboxylic acid
CMC Ca2þ, SO2-
4 1.0e30.0 Flux decline was greatly decreased as Inhibit bulk crystallization; Yu et al.
dose  2.0 mg/L; aggravated fouling as crystal distortion. (2019)
dose > 10.0 mg/L.
PAA 2.0e6.0 Flux decline was greatly decreased as
dose  4.0 mg/L.
ATMP 2.0e6.0 Flux decline was greatly decreased as dose
4.0 mg/L.
2þ 2þ
Zinc ions Ca , Mg , 0.075, The flux decline was lower compared to that Inducing increases in nucleation Aziz and

SO2-
4 , HCO3 0.120 without antiscalant. times. Kasongo
Vitec 3000 Phosphonic acid (H2O3Pþ) The flux decline was reduced to just 5%. Retardation of the onset of (2019)
surface crystallization or
extremely slow growth of salt
crystals on the membrane
surface.
Zinc ions Ca2þ, Mg2þ, 2.0, 5.0, The presence of 2.0 mg/L zinc ion was able to Hinder scale formation. Lisitsin et al.

SO2-
4 , HCO3 10.0 prevent bulk precipitation and reduce (2005)
permeability decline.

Although various phosphorus antiscalants typically exhibit good importance of the molecular structures of polymeric antiscalants,
inhibition performance in different fields, the use of this type of such as MW, in their scale inhibition as previously discussed
antiscalants should be restricted to reduce phosphorus emissions (Boerlage et al., 2002).
and the risk of eutrophication of surface waters (Andrews et al., Copolymers with other functional groups apart from carboxylic
2008; Boels and Witkamp, 2011; Li et al., 2017). groups have been developed and applied to improve the inhibition
performance of synthetic polymeric antiscalants (Dietzsch et al.,
5.2. Synthetic polymeric antiscalants 2013; Shakkthivel and Vasudevan, 2006; Wang et al., 2009).
Tzotzi et al. (2007) assessed the effectiveness of various commercial
The release of phosphorus from phosphorus antiscalants will antiscalants in inhibiting calcium carbonate scale on RO and NF
cause unfavorable environmental effects; therefore, the develop- membranes. Most of these antiscalants exhibited considerable in-
ment of new phosphorus-free antiscalants has elicited considerable hibition performance against calcium carbonate on membrane
attention (Andrews et al., 2008; Boels and Witkamp, 2011; Li et al., surface by changing the morphology of crystals and reducing the
2017). Various synthetic polymeric antiscalants with different amount of deposits. The Goodrite K798 additives (PAA/sulfonic
antiscaling functional groups have been prepared based on the acid/sulfonated styrene) demonstrated significantly better perfor-
homopolymerization or copolymerization of acrylic acid (AA), mance compared with K702 (PAA), which may be due to the syn-
maleic acid (MA), and other functional monomers (Ghani and Al- ergistic effect of various functional groups (Tzotzi et al., 2007).
Deffeeri, 2010). They can effectively prevent scale formation due In summary, polycarboxylic acid antiscalants are important
to chelation, dispersion, and crystal lattice distortion, and thus, are types of synthetic polymeric antiscalants because carboxylic acid
used as important ingredients of commercial antiscalants (Ang groups play an important role in scale inhibition (Yang et al., 2017;
et al., 2016; Ghani and Al-Deffeeri, 2010; He et al., 2009). Zhang et al., 2016c). The introduction of multiple functional groups
Polycarboxylic acid is a major class of synthetic polymeric improves the inhibition performance of this type of antiscalants
antiscalants, including PAA, polymethacrylic acid, polymaleic acid, due to the synergistic inhibition effect. However, the
and their bi-, ter-, and multi-copolymers that contain carboxylic structureeactivity relationship of polymeric antiscalants cannot be
groups (Al-Hamzah and Fellows, 2015; Ketrane et al., 2009). PAA is established clearly, and the effects of various structural factors,
the most common active ingredient in commercial synthetic including MW; the contents, types, and distributions of various
polymer antiscalants (Al-Hamzah and Fellows, 2015; Ketrane et al., scaling functional groups; and the sequence structures, particularly
2009; Shmulevsky et al., 2017). Shmulevsky et al. (2017) revealed for multi-copolymer inhibitors, on inhibition performance are still
that increased concentrations of a polycarboxylic acid-based anti- not comprehensively understood. Insufficient knowledge about
scalant prolonged the induction period obtained from permeate polymeric antiscalants in scaling control at the molecular levels
flux data in inhibition of calcium sulfate scaling on RO membranes largely limits their practical applications.
and led to rough formed crystals due to the crystal distortion effect.
Boerlage et al. (2002) reported the application of two PA polymers 5.3. Environment-friendly antiscalants
with different MWs in RO pilot plants. Scaling frequently occurred
during the 6 months of operation with antiscalant A2 (PA, Synthetic polymeric antiscalants are not easily degradable,
MW ¼ 1600 Da), whereas no scaling was detected during the which may cause a difficulty in the subsequent concentrate
following 6 months of operation with antiscalant B2 (PA, disposal and a negative environmental effect due to their emissions
MW ¼ 3500 Da). The different inhibition performance of the two (Hasson et al., 2011). Thus, environment-friendly antiscalants have
PAs was partially due to their different MWs, in addition to changes become a popular topic because of their phosphorus-free, biode-
in environmental conditions (Boerlage et al., 2002). Antiscalant B2 gradable, and nonharmful properties. The inhibition efficiency of
with a relatively high MW exhibited an enhanced dispersive PASP, PESA, and their various derivatives, as the most promising
property, resulting in high performance. This finding indicated the green antiscalants, has received widespread attention and been
W. Yu et al. / Water Research 183 (2020) 115985 17

reported in many studies (Chaussemier et al., 2015; Chen et al., effective inhibition performance, compared with the conventional
2015a; Hasson et al., 2011; Liu et al., 2012a; Pramanik et al., 2017). pretreatment methods mentioned earlier (Darton, 2000; Li et al.,
Yang et al., 2010b,Yang et al., 2010c reported the control of OM 2011; Tong et al., 2019). However, additional knowledge about
fouling (e.g., BSA fouling and HA fouling) with feed water con- polymeric antiscalants in scaling control, particularly at the mo-
taining Ca2þ ions by PASP and several commercial compounds [LB- lecular levels, is urgently required for optimized applications in
0100 mentioned earlier for BSA fouling and disodium ethylenedi- their practical operations.
aminetetraacetate (Na2-EDTA) for HA fouling] in RO systems. OM
fouling could be reduced by using a suitable dose of PASP, but 6. Existing problems and perspectives
overdosing had a detrimental effect. The presence of Ca2þ improved
the inhibition performance of PASP against OM fouling, which may Various antiscalants, their applications to RO membrane scaling
be due to the formation of a stable water-soluble complex control, and their advantages and disadvantages have been
OMeCaePASP through Ca2þ bridging (Yang et al., 2010b, Yang et al., reviewed and discussed in this paper. Although antiscalants play a
2010c). Compared with the two commercial compounds, PASP significant role in preventing membrane scaling, many challenges,
exhibited better inhibition effect, which might be due to its distinct practical problems, and weakness of existing antiscalants, such as
chemical structures. Gwak and Hong (2017) used an antiscalant- enhanced bacterial growth, appropriate dosing control, and final
blended draw solution with a mixture of NaCl and poly(aspartic concentrate disposal, exist in their applications (Greenlee et al.,
acid sodium salt) (PASP-Na) in forward osmosis experiments to 2009b, Greenlee et al., 2010a; Greenlee et al., 2010b, 2011; Joo
examine the scaling control effect of the novel strategy. Gypsum and Tansel, 2015; Malaeb and Ayoub, 2011; McCool et al., 2012,
fouling tests demonstrated that the addition of PASP-Na to the draw 2013; Pe rez-Gonzalez et al., 2012; Subramani and Jacangelo, 2014;
solution can effectively control membrane scaling due to the inhi- Sweity et al., 2013, 2014; 2015; Turek et al., 2017; Vrouwenvelder
bition effect of reversely flowed antiscalant from the blended draw et al., 2000). To improve the effectiveness of antiscalants in RO
solution (Gwak and Hong, 2017). membrane scaling control, ongoing research can focus on the
To further improve the inhibition performance of PASP, various development of novel, high-performance, and environment-
attempts are ongoing, and one of the most common methods is to friendly antiscalants on the basis of an in-depth study of the inhi-
introduce functional groups via chemical modification (Chen et al., bition mechanisms and well-established structureeactivity re-
2015a; Gao et al., 2015; Pramanik et al., 2017; Zhang et al., 2016c, lationships obtained from a comprehensive understanding of the
2017). Pramanik et al. (2017) synthesized a PASP derivative (PASP- molecular structures of antiscalants and the involved interactions
SEA-ASP) by introducing sulfonic acid and carboxylic acid to the among antiscalants, scale-forming substances, membranes, and
side chain of PASP and tested its scale inhibition performance on other coexisting matters in solutions (Karabelas et al., 2020; Liu
the desalination of synthetic brackish water using a polyamide RO et al., 2019; Matin et al., 2019).
membrane. The results showed that compared with a commercial
phosphorus antiscalant, PASP-SEA-ASP and PASP achieved higher 6.1. Existing problems
improvement in water recovery. PASP-SEA-ASP was more effective
than PASP due to the enhanced chelation ability from the intro- With regard to the practical applications of antiscalants to RO
duced sulfonic acid groups (Pramanik et al., 2017). membrane processes, several problems that limit and influence the
Apart from PASP, the scale inhibition efficiency of other green utilization of antiscalants, such as enhanced bacterial growth,
antiscalants derived from natural polymers (Chaussemier et al., appropriate dosing control, and concentrate disposal, should be
2015; Hasson et al., 2011), such as oxidized starch (Guo et al., emphasized (Greenlee et al., 2009b, Greenlee et al., 2010a;
2013), CMS (Wang et al., 2017b), starch-graft-polyacrylic acid (St- Greenlee et al., 2010b, 2011; Joo and Tansel, 2015; Malaeb and
g-PAA) (Yu et al., 2018), CMC (Kavitskaya et al., 2000; Yu et al., Ayoub, 2011; McCool et al., 2012, 2013; Pe rez-Gonzalez et al.,
2019), and CMI (Akin et al., 2008; Boels and Witkamp, 2011), has 2012; Subramani and Jacangelo, 2014; Sweity et al., 2013, 2014;
also been investigated. The inhibition performance of these anti- 2015; Turek et al., 2017; Vrouwenvelder et al., 2000).
scalants has been evaluated using static jar tests and RO tests, and
the results are comparable with those of commercial antiscalants. 6.1.1. Enhanced biofouling
Yu et al. (2019) prepared a series of CMC with different degrees of Antiscalants can effectively inhibit scale formation through the
substitution (DS) and investigated its role in mitigating gypsum interactions between functional groups and scale-forming sub-
scaling on RO membranes. Similar to reported results (Ang et al., stances. However, antiscalants may also alter the
2016; Gryta, 2012; Yang et al., 2010b, Yang et al., 2010c), an physicalechemical surface properties of membranes via adsorp-
appropriate dose was necessary for CMC to achieve remarkable tion, particularly at overdoses, which will enhance membrane
inhibition performance. The optimal dose decreased with an in- biofouling (Sweity et al., 2013, 2014, 2015). In particular, membrane
crease in DS but remained nearly unchanged when DS was more surface charge, roughness, and hydrophobicity are important fac-
than a certain value; this result indicated that a suitable DS was tors that influence membrane organic fouling and biofilm forma-
important to achieve good inhibition efficiency and minimize the tion; these properties of membrane surface can be affected by the
production cost of antiscalants (Yu et al., 2019). Moreover, the in- adsorption of antiscalants, promoting membrane organic fouling
hibition performance of CMC was as good as those of PAA and and attachment of bacterial cells onto the surface (Ang et al., 2016;
amino trimethylene phosphonic acid (ATMP), implying that natural Sweity et al., 2013, 2014, 2015). Common commercial phosphorus-
polymer derivatives have remarkable potential as efficient based antiscalants are also a nutritional source for bacterial growth
environment-friendly antiscalants (Yu et al., 2019). (Sweity et al., 2013, 2015; Vrouwenvelder et al., 2010).
In summary, the applications of antiscalants to membrane Sweity et al. (2013) tested the contribution of two common
scaling control, particularly field operation data, are still limited, antiscalants, i.e., polyphosphonate- and PA-based ones, to mem-
and widespread use of scale inhibitors in RO systems awaits brane biofouling in the RO desalination process, and the effects of
considerable laboratory and field operation experiences. Anti- antiscalants on membrane surface properties and bacterial adhe-
scalants, especially environment-friendly inhibitors, applied to RO sion were studied. The results indicated that the addition of both
membrane processes, exhibits the advantages of convenient oper- types of antiscalants increased biofilm formation rate on mem-
ation, low device cost, minimal equipment damage, and highly brane surface but followed different mechanisms. The PA-based
18 W. Yu et al. / Water Research 183 (2020) 115985

antiscalant may enhance initial cell attachment and deposition by online monitoring of antiscalant concentration using simple, online
altering the polyamide surface more hydrophobic, promoting bio- equipment (Kelle Zeiher et al., 2003). The results showed that the
film formation; the polyphosphonate-based antiscalant was use of fluorescent technology made the automated monitoring and
considered as an additional source of phosphorous to enhance control of antiscalant dosing effective (Kelle Zeiher et al., 2003).
biofilm formation rate (Sweity et al., 2013). Although fluorescent molecule technology has been successfully
Accordingly, bacterial growth can be enhanced not only by applied to cooling water systems, the compatibility of added fluo-
phosphorous-based antiscalants but also by synthetic polymeric rescent molecules with the membrane surface must be considered
 et al., 2000; Vrouwenvelder
antiscalants (Ashfaq et al., 2019; Bonne for a membrane system. Apart from combining fluorescent mole-
et al., 2000). Vrouwenvelder et al. (2000) tested the biofouling cules with antiscalants, the synthesis of fluorescent antiscalant
potential of 14 antiscalants with an easily assimilable organic car- through the copolymerization of a monomer with a fluorescent
bon test and a biomass production potential test. The used chem- chromophore and other monomers or modifying polymers using
icals exerted different effects on promoting biological growth, with fluorescent groups is another useful method for introducing fluo-
the PAA-based antiscalant A achieving the greatest promoting rescence characteristics (Feng et al., 2014; Wang et al., 2014). Feng
growth potential. Antiscalant A (2.5 mg/L) caused a significant in- et al. (2014) introduced carbazole groups into PASP to obtain
crease in the normalized pressure drop, approximately 60% within fluorescent poly(aspartic acid) (FPASP). The results showed that the
11 days (Vrouwenvelder et al., 2000). Bonne  et al. (2000) used a relationship between the concentration of FPASP and the fluores-
combination of acid and antiscalant Flocon 100 to prevent the cence intensity was linear, which provided the possibility of auto-
formation of CaCO3 and BaSO4 scales in a RO plant. Although the dosing with FPASP (Feng et al., 2014). Appropriate dosing control
combination was successful for scale inhibition with acid and reduces the risk of underdosing and overdosing, making the use of
antiscalant dosing, the frequency of membrane cleaning was once antiscalants in RO systems effective. Apart from fluorescent tracer
every 4e6 weeks. Flocon 100 may be related to frequent cleaning technology, antiscalant detection requires further research to
because it was beneficial for biological growth and biofouling as a explore more effective dosing control methods.
nutrient (Bonne  et al., 2000).
These widely used antiscalants mostly consist of organic com- 6.1.3. Concentrate disposal
pounds; combined with the influence on membrane surface The generated concentrate from RO systems requires further
properties, the problem of enhanced biofouling should be focused reasonable treatment (Joo and Tansel, 2015; Subramani and
on and mitigated. As mentioned earlier, the use of phosphorus Jacangelo, 2014), and controlled precipitation is a promising
antiscalants should be reduced to avoid their effect on enhancing approach (Greenlee et al., 2010b). Nevertheless, the presence of
bacterial growth in addition to decreasing the risk of water eutro- antiscalants in the concentrate may adversely affect concentrate
phication and algal bloom. The development of novel environment- disposal due to the anti-precipitation effect of antiscalants
friendly antiscalants with high purification and enhanced inhibi- (Greenlee et al., 2010a, Greenlee et al., 2010b; 2011; McCool et al.,
tion performance that require low doses is desirable for mitigating 2012). Greenlee et al., 2010a performed precipitation experiments
enhanced biofouling. In addition, the combination of antibacterial using a synthetic RO concentrate to investigate the effects of anti-
units with antiscalants can be optimized to control membrane scalant type, antiscalant concentration, and precipitation pH on
biofouling and scaling. CaCO3 precipitation and filtration processes. The results indicated
that adding antiscalants can change the shape of precipitate par-
6.1.2. Dosing control ticles and decrease particle size, increasing MF flux decline in the
As mentioned earlier, dose plays an important role in the inhi- succeeding separation step. The mass of precipitated particles was
bition performance of antiscalants, and inappropriate dosing will also reduced due to the presence of antiscalants during precipita-
not only lead to increased membrane fouling but will also reduce tion (Greenlee et al., 2010a). Greenlee et al., 2010b also tried several
inhibition efficiency (Kelle Zeiher et al., 2003). In addition to synthetic concentrates in the presence of different compositions,
various methods for identifying the incipient scaling as mentioned such as Mg2þ and SO2 4 , in which antiscalant dosing increased
in Section 2.1.2 (Karabelas et al., 2020; van de Lisdonk et al., 2001), magnesium and sulfate precipitation and decreased calcium pre-
antiscalant monitoring is essential for the appropriate dosing of cipitation. The adsorption of antiscalants onto the formed CaCO3
antiscalants also. Conventional monitoring methods include crystals was speculated to inhibit crystal growth and allow
colorimetric, spectrometric, turbidimetric, and potentiometric increased magnesium and sulfate to be incorporated into the
techniques and the use of fluorescent tracers (Andrews et al., 2008; formed CaCO3 (Greenlee et al., 2010b). The same group applied the
Darton, 2000; Feng et al., 2014; Kelle Zeiher et al., 2003; Wang et al., advanced oxidation process (AOP) of ozonation and hydrogen
2014; Yuchi et al., 2007). In contrast with acid dosing or softening, peroxide to the degradation of phosphonate antiscalants to elimi-
which can be easily monitored using pH or ion-selective electrodes, nate the detrimental effect of antiscalants in the concentrate pre-
most of these monitoring methods are troublesome and require cipitation process (Greenlee et al., 2011). The results showed that
measurements using large-scale instruments (Kelle Zeiher et al., AOP removed the effects of antiscalants on the mass, size, and
2003). Yuchi et al. (2007) summarized the conventional methods morphology of precipitation particles and improved MF perfor-
for monitoring the concentration of a widely used antiscalant PA, mance in the succeeding separation stage in some cases (Greenlee
including (1) monitoring a spectral change based on the interaction et al., 2011).
of PA with certain substances (e.g., metalereagent complex; met- Xu et al. (2019) used ozonation treatment to remove 2-
achromic or fluorochromic dye); (2) monitoring the concentration phosphonobutane-1,2,4-tricarboxylic acid (PBTCA), and the re-
of an inert tracer added in proportion to PA via potentiometry or sults showed that the antiprecipitation effect of the antiscalant was
spectrophotometry (e.g., Liþ, Kþ, Br, I, transition metal ion, dye); gradually reduced with increasing ozonation time. Boels et al.
(3) monitoring the intensity of fluorescence emitted from a tag (2012) investigated the adsorption and desorption of ATMP onto
covalently bound to PA; and (4) monitoring turbidity change on the granular ferric hydroxide (GFH) from RO membrane concentrate.
interaction of PA with a cationic surfactant. Several of these The results indicated that the presence of calcium increased the
methods are inapplicable to rapid, on-site, and online evaluation. surface coverage of phosphonates and substantially increased the
Fluorescent molecule monitoring based on combining a fluo- adsorption rate. GFH is reusable after regeneration with sodium
rescent molecule with antiscalants has been reported to allow the hydroxide solution, indicating that it is a promising adsorbent for
W. Yu et al. / Water Research 183 (2020) 115985 19

the removal and recovery of ATMP from RO membrane concen- cellulose, chitosan, and alginate, are considered good candidates
trates (Boels et al., 2012). used as efficient antiscalants because of their significantly
Antiscalants increase the difficulties of concentrate disposal; environment-friendly features. However, studies on the develop-
hence, an appropriate concentrate pretreatment that can remove or ment of novel polysaccharide-based antiscalants and their appli-
degrade antiscalants is favorable for improving the performance of cations to RO systems, particularly field operation, are lacking.
the succeeding concentrate treatment. Antiscalants, particularly
biodegradable ones, easily degrade under certain conditions, such 6.2.3. Fabrication of high-performance antiscalants
as using enzyme catalysis, which may be notably advantageous to A high-performance material must be designed and fabricated
final concentrate disposal. In summary, the effects of antiscalants on the basis of the structureeactivity relationship and its applica-
on concentrate disposal have not been fully studied and effective tion mechanisms (Ashby et al., 2013; Wessel, 2004). Suitable anti-
methods to mitigate their adverse effects are still rare. In-depth scalants with desired molecular structures can be accordingly
studies are necessary to improve the understanding of the role of obtained by using various chemical methods (Chen et al., 2015b;
antiscalants in concentrate disposal and the optimization of the Edgar et al., 2001; Heinze and Liebert, 2001; Kang et al., 2015; Roy
entire RO system. et al., 2009). However, previous studies related to inhibition
mechanisms were mostly superficial and focused on qualitative
6.2. Perspectives description; by contrast, quantitative analyses are rare (Wang et al.,
2017b; Yu et al., 2018, 2019). This condition may be attributed to
6.2.1. Optimizing the application of antiscalants in RO process two facts. The first is the lack of an in-depth understanding of
In spite of the effective scaling control of antiscalants in RO antiscalants’ molecular structures, particularly polymeric anti-
systems, many problems are associated with the real applications scalants, causing difficulty in studying the inhibition performance
of antiscalants as mentioned earlier. Considerable research effort of antiscalants at the molecular levels and establishing the
has been devoted to the improvement and optimization of anti- structureeactivity relationship. For example, the sequence struc-
scalants applied to RO systems (Ashfaq et al., 2019; Feng et al., 2014; tures of copolymers with more than two comonomers are difficult
Sweity et al., 2015; Xu et al., 2019). First, suitable antiscalants to determine using currently available structural characterization
should be selected in accordance with feed water characteristics methods (Kang et al., 2015; Roy et al., 2009). Hence, precise char-
(such as pH and composition) and the structural features of the acterization awaits the development of advanced technologies and
selected membrane. The optimization of RO operational conditions, sophisticated analytical instruments.
such as crossflow velocity, applied pressure, hydrodynamic condi- The other fact is the lack of understanding of the complicated
tions, and target recovery setting, is important in practical appli- physicochemical interactions involved among dominant scale-
cations (Antony et al., 2011; Goh et al., 2018; Greenlee et al., 2009a; forming substances, antiscalants, membranes, and other coexist-
Tong et al., 2019). For the operation of antiscalants, rapid adjust- ing substances in solutions (Karabelas et al., 2020; Matin et al.,
ment of antiscalants’ dose, in accordance with the results of 2019). In previous studies, researchers derived various inhibition
membrane scaling monitoring and antiscalant dosing monitoring, mechanisms, including chelation, dispersion, and crystal distortion,
can avoid potential aggravated scaling due to available changes in qualitatively based on apparent inhibition behavior and available
feed water characteristics and environmental conditions; thus, interactions. In fact, quantitative or semiquantitative details, such
monitoring methods for membrane scaling and antiscalant dosing as the degrees of their various interactions and the delayed and/or
control technologies are essential and should be further developed inhibited kinetic processes of inorganic scale crystallization, should
(Feng et al., 2014; Karabelas et al., 2020; Kelle Zeiher et al., 2003; be studied in depth at the molecular levels to guide the develop-
van de Lisdonk et al., 2001). The combination of antiscalants with ment of effective antiscalants. In addition to the effects of anti-
other RO-related pretreatments also plays an important role in scalants on the scaling ions and/or formed scales when they are
optimizing the utilization of antiscalants; besides, the inhibition applied to RO systems, the effects of antiscalants on the membrane
performance of antiscalants can be further improved using modi- should not be disregarded. Possible interfacial interactions between
fied membrane installation design and operation, both of which antiscalants and membrane surface, such as electrostatic attraction,
can more efficiently remove scale-forming substances from treated hydrogen bonding, hydrophobic interaction, and pep electro-
water and largely reduce the antiscalants’ doses. However, re- nedonoreacceptor interactions, can alter membrane surface
searches on exploring the synergistic effects of antiscalants with properties, influence membrane separation performance, and even
other pretreatments and different RO units on the improvement of cause membrane fouling. The design and preparation of anti-
the whole RO operation performance is notably necessary. scalants with enhanced inhibition effects but weakened fouling on
the basis of the detailed evaluation of their available involved
6.2.2. Environment friendliness of antiscalants multiple interactions are interesting topics. This knowledge can be
Apart from the aforementioned strategies in practical opera- obtained from the comprehensive analysis of the molecular struc-
tions, the development of novel, high-performance, and tural characteristics of scaling ions, formed scales, membranes, and
environment-friendly antiscalants should be the most critical point antiscalants. However, it is a quite complicated process to acquire a
that requires additional attention to obtain high scaling control high-performance practical antiscalant from laboratory to market,
performance. Phosphorus antiscalants should be used less or which still require close cooperation between relevant scientists
completely discarded because of their evident harmfulness to the and engineers.
environment, regardless of their good inhibition performance.
Phosphorus-free antiscalants, particularly biodegradable ones, 7. Conclusion
have greater prospects in real applications due to their significant
environment impact and their distinct antiscaling performance. In summary, (1) antiscalants, particularly environment-friendly
Besides, the difficulties of concentrate disposal may be solved by ones, have a wide application prospect in RO membrane scaling
using easily biodegradable antiscalants instead of those synthetic control due to their notable advantages as mentioned earlier; (2)
antiscalants, which would avoid that the presence of antiscalants The novel and high-performance antiscalants can be designed and
can hamper/complicate mining the elements in the brine. In fabricated on the basis of the well-established structure-activity;
addition to PASP and PESA, polysaccharides, including starch, and (3) remarkable achievements will be obtained through efforts
20 W. Yu et al. / Water Research 183 (2020) 115985

in the optimization of practical applications of antiscalants also. Boels, L., Witkamp, G.J., 2011. Carboxymethyl inulin biopolymers: a green alterna-
tive for phosphonate calcium carbonate growth inhibitors. Cryst. Growth Des.
Such achievements can favor the extensive applications of anti-
11 (9), 4155e4165.
scalants to membrane scaling control. Boerlage, S.F.E., Kennedy, M.D., Bremere, I., Witkamp, G.J., Van der Hoekc, J.P.S.,
Schippers, J.C., 2002. The scaling potential of barium sulphate in reverse
osmosis systems. J. Membr. Sci. 197, 151e168.
Declaration of competing interest Bonne , P.A.C., Hofman, J.A.M.H., van der Hoek, J.P., 2000. Scaling control of RO
membranes and direct treatment of surface water. Desalination 132, 109e119.
The authors declare that they have no known competing Bott, T.R., 1997. Aspects of crystallization fouling. Exp. Therm. Fluid Sci. 14 (4),
356e360.
financial interests or personal relationships that could have Bush, J.A., Vanneste, J., Gustafson, E.M., Waechter, C.A., Jassby, D., Turchi, C.S.,
appeared to influence the work reported in this paper. Cath, T.Y., 2018. Prevention and management of silica scaling in membrane
distillation using pH adjustment. J. Membr. Sci. 554, 366e377.
Butt, F.H., Rahman, F., Baduruthamal, U., 1997. Evaluation of SHMP and advanced
References scale inhibitors for control of CaSO4, SrSO4, and CaCO3 scales in RO desalination.
Desalination 109, 323e332.
Abdel-Aal, E.A., Abdel-Ghafar, H.M., El Anadouli, B.E., 2015. New findings about  M., Honzajkova
Bystrianský, M., Nir, O., Sír, , Z., Vurm, R., Hrychova, P., Bervic, A., van
nucleation and crystal growth of reverse osmosis desalination scales with and der Bruggen, B., 2016. The presence of ferric iron promotes calcium sulphate
without inhibitor. Cryst. Growth Des. 15 (10), 5133e5137. scaling in reverse osmosis processes. Desalination 393, 115e119.
Abdul Azis, P.K., Al-Tisan, I., Sasikumar, N., 2001. Biofouling potential and envi- Chauhan, K., Kumar, R., Kumar, M., Sharma, P., Chauhan, G.S., 2012. Modified pectin-
ronmental factors of seawater at a desalination plant intake. Desalination 135. based polymers as green antiscalants for calcium sulfate scale inhibition.
Agarwal, P., Berglund, K.A., 2004. Effect of polymeric additives on calcium carbonate Desalination 305, 31e37.
crystallization as monitored by nephelometry. Cryst. Growth Des. 4 (3), Chaussemier, M., Pourmohtasham, E.G., Dominique, Pe coul, N., Perrot, H., Le
dion, J.,
479e483. Cheap-Charpentier, H., Horner, O., 2015. State of art of natural inhibitors of
Agashichev, S.P., 2005. Reverse osmosis at elevated temperatures: influence of calcium carbonate scaling. A review article. Desalination 356, 47e55.
temperature on degree of concentration polarization and transmembrane flux. Chen, C.Y., Bai, N., Zhang, Y., Jiao, L.N., Xia, M.Z., Chen, G., 2017. A theoretical study of
Desalination 179 (1e3), 61e72. the inhibition effect of PAMAM molecule on silica scale. J. Mol. Model. 23 (1),

Akin, B., Oner, M., Bayram, Y., Demadis, K.D., 2008. Effects of carboxylate-modified, 32.
“green” inulin biopolymers on the crystal growth of calcium oxalate. Cryst. Chen, C.Y., Lei, W., Xia, M.Z., Wang, F.Y., Gong, X.D., 2013. Molecular modeling of
Growth Des. 8, 1997e2005. several phosphonates onto the stepped calcite (011) surface. Desalination 309,
Al-Amoudi, A.S., 2010. Factors affecting natural organic matter (NOM) and scaling 208e212.
fouling in NF membranes: a review. Desalination 259, 1e10. Chen, J.X., Xu, L.H., Han, J., Su, M., Wu, Q., 2015a. Synthesis of modified polyaspartic
Al-Hamzah, A.A., East, C.P., Doherty, W.O.S., Fellows, C.M., 2014. Inhibition of ho- acid and evaluation of its scale inhibition and dispersion capacity. Desalination
mogenous formation of calcium carbonate by poly(acrylic acid). The effect of 358, 42e48.
molar mass and end-group functionality. Desalination 338, 93e105. Chen, Q., Yu, H.J., Wang, L., ul Abdin, Z., Chen, Y.S., Wang, J.H., Zhou, W.D., Yang, X.P.,
Al-Hamzah, A.A., Fellows, C.M., 2015. A comparative study of novel scale inhibitors Khan, R.U., Zhang, H.T., Chen, X., 2015b. Recent progress in chemical modifi-
with commercial scale inhibitors used in seawater desalination. Desalination cation of starch and its applications. RSC Adv. 5 (83), 67459e67474.
359, 22e25. Dao, V.H., Cameron, N.R., Saito, K., 2016. Synthesis, properties and performance of
Al-Rammah, A., 2000. The application of acid free antiscalant to mitigate scaling in organic polymers employed in flocculation applications. Polym. Chem. 7 (1),
reverse osmosis membranes. Desalination 132, 83e87. 11e25.
Al-Roomi, Y.M., Hussain, K.F., 2016. Potential kinetic model for scaling and scale Darton, E.G., 2000. Membrane chemical research: centuries apart. Desalination 132,
inhibition mechanism. Desalination 393, 186e195. 121e131.
Al-Roomi, Y.M., Hussain, K.F., 2015. Application and evaluation of novel acrylic Demadis, K.D., Neofotistou, E., 2007. Synergistic effects of combinations of cationic
based CaSO4 inhibitors for pipes. Desalination 355, 33e44. polyaminoamide dentrimer anionic polyelectrolytes on amorphous silica for-
Al-Roomi, Y.M., Hussain, K.F., Al-Rifaie, M., 2015. Performance of inhibitors on mation: a bioinspired approach. Chem. Mater. 19, 581e587.
CaCO3 scale deposition in stainless steel & copper pipe surface. Desalination Demadis, K.D., Stathoulopoulou, A., 2006. Solubility enhancement of silicate with
375, 138e148. polyamine/polyammonium cationic macromolecules: relevance to silica-laden
Al-Shammiri, M., Safar, M., Al-Dawas, M., 2000. Evaluation of two different anti- process waters. Ind. Eng. Chem. Res. 45, 4436e4440.
scalants in real operation at the Doha research plant. Desalination 128 (1), 1e16. Dietzsch, M., Barz, M., Schuler, T., Klassen, S., Schreiber, M., Susewind, M., Loges, N.,
Ali, S.A., Kazi, I.W., Rahman, F., 2015. Synthesis and evaluation of phosphate-free Lang, M., Hellmann, N., Fritz, M., Fischer, K., Theato, P., Kuhnle, A., Schmidt, M.,
antiscalants to control CaSO4$2H2O scale formation in reverse osmosis desali- Zentel, R., Tremel, W., 2013. PAA-PAMPS copolymers as an efficient tool to
nation plants. Desalination 357, 36e44. control CaCO3 scale formation. Langmuir 29 (9), 3080e3088.
Amjad, Z., 2014. Mineral Scales in Biological and Industrial Systems. CRC Press, New Dobberschütz, S., Nielsen, M.R., Sand, K.K., Civioc, R., Bovet, N., Stipp, S.L.S.,
York. Andersson, M.P., 2018. The mechanisms of crystal growth inhibition by organic
Andrews, B., Dave , B., Lo
pez-Serrano, P., Tsai, S.P., Frank, R., Wilf, M., Koutsakos, E., and inorganic inhibitors. Nat. Commun. 9 (1), 1578.
2008. Effective scale control for seawater RO operating with high feed water pH Drak, A., Glucina, K., Busch, M., Hasson, D., Laine, J.M., Semiat, R., 2000. Laboratory
and temperature. Desalination 220 (1e3), 295e304. technique for predicting the scaling propensity of RO feed waters. Desalination
Ang, W.L., Mohammad, A.W., Benamor, A., Hilal, N., Leo, C.P., 2016. Hybrid 132, 233e242.
coagulation-NF membrane process for brackish water treatment: effect of Dydo, P., Turek, M., Ciba, J., Wandachowicz, K., Misztal, J., 2004. The nucleation
antiscalant on water characteristics and membrane fouling. Desalination 393, kinetic aspects of gypsum nanofiltration membrane scaling. Desalination 164,
144e150. 41e52.
Anis, S.F., Hashaikeh, R., Hilal, N., 2019. Reverse osmosis pretreatment technologies Edgar, K.J., Buchanan, C.M., Debenham, J.S., Rundquist, P.A., Seiler, B.D.,
and future trends: a comprehensive review. Desalination 452, 159e195. Shelton, M.C., Tindall, D., 2001. Advances in cellulose ester performance and
Antony, A., Low, J.H., Gray, S., Childress, A.E., Le-Clech, P., Leslie, G., 2011. Scale application. Prog. Polym. Sci. 26 (9), 1605e1688.
formation and control in high pressure membrane water treatment systems: a El-Shall, H., Rashad, M.M., Abdel-Aal, E.A., 2005. Effect of cetyl pyridinium additive
review. J. Membr. Sci. 383 (1e2), 1e16. on crystallization of gypsum in phosphoric and sulfuric acids medium Cryst.
Ashby, M., Shercliff, H., Cebon, D., 2013. Materials: Engineering, Science, Processing Respir. Technol. 40, 860e866.
and Design. Butterworth-Heinemann. Elias, H.G., 1984. Macromolecules, Structure and Properties, vol. 1. Plenum Press,
Ashfaq, M.Y., Al-Ghouti, M.A., Qiblawey, H., Zouari, N., 2019. Evaluating the effect of New York.
antiscalants on membrane biofouling using FTIR and multivariate analysis. Feng, J.Y., Gao, L.J., Wen, R.Z., Deng, Y.Y., Wu, X.J., Deng, S.L., 2014. Fluorescent
Biofouling 35 (1), 1e14. polyaspartic acid with an enhanced inhibition performance against calcium
Aziz, M., Kasongo, G., 2019. Scaling prevention of thin film composite polyamide phosphate. Desalination 345, 72e76.
Reverse Osmosis membranes by Zn ions. Desalination 464, 76e83. Fischer, V., Landfester, K., Munoz-Espí, R., 2011. Stabilization of calcium oxalate
Badruzzaman, M., Voutchkov, N., Weinrich, L., Jacangelo, J.G., 2019. Selection of metastable phases by oligo(L-glutamic acid): effect of peptide chain length.
pretreatment technologies for seawater reverse osmosis plants: a review. Cryst. Growth Des. 11, 1880e1890.
Desalination 449, 78e91. Flory, P.J., 1953. Principles of Polymer Chemistry. Cornell University Press, New York.
Bartman, A.R., Lyster, E., Rallo, R., Christofides, P.D., Cohen, Y., 2011. Mineral scale Fritzmann, C., Lowenberg, J., Wintgens, T., Melin, T., 2007. State-of-the-art of reverse
monitoring for reverse osmosis desalination via real-time membrane surface osmosis desalination. Desalination 216 (1e3), 1e76.
image analysis. Desalination 273 (1), 64e71. Gao, Y.H., Fan, L.H., Ward, L., Liu, Z.F., 2015. Synthesis of polyaspartic acid derivative
Benecke, J., Rozova, J., Ernst, M., 2018. Anti-scale effects of select organic macro- and evaluation of its corrosion and scale inhibition performance in seawater
molecules on gypsum bulk and surface crystallization during reverse osmosis utilization. Desalination 365, 220e226.
desalination. Separ. Purif. Technol. 198, 68e78. GB/T 16632-2008, 2008. Determination of scale inhibition performance of water
Boels, L., Keesman, K.J., Witkamp, G.J., 2012. Adsorption of phosphonate antiscalant treatment agentsddcacium carbonate precipitation method. Issued by general
from reverse osmosis membrane concentrate onto granular ferric hydroxide. administration of quality supervision, inspection and quarantine of the people’s
Environ. Sci. Technol. 46 (17), 9638e9645. Republic of China and standardization administration of the people’s Republic
W. Yu et al. / Water Research 183 (2020) 115985 21

of China, 1. Adopted on Septemper. for cooling water system application. Desalination 268 (1e3), 38e45.
Ghani, S., Al-Deffeeri, N.S., 2010. Impacts of different antiscalant dosing rates and Kavitskaya, A.A., Knyazkova, T.V., Maynarovich, A.A., 2000. Reverse osmosis of
their thermal performance in Multi stage Flash (MSF) distiller in Kuwait. concentrated calcium sulphate solutions in the presence of iron (III) ions using
Desalination 250, 463e472. composite membranes. Desalination 132, 281e286.
Ghizellaoui, S., Euvrard, M., 2008. Assessing the effect of zinc on the crystallization Kelle Zeiher, E.H., Ho, B., Williams, K.D., 2003. Novel antiscalant dosing control.
of calcium carbonate. Desalination 220 (1e3), 394e402. Desalination 157, 209e216.
Gill, J.S., 1999. A novel inhibitor for scale control in water desalination. Desalination Ketrane, R., Saidani, B., Gil, O., Leleyter, L., Baraud, F., 2009. Efficiency of five scale
124 (1e3), 43e50. inhibitors on calcium carbonate precipitation from hard water: effect of tem-
Goh, P.S., Lau, W.J., Othman, M.H.D., Ismail, A.F., 2018. Membrane fouling in desa- perature and concentration. Desalination 249, 1397e1404.
lination and its mitigation strategies. Desalination 425, 130e155. Ketsetzi, A., Stathoulopoulou, A., Demadis, K.D., 2008. Being "green" in chemical
Goncharuk, V.V., Kavitskaya, A.A., Skilskaya, M.D., 2012. Sodium carboxymethyl water treatment technologies: issues, challenges and developments. Desalina-
cellulose as an inhibitor of scale formation in nanofiltration of hard artesian tion 223 (1e3), 487e493.
waters. Desalin. Water Treat. 47 (1e3), 235e242. Kim, I.W., Robertson, R.E., Zand, R., 2005. Effects of some nonionic polymeric ad-
Greenlee, L.F., Lawler, D.F., Freeman, B.D., Marrot, B., Moulin, P., 2009a. Reverse ditives on the crystallization of calcium carbonate. Cryst. Growth Des. 5 (2),
osmosis desalination: water sources, technology, and today’s challenges. Water 513e522.
Res. 43 (9), 2317e2348. Kırbog €
a, S., Oner, M., 2009. Inhibition of calcium oxalate crystallization by graft
Greenlee, L.F., Marrot, B., Moulin, P., Freeman, B.D., Lawler, D.F., 2009b. Treatment of copolymers. Cryst. Growth Des. 9 (5), 2159e2167.
reverse osmosis concentrate to improve overall recovery: parameter effects on Lakshminarayanan, R., Valiyaveettil, S., Loy, G.L., 2003. Selective nucleation of cal-
antiscalant oxidation and subsequent precipitation. Abstr. Pap. Am. Chem. Soc. cium carbonate polymorphs: role of surface functionalization and poly(vinyl
237, 1. alcohol) additive. Cryst. Growth Des. 3, 953e958.
Greenlee, L.F., Testa, F., Lawler, D.F., Freeman, B.D., Moulin, P., 2010a. The effect of Lakshtanov, L.Z., Bovet, N., Stipp, S.L.S., 2011. Inhibition of calcite growth by alginate.
antiscalant addition on calcium carbonate precipitation for a simplified syn- Geochem. Cosmochim. Acta 75, 3945e3955.
thetic brackish water reverse osmosis concentrate. Water Res. 44 (9), Le Gouellec, Y.A., Elimelech, M., 2002a. Calcium sulfate (gypsum) scaling in nano-
2957e2969. filtration of agricultural drainage water. J. Membr. Sci. 205, 279e291.
Greenlee, L.F., Testa, F., Lawler, D.F., Freeman, B.D., Moulin, P., 2010b. Effect of Le Gouellec, Y.A., Elimelech, M., 2002b. Control of calcium sulfate (gypsum) scale in
antiscalants on precipitation of an RO concentrate: metals precipitated and nanofiltration of saline agricultural drainage water. Environ. Eng. Sci. 19 (6),
particle characteristics for several water compositions. Water Res. 44 (8), 387e397.
2672e2684. Lee, S., Choi, J.S., Lee, C.H., 2009. Behaviors of dissolved organic matter in membrane
Greenlee, L.F., Testa, F., Lawler, D.F., Freeman, B.D., Moulin, P., 2011. Effect of anti- desalination. Desalination 238, 109e116.
scalant degradation on salt precipitation and solid/liquid separation of RO Lee, T.J., Hong, S.J., Park, J.Y., Kim, H.J., 2015. Effects of anionic polyacrylamide on
concentrate. J. Membr. Sci. 366 (1e2), 48e61. carbonation for the crystallization of precipitated calcium carbonate. Cryst.
Gryta, M., 2012. Polyphosphates used for membrane scaling inhibition during water Growth Des. 15 (4), 1652e1657.
desalination by membrane distillation. Desalination 285, 170e176. Li, H., Hsieh, M.K., Chien, S.H., Monnell, J.D., Dzombak, D.A., Vidic, R.D., 2011. Control
Guo, W.S., Ngo, H.H., Li, J.X., 2012. A mini-review on membrane fouling. Bioresour. of mineral scale deposition in cooling systems using secondary-treated
Technol. 122, 27e34. municipal wastewater. Water Res. 45 (2), 748e760.
Guo, X.R., Qiu, F.X., Dong, K., He, K.C., Rong, X.S., Yang, D.Y., 2013. Scale inhibitor Li, J., Zhou, Y.M., Yao, Q.Z., Wang, T.T., Zhang, A., Chen, Y.Y., Wu, W.D., Sun, W., 2017.
copolymer modified with oxidized starch: synthesis and performance on scale Preparation and evaluation of a polyether-based polycarboxylate as a kind of
inhibition. Polym. Plast. Technol. 52, 261e267. inhibitor for water systems. Ind. Eng. Chem. Res. 56 (10), 2624e2633.
Gwak, G., Hong, S., 2017. New approach for scaling control in forward osmosis (FO) Li, S.L., Qu, Q., Li, L., Xia, K., Li, Y., Zhu, T.T., 2019. Bacillus cereus s-EPS as a dual bio-
by using an antiscalant-blended draw solution. J. Membr. Sci. 530, 95e103. functional corrosion and scale inhibitor in artificial seawater. Water Res. 166,
Hasson, D., Bramson, D., Litnoni-Relis, B., Semiat, R., 1996. Influence of the flow 115094.
system on the inhibitory action of CaCO3 scale prevention additives. Desalina- Lin, Y.P., Singer, P.C., 2005. Inhibition of calcite crystal growth by polyphosphates.
tion 108, 67e79. Water Res. 39 (19), 4835e4843.
Hasson, D., Drak, A., Semiat, R., 2001. Inception of CaSO4 scaling on RO membranes Lisitsin, D., Yang, Q.F., Hasson, D., Semiat, R., 2005. Inhibition of CaCO3 scaling on RO
at various water recovery levels. Desalination 139, 73e81. membranes by trace amounts of zinc ions. Desalination 183 (1e3), 289e300.
Hasson, D., Drak, A., Semiat, R., 2003. Induction times induced in an RO system by Liu, Y., Cui, Y.J., Mao, H.Y., Guo, R., 2012b. Calcium carbonate crystallization in the
antiscalants delaying CaSO4 precipitation. Desalination 157, 193e207. presence of Casein. Cryst. Growth Des. 12 (10), 4720e4726.
Hasson, D., Semiat, R., Bramson, D., Busch, M., Limoni-Relis, B., 1998. Suppression of Liu, D., Dong, W.B., Li, F., Hui, F.T., Le dion, J., 2012a. Comparative performance of
CaCO3 scale deposition by anti-scalants. Desalination 118, 285e296. polyepoxysuccinic acid and polyaspartic acid on scaling inhibition by static and
Hasson, D., Shemer, H., Sher, A., 2011. State of the art of friendly “green” scale rapid controlled precipitation methods. Desalination 304, 1e10.
control inhibitors: a review article. Ind. Eng. Chem. Res. 50 (12), 7601e7607. Liu, Q., Xu, G.R., Das, R., 2019. Inorganic scaling in reverse osmosis (RO) desalina-
He, F., Sirkar, K.K., Gilron, J., 2009. Effects of antiscalants to mitigate membrane tion: mechanisms, monitoring, and inhibition strategies. Desalination 468,
scaling by direct contact membrane distillation. J. Membr. Sci. 345 (1e2), 114065.
53e58. Lyster, E., Kim, M.M., Au, J., Cohen, Y., 2010. A method for evaluating antiscalant
Heinze, T., Liebert, T., 2001. Unconventional methods in cellulose functionalization. retardation of crystal nucleation and growth on RO membranes. J. Membr. Sci.
Prog. Polym. Sci. 26 (9), 1689e1762. 364 (1e2), 122e131.
Henthorne, L., Boysen, B., 2015. State-of-the-art of reverse osmosis desalination Malaeb, L., Ayoub, G.M., 2011. Reverse osmosis technology for water treatment:
pretreatment. Desalination 356, 129e139. state of the art review. Desalination 267 (1), 1e8.
Her, N., Amy, G., Jarusutthirak, C., 2000. Seasonal variations of nanofiltration (NF) Manoli, F., Dalas, E., 2001. Calcium carbonate crystallization in the presence of
foulants: identification and control. Desalination 132, 143e160. glutamic acid. J. Cryst. Growth 222 (1e2), 293e297.
HG/T 2024-2009, 2009. Determination of scaIe inhibition performance of water Matin, A., Khan, Z., Zaidi, S.M.J., Boyce, M.C., 2011. Biofouling in reverse osmosis
treatment agentsddbubble method. Issued by the Ministry of Industry and membranes for seawater desalination: phenomena and prevention. Desalina-
Information Technology of the People’s Republic of China, 1. Adopted on June. tion 281, 1e16.
Jamaly, S., Darwish, N.N., Ahmed, I., Hasan, S.W., 2014. A short review on reverse Matin, A., Rahman, F., Shafi, H.Z., Zubair, S.M., 2019. Scaling of reverse osmosis
osmosis pretreatment technologies. Desalination 354, 30e38. membranes used in water desalination: phenomena, impact, and control;
Jawor, A., Hoek, E.M.V., 2009. Effects of feed water temperature on inorganic fouling future directions. Desalination 455, 135e157.
of brackish water RO membranes. Desalination 235, 44e57. McCool, B.C., Rahardianto, A., Cohen, Y., 2012. Antiscalant removal in accelerated
Ji, Y., Chen, Y., Le, J.X., Qian, M.Q., Huan, Y., Yang, W.Z., Yin, X.S., Liu, Y., Wang, X.R., desupersaturation of RO concentrate via chemically-enhanced seeded precipi-
Chen, Y.Z., 2017. Highly effective scale inhibition performance of amino trime- tation (CESP). Water Res. 46 (13), 4261e4271.
thylenephosphonic acid on calcium carbonate. Desalination 422, 165e173. McCool, B.C., Rahardianto, A., Faria, J.I., Cohen, Y., 2013. Evaluation of chemically-
Jiang, S.X., Li, Y.N., Ladewig, B.P., 2017. A review of reverse osmosis membrane enhanced seeded precipitation of RO concentrate for high recovery desalting
fouling and control strategies. Sci. Total Environ. 595, 567e583. of high salinity brackish water. Desalination 317, 116e126.
Joo, S.H., Tansel, B., 2015. Novel technologies for reverse osmosis concentrate Milne, N.A., O’Reilly, T., Sanciolo, P., Ostarcevic, E., Beighton, M., Taylor, K.,
treatment: a review. J. Environ. Manag. 150, 322e335. Mullett, M., Tarquin, A.J., Gray, S.R., 2014. Chemistry of silica scale mitigation for
Kang, G.D., Cao, Y.M., 2012. Development of antifouling reverse osmosis membranes RO desalination with particular reference to remote operations. Water Res. 65,
for water treatment: a review. Water Res. 46 (3), 584e600. 107e133.
Kang, H.L., Liu, R.G., Huang, Y., 2015. Graft modification of cellulose: methods, Mullin, J.W., 2001. Nucleation. Crystallization. Butterworth-Heinemann, Oxford,
properties and applications. Polymer 70, A1eA16. pp. 181e215.
Karabelas, A.J., Karanasiou, A., Sioutopoulos, D.C., 2017. Experimental study on the Neofotistou, E., Demadis, K.D., 2004a. Use of antiscalants for mitigation of silica
effect of polysaccharides on incipient membrane scaling during desalination. (SiO2) fouling and deposition: fundamentals and applications in desalination
Desalination 416, 106e121. systems. Desalination 167 (1e3), 257e272.
Karabelas, A.J., Mitrouli, S.T., Kostoglou, M., 2020. Scaling in reverse osmosis desa- Neofotistou, E., Demadis, K.D., 2004b. Silica scale inhibition by polyaminoamide
lination plants: a perspective focusing on development of comprehensive STARBURST® dendrimers. Colloid. Surface. A Physicochem. Eng. Aspect. 242
simulation tools. Desalination 474, 114193. (1e3), 213e216.
Kavitha, A.L., Vasudevan, T., Prabu, H.G., 2011. Evaluation of synthesized antiscalants Nielsen, J.W., Sand, K.K., Pedersen, C.S., Lakshtanov, L.Z., Winther, J.R.,
22 W. Yu et al. / Water Research 183 (2020) 115985

Willemoes, M., Stipp, S.L.S., 2012. Polysaccharide effects on calcite growth: the mechanism: a case of itaconic acid-epoxysuccinate copolymer. Comput. Mater.
influence of composition and branching. Cryst. Growth Des. 12 (10), Sci. 136, 118e125.
4906e4910. Shih, W.Y., Rahardianto, A., Lee, R.W., Cohen, Y., 2005. Morphometric character-
Ning, R.Y., 2002. Discussion of silica speciation, fouling, control and maximum ization of calcium sulfate dihydrate (gypsum) scale on reverse osmosis mem-
reduction. Desalination 151, 67e73. branes. J. Membr. Sci. 252 (1e2), 253e263.
Nowack, B., 2003. Environmental chemistry of phosphonates. Water Res. 37, Shirazi, S., Lin, C.J., Chen, D., 2010. Inorganic fouling of pressure-driven membrane
2533e2546. processes d a critical review. Desalination 250 (1), 236e248.
Oh, H.J., Choung, Y.K., Lee, S., Choi, J.S., Hwang, T.M., Kim, J.H., 2009. Scale formation Shmulevsky, M., Li, X.H., Shemer, H., Hasson, D., Semiat, R., 2017. Analysis of the
in reverse osmosis desalination: model development. Desalination 238 (1e3), onset of calcium sulfate scaling on RO membranes. J. Membr. Sci. 524, 299e304.
333e346. € berg, S., 1996. Silica in aqueous environments. J. Non-Cryst. Solids 196, 51e57.
Sjo
Olderøy, M.Ø., Xie, M.L., Strand, B.L., Flaten, E.M., Sikorski, P., Andreassen, J.P., 2009. Subramani, A., Jacangelo, J.G., 2014. Treatment technologies for reverse osmosis
Growth and nucleation of calcium carbonate vaterite crystals in presence of concentrate volume minimization: a review. Separ. Purif. Technol. 122,
alginate. Cryst. Growth Des. 9 (12), 5176e5183. 472e489.
rez-Gonz
Pe alez, A., Urtiaga, A.M., Ib
an~ ez, R., Ortiz, I., 2012. State of the art and re- Sutzkover, I., Hasson, D., Semiat, R., 2000. Simple technique for measuring the
view on the treatment technologies of water reverse osmosis concentrates. concentration polarization level in a reverse osmosis system. Desalination 131,
Water Res. 46 (2), 267e283. 117e127.
Pramanik, B.K., Gao, Y.H., Fan, L.H., Roddick, F.A., Liu, Z.F., 2017. Antiscaling effect of Sweity, A., Oren, Y., Ronen, Z., Herzberg, M., 2013. The influence of antiscalants on
polyaspartic acid and its derivative for RO membranes used for saline waste- biofouling of RO membranes in seawater desalination. Water Res. 47 (10),
water and brackish water desalination. Desalination 404, 224e229. 3389e3398.
Prihasto, N., Liu, Q.F., Kim, S.H., 2009. Pre-treatment strategies for seawater desa- Sweity, A., Ronen, Z., Herzberg, M., 2014. Induced organic fouling with antiscalants
lination by reverse osmosis system. Desalination 249 (1), 308e316. in seawater desalination. Desalination 352, 158e165.
Qasim, M., Badrelzaman, M., Darwish, N.N., Darwish, N.A., Hilal, N., 2019. Reverse Sweity, A., Zere, T.R., David, I., Bason, S., Oren, Y., Ronen, Z., Herzberg, M., 2015. Side
osmosis desalination: a state-of-the-art review. Desalination 459, 59e104. effects of antiscalants on biofouling of reverse osmosis membranes in brackish
Qiang, X.h., Sheng, Z.h., Zhang, H., 2013. Study on scale inhibition performances and water desalination. J. Membr. Sci. 481, 172e187.
interaction mechanism of modified collagen. Desalination 309, 237e242. Tang, Y.M., Yang, W.Z., Yin, X.S., Liu, Y., Yin, P.W., Wang, J.T., 2008. Investigation of
Qin, J.J., Oo, M.H., Wong, F.S., 2005. Effects of pH and antiscalant on fouling of RO CaCO3 scale inhibition by PAA, ATMP and PAPEMP. Desalination 228 (1e3),
membrane for reclamation of spent rinse water from metal plating. Separ. Purif. 55e60.
Technol. 46 (1e2), 46e50. Thompson, J., Rahardianto, A., Kim, S., Bilal, M., Breckenridge, R., Cohen, Y., 2017.
Rabizadeh, T., Stawski, T.M., Morgan, D.J., Peacock, C.L., Benning, L.G., 2017. The Real-time direct detection of silica scaling on RO membranes. J. Membr. Sci. 528,
effects of inorganic additives on the nucleation and growth kinetics of calcium 346e358.
sulfate dihydrate crystals. Cryst. Growth Des. 17 (2), 582e589. Tong, T.Z., Wallace, A.F., Zhao, S., Wang, Z., 2019. Mineral scaling in membrane
Rahardianto, A., Gao, J.B., Gabelich, C.J., Williams, M.D., Cohen, Y., 2007. High re- desalination: mechanisms, mitigation strategies, and feasibility of scaling-
covery membrane desalting of low-salinity brackish water: integration of resistant membranes. J. Membr. Sci. 579, 52e69.
accelerated precipitation softening with membrane RO. J. Membr. Sci. 289 Topçu, G., Çelik, A., Baba, A., Demir, M.M., 2017. Design of polymeric antiscalants
(1e2), 123e137. based on functional vinyl monomers for (Fe, Mg) silicates. Energy Fuels 31 (8),
Rahardianto, A., McCool, B.C., Cohen, Y., 2008. Reverse osmosis desalting of inland 8489e8496.
brackish water of high gypsum scaling propensity: kinetics and mitigation of Turek, M., Mitko, K., Piotrowski, K., Dydo, P., Laskowska, E., Jako  bik-Kolon, A., 2017.
membrane mineral scaling. Environ. Sci. Technol. 42 (12), 4292e4297. Prospects for high water recovery membrane desalination. Desalination 401,
Rahardianto, A., Shih, W.Y., Lee, R.W., Cohen, Y., 2006. Diagnostic characterization of 180e189.
gypsum scale formation and control in RO membrane desalination of brackish Tzotzi, C., Pahiadaki, T., Yiantsios, S.G., Karabelas, A.J., Andritsos, N., 2007. A study of
water. J. Membr. Sci. 279 (1e2), 655e668. CaCO3 scale formation and inhibition in RO and NF membrane processes.
Rahman, F., 2013. Calcium sulfate precipitation studies with scale inhibitors for J. Membr. Sci. 296 (1e2), 171e184.
reverse osmosis desalination. Desalination 319, 79e84. Uchymiak, M., Lyster, E., Glater, J., Cohen, Y., 2008. Kinetics of gypsum crystal
Rashchi, F., Finch, J.A., 2000. Polyphosphates: a review their chemistry and appli- growth on a reverse osmosis membrane. J. Membr. Sci. 314 (1e2), 163e172.
cation with particular reference to mineral processing. Miner. Eng. 13 (10e11), van de Lisdonk, C.A.C., Rietman, B.M., Heijman, S.G.J., Sterk, G.R., Schippers, J.C.,
1019e1035. 2001. Prediction of supersaturation and monitoring of scaling in reverse
Roy, D., Semsarilar, M., Guthrie, J.T., Perrier, S., 2009. Cellulose modification by osmosis and nanofiltration membrane systems. Desalination 138, 259e270.
polymer grafting: a review. Chem. Soc. Rev. 38 (7), 2046e2064. Vrouwenvelder, J.S., Beyer, F., Dahmani, K., Hasan, N., Galjaard, G., Kruithof, J.C., Van
Ruiz-Agudo, C., Putnis, C.V., Iban ~ ez-Velasco, A., Ruiz-Agudo, E., Putnis, A., 2016. Loosdrecht, M.C.M., 2010. Phosphate limitation to control biofouling. Water Res.
A potentiometric study of the performance of a commercial copolymer in the 44 (11), 3454e3466.
precipitation of scale forming minerals. CrystEngComm 18 (30), 5744e5753. Vrouwenvelder, J.S., Manolarakis, S.A., Veenendaal, H.R., van der Kooij, D., 2000.
S
anchez, O., 2018. Microbial diversity in biofilms from reverse osmosis membranes: Biofouling potential of chemicals used for scale control in RO and NF mem-
a short review. J. Membr. Sci. 545, 240e249. branes. Desalination 132, 1e10.
Schindler, P.W., Fürst, B., Dick, R., Wolf, P.U., 1976. Ligand properties of surface Waly, T., Kennedy, M.D., Witkamp, G.J., Amy, G., Schippers, J.C., 2012. The role of
silanol groups. I. surface complex formation with Fe3þ, Cu2þ, Cd2þ, and Pb2þ. inorganic ions in the calcium carbonate scaling of seawater reverse osmosis
J. Colloid Interface Sci. 55, 469e475. systems. Desalination 284, 279e287.
Semiat, R., Sutzkover, I., Hasson, D., 2001. Technique for evaluating silica scaling and Wang, L.C., Cui, K., Wang, L.B., Li, H.X., Li, S.F., Zhang, Q.L., Liu, H.B., 2016b. The effect
its inhibition in RO desalting. Desalination 140, 181e193. of ethylene oxide groups in alkyl ethoxy carboxylates on its scale inhibition
Semiat, R., Sutzkover, I., Hasson, D., 2003. Characterization of the effectiveness of performance. Desalination 379, 75e84.
silica anti-sealants. Desalination 159, 11e19. Wang, C., Li, S.P., Li, T.D., 2009. Calcium carbonate inhibition by a phosphonate-
Senthilmurugan, B., Ghosh, B., Kundu, S.S., Haroun, M., Kameshwari, B., 2010. Maleic terminated poly(maleic-co-sulfonate) polymeric inhibitor. Desalination 249,
acid based scale inhibitors for calcium sulfate scale inhibition in high temper- 1e4.
ature application. J. Petrol. Sci. Eng. 75, 189e195. Wang, Y.N., Li, X.S., Wang, R., 2017a. Silica scaling and scaling control in pressure
Shahid, M.K., Pyo, M., Choi, Y.G., 2018. The operation operation of reverse osmosis retarded osmosis processes. J. Membr. Sci. 541, 73e84.
system with CO2 as a scale inhibitor: a study on operational behavior and Wang, Y.W., Li, A.M., Yang, H., 2017b. Effects of substitution degree and molecular
membrane morphology. Desalination 426, 11e20. weight of carboxymethyl starch on its scale inhibition. Desalination 408, 60e69.
Shakkthivel, P., Sathiyamoorthi, R., Vasudevan, T., 2004. Development of acryloni- Wang, J.X., Wang, L., Miao, R., Lv, Y.T., Wang, X.D., Meng, X.R., Yang, R.S., Zhang, X.T.,
trile copolymers for scale control in cooling water systems. Desalination 164, 2016a. Enhanced gypsum scaling by organic fouling layer on nanofiltration
111e123. membrane: characteristics and mechanisms. Water Res. 91, 203e213.
Shakkthivel, P., Vasudevan, T., 2006. Acrylic acid-diphenylamine sulphonic acid Wang, C.Y., Zhao, J.Z., Zhao, X., Bala, H., Wang, Z.C., 2006. Synthesis of nanosized
copolymer threshold inhibitor for sulphate and carbonate scales in cooling calcium carbonate (aragonite) via a polyacrylamide inducing process. Powder
water systems. Desalination 197, 179e189. Technol. 163 (3), 134e138.
Shen, C., Xu, X., Hou, X.Y., Wu, D.X., Yin, J.H., 2018. Molecular weight effect on PAA Wang, H.C., Zhou, Y.M., Yao, Q.Z., Ma, S.S., Wu, W.D., Sun, W., 2014. Synthesis of
antiscale performance in LT-MED desalination system: static experiment and fluorescent-tagged scale inhibitor and evaluation of its calcium carbonate
MD simulation. Desalination 445, 1e5. precipitation performance. Desalination 340, 1e10.
Shenvi, S.S., Isloor, A.M., Ismail, A.F., 2015. A review on RO membrane technology: Weng, P.F., 1995. Silica scale inhibition and colloidal silica dispersion for reverse
developments and challenges. Desalination 368, 10e26. osmosis systems. Desalination 103, 59e67.
Shi, W.Y., Ding, C., Yan, J.L., Han, X.Y., Lv, Z.M., Lei, W., Xia, M.Z., Wang, F.Y., 2012. Wessel, J.K., 2004. Handbook of Advanced Materials: Enabling New Designs. John
Molecular dynamics simulation for interaction of PESA and acrylic copolymers Wiley & Sons, Inc., Hoboken, New Jersey, USA.
with calcite crystal surfaces. Desalination 291, 8e14. Xu, Y., Zhao, L.L., Wang, L.N., Xu, S.Y., Cui, Y.C., 2012. Synthesis of polyaspartic
Shi, W.Y., Xia, M.Z., Lei, W., Wang, F.Y., 2013. Molecular dynamics study of polyether acidemelamine grafted copolymer and evaluation of its scale inhibition per-
polyamino methylene phosphonates as an inhibitor of anhydrite crystal. formance and dispersion capacity for ferric oxide. Desalination 286, 285e289.
Desalination 322, 137e143. Xu, Z.B., Wang, W.L., Huang, N., Wu, Q.Y., Lee, M.Y., Hu, H.Y., 2019. 2-
Shi, W.Y., Xu, W., Cang, H., Yan, X.H., Shao, R., Zhang, Y.H., Xia, M.Z., 2017. Design and Phosphonobutane-1,2,4-tricarboxylic acid (PBTCA) degradation by ozonation:
synthesis of biodegradable antiscalant based on MD simulation of antiscale kinetics, phosphorus transformation, anti-precipitation property changes and
W. Yu et al. / Water Research 183 (2020) 115985 23

phosphorus removal. Water Res. 148, 334e343. Zhang, B.R., Chen, Y.N., Li, F.T., 2011. Inhibitory effects of poly(adipic acid/amine-
Yang, Q.F., Liu, Y.Q., Li, Y.J., 2010c. Humic acid fouling mitigation by antiscalant in terminated polyether D230/diethylenetriamine) on colloidal silica formation.
reverse osmosis system. Environ. Sci. Technol. 44 (13), 5153e5158. Colloid. Surface. Physicochem. Eng. Aspect. 385, 11e19.
Yang, Q.F., Liu, Y.Q., Li, Y.J., 2010b. Control of protein (BSA) fouling in RO system by Zhang, B.R., Sun, P.D., Chen, F., Li, F.T., 2012. Synergistic inhibition effect of poly-
antiscalants. J. Membr. Sci. 364 (1e2), 372e379. aminoamide dendrimers and PESA on silica polymerization. Colloid. Surface.
Yang, L., Yang, W.Z., Xu, B., Yin, X.S., Chen, Y., Liu, Y., Ji, Y., Huan, Y., 2017. Synthesis Physicochem. Eng. Aspect. 410, 159e169.
and scale inhibition performance of a novel environmental friendly and hy- Zhang, H.X., Wang, F., Jin, X.H., Zhu, Y.C., 2013. A botanical polysaccharide extracted
drophilic terpolymer inhibitor. Desalination 416, 166e174. from abandoned corn stalks. Desalination 326, 55e61.
Yang, H.L., Huang, C.P., Lin, J.C.T., 2010a. Seasonal fouling on seawater desalination Zhang, R.N., Liu, Y., He, M.R., Su, Y.L., Zhao, X.T., Elimelech, M., Jiang, Z.Y., 2016b.
RO membrane. Desalination 250 (2), 548e552. Antifouling membranes for sustainable water purification: strategies and
Yang, R., Li, H.J., Huang, M., Yang, H., Li, A.M., 2016. A review on chitosan-based mechanisms. Chem. Soc. Rev. 45 (21), 5888e5924.
flocculants and their applications in water treatment. Water Res. 95, 59e89. Zhang, H.P., Luo, X.G., Lin, X.Y., Tang, P.P., Lu, X., Yang, M.J., Tang, Y.H., 2016a.
Yu, Q., Ou, H.D., Song, R.Q., Xu, A.W., 2006. The effect of polyacrylamide on the Biodegradable carboxymethyl inulin as a scale inhibitor for calcite crystal
crystallization of calcium carbonate: synthesis of aragonite single-crystal growth: molecular level understanding. Desalination 381, 1e7.
nanorods and hollow vatarite hexagons. J. Cryst. Growth 286 (1), 178e183. Zhang, S.P., Qu, H.J., Yang, Z., Fu, C.E., Tian, Z.Q., Yang, W.B., 2017. Scale inhibition
Yu, W., Song, D., Li, A.M., Yang, H., 2019. Control of gypsum-dominated scaling in performance and mechanism of sulfamic/amino acids modified polyaspartic
reverse osmosis system using carboxymethyl cellulose. J. Membr. Sci. 577, acid against calcium sulfate. Desalination 419, 152e159.
20e30. Zhang, X.Y., 2011. Practical Chemistry Handbook. National Defense Industry Press,
Yu, W., Wang, Y.W., Li, A.M., Yang, H., 2018. Evaluation of the structural morphology Beijing.
of starch-graft-poly(acrylic acid) on its scale-inhibition efficiency. Water Res. Zhang, Y., Yin, H.Q., Zhang, Q.S., Li, Y.Z., Yao, P.J., 2016c. Synthesis and character-
141, 86e95. ization of novel polyaspartic acid/urea graft copolymer with acylamino group
Yuchi, A., Gotoh, Y., Itoh, S., 2007. Potentiometry of effective concentration of pol- and its scale inhibition performance. Desalination 395, 92e98.
yacrylate as scale inhibitor. Anal. Chim. Acta 594 (2), 199e203.

You might also like