Sie sind auf Seite 1von 9

Desalination 335 (2014) 5563

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

Evaluation of maleic acid based polymers as scale inhibitors


and dispersants for industrial water applications
Zahid Amjad a, Petros G. Koutsoukos b,
a
b

Walsh University, Division of Mathematics and Sciences, North Canton, OH 44720, USA
University of Patras, Department of Chemical Engineering, Patras, University Campus, 26500 Patras, Greece and FORTH-ICEHT, Patras, GREECE

H I G H L I G H T S
Prevention of scale formation and particles dispersion important for fouling minimization.
Polymers and copolymers of maleic acid are excellent compounds for these purposes.
Polymer molecular weight and functional groups control fouling and dispersion.

a r t i c l e

i n f o

Article history:
Received 17 September 2013
Received in revised form 11 December 2013
Accepted 14 December 2013
Available online 9 January 2014
Keywords:
Inorganic scale
Calcium carbonate
Calcium sulfate dihydrate
Maleic acid polymers
Inhibitors of crystal growth
Hematite dispersion

a b s t r a c t
The formation of inorganic scale deposits of the alkaline earth metals is a persistent problem. Dispersion of solid
particles separating out from the uids is also very important for fouling due to deposition. Scale formation and
stabilization of suspensions are often overcome through the use of water soluble polymers. In the present work, a
series of polymeric compounds were tested as inhibitors of calcium carbonate and calcium sulfate dihydrate
(gypsum) precipitation and as dispersion agents of hematite (Fe2O3) suspensions in electrolyte solutions. Acrylic
acid (AA) and maleic acid (MA) polymers were found to inhibit the precipitation of calcium carbonate and gypsum from supersaturated solutions to extents exceeding 90% at concentrations as low as 2 ppm. The molecular
weight (MW) was an important parameter in determining the activity of the tested inhibitors. Lower MW polymers (ca. 2000) proved to be more efcient than higher MW polymers. A similar trend was exhibited for the dispersion ability of the hematite particles. AA and MA copolymers in which functional groups were introduced
(pyrrole, sulfono and amide groups) were efcient precipitation inhibitors and dispersion agents, but the efciency depended strongly on the co-polymer architecture. Copolymers containing sulfono groups improved inhibitory activity and dispersion ability and showed higher calcium ion tolerance.
2013 Elsevier B.V. All rights reserved.

1. Introduction
The formation of sparingly soluble salts in steam generators, boilers,
cooling towers, pipes, tubing and other equipment commonly employed
in water intensive processes is a serious problem, often impairing significantly the overall process and in all cases increasing the cost of production due to the concomitant maintenance cost. Carbonate and sulfate
scaling of alkaline earth metal ions is of particular concern because
these salts, as a rule show inverse solubility, i.e. their solubility decreases
with increasing temperature. Moreover, in the case of calcium carbonate
polymorphism is a complicating factor [1]. Scale deposits, according to
their mode of formation, may be distinguished in two categories: Salts
depositing more or less selectively onto the surfaces of the equipment
in contact with the aqueous uids (usually at elevated temperature)
and precipitates accumulating because of sedimentation or transport

Corresponding author.
0011-9164/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.desal.2013.12.012

by uid ow. As a rule, in this latter case the deposits are formed in
the bulk spontaneously due to the increase of the solution supersaturation, or they form as corrosion by-products which at a second stage sediment out. Among the various strategies adopted to retard or prevent
scaling is the use of chemical additives [2,3] which, depending on their
chemistry and the nature of the solids forming, inhibit nucleation
(threshold inhibitors [4]), crystal growth or both [5]. Scale deposits are
in most cases crystalline, with a surface charge directly or indirectly
pH dependent. The chemical additives through the ionization of their
functional groups, cause changes in the distribution of electrical charges
on the surface of the suspended particles, thus determining the stability
of the respective suspensions.
Polymaleic acid polymers have been shown to inhibit the formation
of scale at very low concentrations [6]. A problem which may be limiting
the efciency of the water soluble polymers as scale retardants is the
formation of salts with alkaline earth metal ions or the formation of micelles if the respective concentration in solution exceeds the critical micellization concentration.

56

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

In the present work, the effects of various polymers and copolymers


containing different functional groups (carboxyl, sulfono, and amide)
were tested with respect to their ability to inhibit calcium carbonate
(pH dependent formation [79]) and calcium sulfate (pH independent
precipitation in the range of 310 [1012]). Since it is established that
for the most part inhibitors interfere with the nucleation and crystal
growth processes through adsorption on the crystals forming from supersaturated solutions, it is anticipated that their presence in the aqueous solutions shall cause changes in the distribution of the surface
charge of the particles thus affecting secondary processes such as aggregation. Experiments were therefore conducted to evaluate the effect of
the presence of polymers in solution on the stability of hematite dispersions. The surface of hematite has a pH dependent charge and the pH
dependent adsorption of water soluble macromolecules is expected to
affect the stability of the respective suspensions.
2. Experimental methods
2.1. Materials
Grade A glassware and analytical grade chemicals were used. Stock
solutions of calcium chloride, sodium sulfate, sodium bicarbonate and
sodium carbonate were prepared from the respective crystalline solids
(Merck) using distilled water, ltered through a 0.22 m lter paper
and standardized as described previously [13]. Iron oxide (Fe2O3)

used in this investigation was obtained from Fisher Scientic Co.


Powder X-ray diffraction showed that it consisted exclusively of
-Fe2O3(hematite). The polymers tested were laboratory preparations
and commercial materials. Polymer stock solutions were prepared on
a dry weight basis. The desired concentrations of the polymers were obtained by dilution. Table 1 lists the polymers tested.
2.2. Inhibitor evaluation protocols
2.2.1. Calcium sulfate dihydrate (CaSO4.2H2O, gypsum) nucleation and
crystal growth
Supersaturated solutions of calcium sulfate for precipitation experiments were prepared by adding a known volume of stock solutions of
sodium sulfate and inhibitor solutions to glass bottles containing
known volume of distilled water maintained at 66 C. Following
temperature equilibration, a known volume of calcium chloride stock
solution was added in such amount that the nal CaSO4 solution concentration was 45.0 mM. The total volume of the supersaturated solutions containing varying concentrations of the test inhibitors was
100 mL. Precipitation in these solutions was monitored by analyzing
aliquots of the ltered (0.22 m lter paper) solution for calcium by
EDTA titrations. The pH value of the calcium sulfate supersaturated solution was adjusted to 7.00 0.05 by the addition of standard HCl
and/or NaOH solutions as needed. At this pH value the additives tested
were fully ionized.

Table 1
Polymers investigated as scale inhibitors in calcium sulfate and calcium carbonate supersaturated solutions and as dispersion agents for Fe2O3 in water.
Polymer

MW

Abbreviation

poly(maleic acid)

Repeat unit

b1 k

HP1

poly(vinyl pyrrolidone)

15 k

HP2

poly(acrylamide)

~6 k

HP3

poly(acrylic acid)

~6 k

HP4

poly(acrylic acid)

~2 k

HP5

poly(methacrylic acid)

~6 k

HP6

poly(methacrylic acid)

30 k

HP7

poly(2-acrylamido 2-methylpropane sulfonic acid)

~10 k

HP8

Sodium salt of maleic acid alt-N vilylpyrrolidone copolymer (AP 28.2)

15 k

CP1

Na/Zn salt of maleic acid-alt-N-vilylpyrrolidone copolymer (AP 28.52.1)

60 k

CP2

Sodium salt of maleic acid-N-vinylpyrrolidone-alt-vinyl acetate ternary copolymer (ATP 6.2)

15 k

CP3

poly(maleic acid:sulfonated styrene)

b10 k

CP4

poly(acrylic acid:2-acrylamido-2-methylpropane sulfonic acid)

b15 k

CP5

poly(acrylic acid:acrylamide)

~10 k

CP6

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

2.2.2. Calcium carbonate


Supersaturated solutions of calcium carbonate for precipitation experiments were prepared by adding a known volume of stock solutions
of sodium bicarbonate (0.258 M), sodium carbonate (0.0029 M) and inhibitor solutions to glass bottles containing known volume of distilled
water maintained at 66 C. The pH of the solutions at 25 C was about
8.2, typically found in alkaline carbonate scale deposits. Following temperature equilibration, a known volume (2.50 mL) of calcium chloride
stock (0.225 M) solution was added. The total volume of calcium carbonate supersaturated solutions containing varying concentrations of
polymer was 100 mL. Precipitation in these solutions was monitored
by analyzing aliquots of the ltered (0.22 m lter paper) solution for
calcium by EDTA titration.
2.2.3. Inhibitor performance
The driving force for the precipitation of salts from supersaturated
solutions depends on the saturation index, SI, dened as:
SI

IAP
K0s

where IAP is the ion activity product and K0s is the thermodynamic solubility product of the salts considered. For the experiments done in the
present work the SI was 0.48 103, 1.07 103 and 1.35 103 for
vaterite, aragonite and calcite polymorphs of calcium carbonate respectively and 1.78 and 2.09 for calcium sulfate anhydrite and gypsum respectively. It should be noted that in the case of calcium sulfate
supersaturated solutions, it was only gypsum that was found in the precipitated solids.
The performance of the tested compounds as calcium sulfate and
calcium carbonate inhibitors was calculated using Eq. (2):
Percent Inhibition %I 100 

Caexp Cafinal
Caini Cafinal

where:
(Ca)exp:
(Ca)nal:
(Ca)ini:

Concentration of calcium in the ltrate in the presence of an


inhibitor at 20 h.
Concentration of calcium in the ltrate in the absence of an
inhibitor at 20 h.
Concentration of calcium at the beginning of the experiment.

At the end of experiments solid samples of precipitates were collected


for characterization by powder X-ray diffraction (XRD) and by scanning
electron microscopy (SEM).
2.2.4. Calcium-inhibitor salt
The CaP (calcium inhibitor) salt precipitation experiments were
performed in a glass bottle (125 mL capacity) placed in a doublewalled glass cell maintained at 45 C. The test solutions (total volume
100 mL) were prepared by adding known volumes of stock solution
(0.1%) of polymer (P) to a known volume of water in the glass bottle.
After allowing the polymer solution to equilibrate at the experimental
temperature for at least 30 min, the solution pH was adjusted to 9.0
using dilute solutions of standard HCl and/or NaOH. This pH value was
sufciently high to ensure full ionization of the functional groups of
the polymers but lower than the value in which Ca(OH)2 precipitate
could form. Next, 5.0 mL of 0.25 M CaCl2 solution was added to the
polymer solution. The solution pH was quickly re-adjusted to the required value with the addition of a standard solution of NaOH as needed.
The bottles were capped and continuously stirred with stirring bars and
magnetic stirrers. At known time intervals (typically 30 min) transmittance readings were taken using a ber optic probe. Duplicate/triplicate
experiments were run to check data reproducibility. In order to avoid

57

faulty signal, extreme care was taken to eliminate air bubbles in the solutions, especially in the vicinity of the ber optic probe.
A known amount (0.12 g) of iron oxide was suspended in an 800 mL
beaker containing 600 mL of simulated industrial water containing
known polymer concentration (dispersant). The simulated industrial
water was prepared by mixing accurately measured volumes of standard solutions of CaCl2, MgCl2, Na2SO4, and NaHCO3. The composition
of the simulated industrial water was: 100 mg/L Ca, 30 mg/L Mg,
314 mg/L Na, 571 mg/L Cl, 192 mg/L SO4, and 60 mg/L HCO3. The pH
of the simulated water was 7.67.8. All dispersion experiments were
done at room temperature (~22 C).
In a typical test, six experiments were run simultaneously using a
gang-stirrer at 110 rpm (supercial uid velocity 1.22 m s 1, or for
the experimental setup the Reynolds number was 9778 [14]). At
known time intervals transmittance readings (%T) were taken with a
Brinkmann Probe Colorimeter equipped with a 420 nm lter. The absorbance of several ltered (0.22 m) suspensions with low to high %T
was measured at 420 nm. It was found that absorbance contribution
due to dissolved species was insignicant (b3%). Polymer performance
as percent iron oxide dispersed (%D) was calculated, after making a correction for the %T reading obtained in the absence of polymer (90%T),
from %T readings (%D = 100 1.1%T) measured past 3 h from the
preparation of the solutions and was expressed as the amount of iron
oxide dispersed. The data presented in this study were reproducible
(5% or better). The performance of the polymer was determined by
comparing the %D values of the slurries containing polymers against
control (no polymer). Higher %D values suggested that the dispersion
was more effective. The following ranking was assigned for dispersant
performance: Poor (b 25%D), Mediocre (b50%D), Good (b75%D), and
Excellent (N85%).
2.2.5. Crystal characterization
The solids precipitated, following air drying at 40 C were characterized by scanning electron microscopy (SEM, Leo Supra 35VP, with
Bruker AXS EDS analyzer) and powder X-ray diffraction (Siemens
D5000 diffractometer). The sample examinations by SEM were done
on aluminum slabs sputtered with gold. The FTIR spectra over the
wave number range 4000400 cm1 were recorded using a FTIR spectrometer (Digilab Excalibur, Randolph, MA, USA).
In domestic and industrial applications (i.e. laundry, cleaners, water
treatment, desalination, geothermal, oil eld, etc.) polymers are used for
a variety of reasons but most importantly they are employed to inhibit
the formation of scale forming salts and in order to keep particulate
matter dispersed. Polymers prevent scale formation either by adsorbing
onto crystal growth sites of micro-crystallites thereby interfering with
crystal growth or by retarding or completely suppressing the formation
of the critical nuclei in the supersaturated uid. Adsorption at specic
sites of the growing crystals may result to changes in crystal morphology depending on the slower growing crystal faces.
2.3. Polymer performance
Over the last three decades different types of polymers have been developed and are currently being used as components of water treatment
formulations. The role of these polymers in such formulations is twofold:
(a) to prevent or inhibit the precipitation of scale forming salts such as
CaCO3, Ca3(PO4)2, CaF2, BaSO4, etc., and (b) to disperse the suspended
matter such as clay, iron oxide, organic debris, etc. In the former case,
prevention refers to the nucleation process and the inhibition of the
growth of scale forming salt crystallites, to the slow-down of the crystal
growth of the crystal faces due to the adsorption of the polymeric species
at the active crystal growth sites present in the crystal faces. The polymers that fall in this category are usually low molecular weight homopolymers containing a carboxyl group (\COOH). The functional groups
in the polymer chain are instrumental for the interaction with the scale
deposit nuclei forming. Homopolymers with 1 (HP4) and 2 (HP1)

58

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

carboxyl groups and with methyl substitution in the acryl group (HP6)
were tested. Different MW polyacrylic acids (HP5, MW 2 k and HP4
MW 6 k) were tested as inhibitors. Dispersants commonly used include
copolymers containing different functional groups i.e., \COOH,
\CONHR, \SO3H, \COOR, etc. These polymers function by adsorbing
onto particles such as iron oxide, silt, etc., present in recirculating waters
and preventing them from settling on the equipment surfaces. Polyamides (HP3) and polyamides substituted with acidic sulfono groups
(HP8) and pyrrolidone substituted poly acrylates (HP2) were also tested.
Finally, co-polymers, which at the molecular level may combine the
properties of both carboxylic acid scale inhibitors with the amide
group containing dispersants were tested. Co-polymers CP1 and CP2
are combinations of HP1 with HP2 (at the monomer level) with the additional presence of zinc ions in CP2 and co-polymer CP3 has a similar repeating unit as CP1 containing however an additional ester group
(acrylic methylester). In co-polymer CP4 the repeating unit is HP1 with
an acrylic unit containing a benzosulfonic acidic group, expected to be
ionized over a very wide range of pH values. Co-polymers CP5 and CP6
have repeating units made of HP4 and HP8 and HP3 and HP4 respectively. The polymers and copolymers tested included functional groups expected to function as calcium carbonate and calcium sulfate dihydrate
(gypsum) scale inhibitors and iron oxide dispersants.
2.4. Inhibition of gypsum crystal growth
In cooling water and desalination of brackish water by reverse osmosis, calcium sulfate dihydrate is the most commonly encountered
calcium sulfate scales [15,16] whereas calcium sulfate hemihydrate
(CaSO4. H2O, plaster of Paris), and calcium sulfate anhydrite (CaSO4)
are the most frequently salts encountered in high temperature processes (i.e., multistage ash distillation, boiler, geothermal) [17]. Several
methods have been applied for the control of scale formation including
the use of acids [18], chelants [19], or crystal growth inhibitors [20]. The
disadvantage of using acids is that at low pH values corrosion is enhanced. The reason for the use of chelants and crystal growth inhibitors
some of which are added at high concentrations is that due to the
specicity of the compounds and the relatively large amounts needed
for effective scale control, the cost of scale prevention increases drastically. The most promising scale control method involves adding substoichiometric dosages, typically a few parts per million (ppm), of
water soluble additives to the feed water. The small quantities needed
in this case, even if the chemicals are expensive, make the overall approach economically feasible.

2.4.2. Polymer architecture


In view of above results suggesting that polymers containing carboxylic acid (COOH) groups inhibit gypsum scale formation, additional
experiments were carried out with copolymers containing different
functional groups i.e., ester (\COOR), sulfonic acid (\SO3H) and
amide, (\CONH2). Fig. 2 shows the results of the % inhibition of crystal
growth of gypsum from supersaturated solutions in the presence of copolymers HP1, HP2, HP3, HP4, HP8, CP1, CP3, CP4, CP5 and CP6.
As may be seen, all copolymers tested showed less inhibition in comparison with poly(maleic acid), HP1, and poly(acrylic acid), HP4, suggesting that the replacement of \COOH groups with amido groups
and/or the introduction of pyrrolidone in the polymer chain resulted
in decreased performance of the copolymers as gypsum inhibitors

100

100

80

80

60
HP1
CP1

40

HP2
CP2

20
0
0.0

% Inhibition

% Inhibition

2.4.1. Polymer dosage


To evaluate the efcacy of homopolymers as gypsum scale inhibitors, a series of precipitation experiments were carried out in the presence of varying concentrations of homo-polymers such as poly(maleic

acid), HP1, poly(vinyl pyrrolidone), HP2, and two copolymers of maleic


acid:vinyl pyrrolidone, CP1 and CP2 respectively. The structures of these
polymers are shown in Table 1.
Fig. 1 presents percent inhibition values (%I) for gypsum precipitation calculated according to Eq. (2) for HP1, HP2, CP1 and CP2.
There are two points worth noting in Fig. 1: (a) %I value increases
with increasing polymer concentrations and (b) %I value depends on
polymer architecture. Based on the data presented in Fig. 1 it is clear
that under similar experimental conditions poly(maleic acid), HP1, containing two \COOH groups on adjacent carbons performs better than
CP1, a copolymer of maleic acid:vinyl pyrrrolidone (MA:VP). In the molecular unit of the latter, although there are two \COOH on adjacent
carbon, one of these groups is sterically hindered by the presence of
the pyrrole unit. Thus, %I values obtained in the presence of 2.0 ppm
of polymers are 83% for HP1 compared to 26% for CP1. Moreover,
the superior performance of HP1 over CP1 may be attributed to:
(a) low molecular weight (MW) for HP1 (b 1 k) versus high MW
(15 k) for CP1 and (b) the presence of a non-ionic group in CP1 (i.e.,
vinyl pyrrrolidone). Making the widely accepted assumption that the
inhibition of gypsum precipitation involves the adsorption and/or interactions of the inhibitor \COOH groups with Ca2+ of gypsum [21] crystallites, the presence of non-ionic and bulkier pyrrole ring in CP1
apparently reduced the interactions of the two \COOH ionized groups
with calcium ions on the surface of the gypsum crystallites. The net effect is the relatively poorer performance of CP1 as a gypsum inhibitor
in comparison with HP1. In the case of CP2, which performed better as
an inhibitor of gypsum precipitation, it may be assumed that the predominant factor was the presence of Zn2+. In view of its low concentration however it may be assumed that this ion has a synergistic effect
with the polymer. Polymer HP2 showed practically no inhibitory activity. As may be seen from the respective chemical structure, in which the
ionizable carboxyl group has been replaced by the pyrrolidone group, it
is not expected that there shall be a signicant electrostatic interaction
of this polymer with the surface of gypsum.

60
40
20
0

2.0

4.0

6.0

8.0

10.0

Polymer, ppm
Fig. 1. Inhibition of gypsum precipitation in the presence of varying dosages of homo- and
copolymers; 66 C; SIgypsum = 2.09.

HP1 HP2 HP3 HP4 HP8

Homopolymers

CP1 CP3 CP4 CP5 CP6

Copolymers

Fig. 2. Inhibition of gypsum precipitation in the presence of 2.0 ppm of homo- and
copolymers; 66 C; SIgypsum = 2.09.

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

even in the case of inclusion of the highly ionizable sulfonic group in the
amides. For example, introduction in the maleic acid chain in HP1 vinyl
pyrrolidone, methacrylic ester or sulfonated styrene resulted in the copolymers CP1, CP3, and CP4. As shown in Fig. 2 these copolymers
showed lower performance as gypsum inhibitors, the worse being
CP4. It may be speculated that the rigidity of the aromatic ring makes
the interaction of this molecule with the gypsum unfavorable. In CP5,
in which the sulfonic acid is attached to the amidic group the molecule
is more exible resulting in stronger interaction with the substrate and
hence stronger inhibition. Moreover, the presence of the amidic nitrogen may favor the formation of hydrogen bonds with the surface sulfate
groups (\S\OHN), thus counterbalancing the lack of carboxyl
group. Hydrogen bonding has been suggested to enhance the inhibitory
activity of certain classes of compounds [22]. It is worth noting that although sulfonic acid containing copolymers (i.e., CP4, CP5) show good
to mediocre performance as gypsum inhibitors, these copolymers
have been found to be excellent calcium phosphate inhibitors [23] and
iron oxide dispersants [24] for industrial water systems. The performance as inhibitors of calcium phosphates (mainly apatitic) may be attributed to stronger interaction with surfaces containing phosphate and
hydroxyl groups. The ability to stabilize iron oxide particles in suspension is related with the anticipated changes of the charge distribution
of the electrical double layer of the particles considering the presence
of acidic ionizable groups (both carboxyl and sulfonic).

2.4.3. Polymer molecular weight


In recent years, a number of studies have been undertaken
concerning the inuence of molecular weight (MW) on the performance of polymers for water treatment applications. Studies of polymers as scale inhibitors have shown that polymer performance in
industrial water treatment applications is strongly affected by the MW
of the polymer present in the aqueous uid. Amjad, in studies on the
evaluation of poly(acrylic acids) of varying MW in controlling the precipitation of BaSO4 [25], CaSO4.2H2O [26], and calcium phosphonate
[27] from aqueous solutions, reported that the MW of the polymer
present in the respective supersaturated solutions, plays an important
role in the inhibition of various scale forming salts. Results of these studies indicated that especially for PAA, the optimum performance was obtained with preparations with MW ca. 2000. Apparently, the low MW
chains are more effectively adsorbed at on the surface of the crystalline
material through strong Ca2+\COO interactions. Signicantly higher
molecular weight of the polymers may change the mode of adsorption
allowing the conformation of the polymer on the crystals to form
loops and tails [28], thus blocking less effectively the active growth
sites. For carboxylic acid containing co- and terpolymers, it appears

59

that precipitation inhibition is greatest for MWs of below 20,000 with


the optimum MW being dependent on the particular polymer composition and the scale formed [2830].
Excellent examples of the MW effect exhibited by poly(acrylic acids)
HP4 (MW 6 k), HP5 (MW 2 k) poly(methacrylic acid) HP6 (MW 6 k),
HP7 (MW 30 k) and copolymers, i.e. CP1 (ME 15 k), CP2 (MW 60 k)
and the respective results on calcium sulfate scale inhibition are presented in Fig. 3.
As may be seen, the polymer performance as gypsum inhibitor decreased with increasing MW for both homo- and copolymers. The better
performance of HP5 (MW 2 k) in comparison to HP4 (MW ca. 6 k) was
anticipated and may be explained perhaps by the mode of interaction
between the polymers with the newly forming gypsum nuclei. HP6,
having the same MW with HP4 but differing by the fact that one of
the hydrogen atoms of the acrylic monomer has been replaced by a
methyl group, which appears to reduce the interaction of the polymer
with the mineral phase forming. The relatively poor performance
shown by high MW polymers may also be attributed to steric hindrance
by the bulkier groups present in copolymers which may result in higher
adsorption entropy for these polymers on gypsum crystallites.

2.5. Calcium carbonate inhibition


Calcium carbonate (CaCO3) is one of the most commonly encountered scale deposits. A special feature of calcium carbonate is polymorphism. It occurs in different crystalline forms in the order of decreasing
stability: calcite, aragonite, and vaterite. Moreover, two hydrated forms
calcium carbonate monohydrate and calcium carbonate hexahydrate
have been reported to form at specic conditions [1,3134]. The precipitation and stabilization of these polymorphs depend upon the precipitation conditions i.e., degree of supersaturation, pH, ionic medium, and
concentration and type of impurities [35,36]. The precipitation and deposition of CaCO3 on equipment surfaces continue to pose serious operational problems in industrial water systems. The formation of these
deposits reduces heat transfer and internal diameter of pipes, increases
operating pressure of pumps, and enhances the probability of corrosion
damage [37,38]. During the last two decades investigations of inhibitors
to prevent or retard CaCO3 scaling have attracted the attention of academic and industrial researchers. Common inhibitors of CaCO3 include
polyphosphates, phosphonates, and acrylic acid/maleic acid based
homo- and copolymers [3942]. Investigations of these and several
other mineral scale inhibitors suggest that inhibitor effectiveness depends upon the functional groups in the inhibitor molecule, polymer
composition, and the molecular weight [43]. In the experiments of the
present work the precipitates consisted of calcite with small amounts

100

100
2k

80

% Inhibition

80

% Inhibition

6k

60
40
6k

20

HP1

40

HP2
CP2

20

15k

30k

60

30k

0
0.0

2.5

5.0

7.5

10.0

Polymer, ppm

0
HP5

HP4

HP6

HP7

CP1

CP2

Fig. 3. Inhibition of gypsum precipitation in the presence of 2.0 ppm of homo- and copolymers of varying molecular weight; 66 C; SIgypsum = 2.09.

Fig. 4. Inhibition of the precipitation of calcium carbonate in the presence of varying concentration of homo- and copolymers; 66 C; SIvaterite = 0.48 103; SIaragonite = 1.07
103; SIcalcite = 1.35 103.

60

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

of aragonite and traces of vaterite, suggesting that the initially formed


precipitate was vaterite, which was transformed rapidly to calcite, as
the supersaturation was rapidly reduced.
The effectiveness of several homo- and copolymers containing different functional groups was investigated according to the procedure
described above. Fig. 4 presents the inhibition results obtained for the
precipitation of calcium carbonate in the presence of HP1, HP2, and CP2.
It is evident that HP1, compared to CP2, which differs (aside from the
difference in MW) chemically in that in the latter the monomer contains
an additional pyrrolidone group, performed much better in inhibiting
CaCO3 formation. HP1 is still superior to CP2 in the inhibition of CaCO3
formation from supersaturated solutions. The lack of signicant inhibition of calcium carbonate precipitation from its supersaturated solutions in the presence of HP2, may be attributed to the hydrophobic
character of this molecule. Moreover, as may be seen from the results
shown in Fig. 4, the incorporation of the relatively more hydrophobic
pyrrolidone functional group in CP2 decreased the respective performance as a CaCO3 carbonate inhibitor in comparison with HP1, a fact
that could be attributed to the increase in molecular weight of the polymer in CP2, despite the fact of the presence of Zn2+, which in the case of
calcium sulfate showed a synergistic effect in the inhibition of precipitation of the respective solid. In this case, apparently the MW effect is predominant. It is interesting to note that results presented in Fig. 2 for
homo- and copolymers are consistent with the results obtained for gypsum inhibition thus indicating that presence of \COOH in polymer
plays an important role in preventing the precipitation of CaCO3. The effect of polymer molecular weight (including polymers HP5, HP6 and
HP7) and also of methacrylic on CaCO3 inhibition has been reported
elsewhere [44].
Inhibition data presented in Fig. 5 clearly show that polymer performance strongly depends upon the ionic charge of the functional group
present in the polymer. The relative performance of HP3, HP4 and HP8,
shown in Fig. 5 suggests that the presence of the ionizable \COOH is
very important for the inhibitory activity of the molecules. Amidic or
sulfono groups attached to amidic groups are obviously not important
in terms of specicity for the inhibition of calcium carbonate.
The stronger effect of HP1 may be attributed both to the presence of
two carboxyl groups which may chelate with surface Ca2+ ions of the
nucleating calcium carbonate and to its signicantly lower MW. It is
also evident from Fig. 5 that polymers containing amidic (HP8)
and pyrrolidone (HP2) functional groups are poor performers for
CaCO3 inhibition. Furthermore, all copolymers evaluated show good to
mediocre performance in inhibiting CaCO3 from supersaturated solutions. It should be noted that the performance trend observed among
various polymers in the case of CaCO3 is consistent with the observation

made in gypsum inhibition (Fig. 2). This nding corroborates the model
assumed for the inhibitorcrystal surface interaction.
2.6. Iron oxide dispersion
The fouling of heat exchangers and reverse osmosis membranes by
suspended matter (i.e., clay, silt, silica, biomass, carbonate and sulfate
salts of alkaline earth metal, corrosion products, etc.) is a critical concern
to water technologists and plant operators. Certain feed waters, especially surface waters require far more extensive pretreatment than
sources such as deep wells. Changes in feed water composition can
occur because of seasonal variations of the water supply. Cooling
water is generally contaminated with various forms of oxidized iron
due to corrosion of steel equipment, and/or its introduction with the
feed water. Maintaining this oxidized iron in soluble and/or in dispersed
forms can possibly prevent heat exchanger surfaces fouling. Although
iron-based deposits are common in industrial water systems (e.g. as
corrosion by-products), no predictive models are available as in the
case of calcium carbonate, calcium sulfate, or barium sulfate. Hematite
(Fe2O3) and magnetite (Fe3O4) are the two most common iron oxide
deposits encountered in industrial water systems.
The suspended particles typically encountered in industrial water
systems (in the pH range 79) generally carry a slightly negative charge
[45]. Therefore, anionic polymers are normally the most effective dispersants because they increase negative surface charge and help keep
particles suspended. Cationic polymers can be used as dispersants, but
this requires relatively high polymer concentrations in order to rst
neutralize the negative surface charges and then to transfer cationic
charge to particles for efcient dispersion.
2.6.1. Dispersion time
The results showing the performance of CP3 as an iron oxide dispersant at varying polymer dosage and as a function of time and polymer
dosage are presented in Fig. 6.
Two points are worth noting in Fig. 6: (a) %D value increases with increasing time and (b) % D value increases with increasing polymer dosages. For example, %D values obtained in the presence of 1.0 ppm at
hr and 1 h are 18 and 30%, respectively. It is evident from Fig. 6 that
increasing the dispersion time by a factor of 3 (i.e., from 1 to 3 h) results
in ~13% increase (from 30 to 43%) in %D value. As noted in Fig. 6 further
increase of the time of suspension (i.e., from 3 h to 4 h) did not yield signicant increase of the %D value. Thus, it is clear from Fig. 6 that dispersion time (i.e., polymer contact time with iron oxide particles) plays an
important role in dispersing iron oxide particles in aqueous solution. It

100

60
50

% Dispersed

% Inhibition

80
60
40
20

40
0.25 ppm

30

0.5 ppm
1 ppm

20

2 ppm

10

0
HP1 HP2 HP3 HP4 HP8

Homopolymer

CP1 CP2 CP3 CP4 CP5 CP6

Copolymers

0
0

50

100

150

200

250

Time, min
Fig. 5. Inhibition of the precipitation of calcium carbonate in the presence of varying
concentration of 3.0 ppm of homo- and copolymers; 66 C; SIvaterite = 0.48 103;
SIaragonite = 1.07 103; SIcalcite = 1.35 103.

Fig. 6. Iron oxide dispersion in the presence of various concentrations of CP3 and as a function of time; room temperature.

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

2.6.2. Polymer dosage


The effect of polymer dosage on iron oxide dispersion was also studied. Results presented in Fig. 6 show that CP3 performance strongly depends on polymer dosages. For example, %D values obtained at 3 h in
the presence of 0.25 and 0.50 ppm of CP3 are 27 and 36%, respectively.
As may be seen, in all cases the extent of dispersion of iron oxide particles does not depend linearly on the amount of polymer added. Thus, increasing CP3 concentration by 2 fold (i.e., from 0.50 to 1.0 ppm)
resulted in only an ~ 20% increase (from 36 to 43%) in the %D value.
This is not surprising, if we take into account the fact that the extent
of adsorption depends largely on the conformation of the polymers on
the particles. More detailed studies on the mechanism of adsorption
through the respective adsorption isotherms are needed to obtain a better understanding of this behavior of the polymers. The data presented
in Fig. 6 clearly show that polymer performance depends on both the
dispersion time and the dispersant concentration. The time dependence
of the dispersion in the presence of the tested polymers is an indication
of the rather slow kinetics of adsorption of the polymers on the
suspected particles with the concomitant changes of the distribution
of charges in their electrical double layer.
2.6.3. Homopolymer performance
The inuence of functional groups (i.e., \COOH, \CONH2,
\CO\N\C, \SO3H) present in homopolymers was studied by
conducting a series of dispersion experiments in the presence of
1.0 ppm of polymers. Dispersion data presented in Fig. 7 clearly show
that the \COOH group containing polymers (i.e., HP1, HP4) exhibits
similar performance as iron oxide dispersants.
As illustrated in Fig. 7 polymers containing neutral monomers, i.e.
with functional groups which may not be ionized in water, such as
HP2 and HP3 were ineffective iron oxide dispersants. It is interesting
to point out that the \SO3H group containing homopolymer, HP8,
showed good performance (72%) as iron oxide dispersant. The signicantly improved performance of HP8 over HP1 or HP4 was attributed
to the more acidic character of \SO3H (pK ca. 3) in comparison with
the acidity of the \COOH group (pK ca. 5). It is interesting to note
that whereas HP8 showed poor performance as gypsum (Fig. 2) and
CaCO3 (Fig. 5) scale inhibitor, it was found to exhibit excellent performance as an iron oxide dispersant. This behavior shows that indeed
the interaction of HP8 with the mineral phases is poor: In the case of nucleation and growth this poor relationship is expected to leave the processes unaffected while from the dispersion point of view, lack of
interaction would allow the high surface potential of the oxide particles

100
80

% Transmittance

should be noted that similar dispersion time dependence of polymer


performance has also been reported for hydroxyapatite (Ca5(PO4)3OH)
[15].

61

60

CP3

40

HP1

CP3

HP2

20
0
0

50

100

150

200

250

300

Polymer, ppm
Fig. 8. Plots of % transmittance vs. polymer concentration for homo- and copolymer for
Fe2O3suspensions in water; room temperature.

to be unchanged, thus facilitating dispersion. Adsorption of polymers in


general reduces the surface potential of suspended oxide particles [28].
2.6.4. Copolymers performance
The performance of a variety acrylic acid and maleic acid copolymers
was investigated in the presence of 1.0 ppm of polymers. Results presented in Fig. 7 clearly show that performance of a copolymer strongly
depends on the ionic character of the monomer functional group. For
example, replacing partly maleic acid monomer in HP1 with a neutral
group containing monomer (i.e., vinyl pyrrolidone) to yield CP1, CP2,
and CP3, does not signicantly impact the performance of HP1. On the
other hand, replacing acrylic acid or maleic acid monomer with sulfonic
acid group containing monomers (i.e., sulfonated styrene, CP4 and 2acryalamido-2-methylpropane sulfonic acid, CP6) improved signicantly the performance of the copolymers. Taking into consideration that
NaHCO3 was used in the water composition, the pH of slurry was
~7.88.0. Moreover, stock solutions (0.1%) of all polymers tested were
adjusted to pH 8, therefore no signicant change in pH is observed.
The pH of several slurries was measured and it was veried that it
remained constant. Further work was done over a wide pH range
(from 5.0 to 9.5) and no signicant dependence on the solution pH
was found.
2.7. Calcium-inhibitor (CaP) salt precipitation
2.7.1. Ca2+ ion compatibility of inhibitors
The tolerance of acrylic acid and maleic acid based homo-and copolymers with calcium ions, was investigated from measurements of the

100
300
>250 ppm

>250 ppm

250

Polymer, ppm

% Dispersed

80
60
40
20

200
150
100
50

0
HP1 HP2 HP3 HP4 HP8

CP1 CP2 CP3 CP4 CP5 CP6

Homopolymers

Copolymers

Fig. 7. Iron oxide dispersion in water in the presence of 1.0 ppm of homo- and copolymers;
room temperature.

0
HP1 HP2 HP3 HP4

CP1 CP2 CP3 CP4 CP5

Homopolymers
Fig. 9. Calcium ion compatibility of various homo- and copolymers.

62

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

turbidity of the respective solutions. Fig. 8 presents the typical % transmittance readings as a function of CP3 concentration. The inection
point in the transmittance-inhibitor prole was used to calculate the
point of onset of turbidity. Fig. 8 shows compatibility data obtained for
the CaCP3 system (500 mg/L Ca, pH 9.0, 45 C).
The reproducibility was satisfactory (7%) as may be seen from the
coincidence of the curves for CP3. The compatibility value calculated for
CP3 was 175 12 ppm polymer/500 mg/L Ca. The compatibility data
for HP1 and HP2 are shown as well in Fig. 8. As may be seen, HP2 is extremely compatible with Ca2+ ions as no noticeable change in %T readings was observed up to concentrations of 250 ppm for the polymer.
Under similar experimental conditions, HP1 showed compatibility
value of 60 ppm/500 mg/L Ca.

and acrylic acid (HP4) show poor compatibility. For example, compatibility values obtained for HP1 and HP4 are 60 ppm and 15 ppm polymer/500 mg/L Ca. The relative poor performance of HP4 over
HP1 may be attributed to the difference in MW (5 k vs., 1 k) and
branching. It is worth noting that under similar experimental conditions, polymers (i.e., HP2, HP3) containing non-ionic group are extremely compatible with Ca2 + ions thus indicating poor interactions
of Ca ions with non-ionic groups present in HP2 and HP3. It should be
noted that the calcium compatibility of the polymers tested is in line
with the respective interactions of these polymers with calcium carbonate surfaces. This effect corroborates further the assumption of surface
interaction of the copolymers with the salts as the mechanism of their
activity.

2.7.2. Ca2+ ion compatibility of homopolymers


Fig. 9 presents compatibility data for a variety of homopolymers
containing different functional groups (i.e., \COOH, \CONH2,
\CON\C\). It can be seen that homopolymers of maleic acid (HP1)

2.7.3. Ca2+ ions compatibility of copolymers


In order to study the impact of monomers containing different functional groups i.e., anionic (i.e., COOH, SO3H, sulfonated styrene), nonionic (i.e., \CONH2, \CON\C), a series of Ca2 + ions tolerance

(a)

(b)

cps/eV
0.30

1.00 * Acquisition 10108

O
0.25

0.20

Au

0.15

Ca

0.10

0.05

0.00
1

10

keV
Fig. 10. (a) Scanning electron micrograph of CaCP3 salt (bar corresponds to 1 m), precipitated past the stability limit. (b) Microanalysis EDX spectrum of the CaCP3 salt(Au peak is due
to gold sputtering of the specimens).

Z. Amjad, P.G. Koutsoukos / Desalination 335 (2014) 5563

200

deposits. At the same time, the polymers tested, both homo- and
copolymers showed satisfactory dispersion ability for iron oxide particles, with the exception of molecules lacking functional groups which
may be ionized. The efciency of the polymers tested with respect to
the dispersion of iron oxide particles increased with concentration and
time. Finally the tolerance of the AA and MA polymers towards calcium
ions in the solutions was minimal. Copolymers in which pyrrole and
sulfonate or amido functional groups were included, showed very
good tolerance towards calcium in solution, irrespective of their MW.
Apparently, it is the presence of the functional groups tested that is responsible for the increase of tolerance towards calcium.

150

Intensity/ a.u.

63

100

50
References

0
10

20

30

40

50

60

20
Fig. 11. Powder X-ray diffraction for the CaCP3 precipitate.

experiments were carried out under similar experimental conditions


(500 mg/L Ca, pH 9.00, 45 C, 30 min). From the data presented in
Fig. 9, it is obvious that Ca2+ ion tolerance strongly depends upon polymer composition.
For example, replacing a portion of MA (maleic acid) in HP1 with
vinyl pyrrolidone (VP, w/non-ionic group), and sulfonated styrene (SS
w/-SO3H, hydrophobic styrene) the resulting copolymers MA:VP (CP1,
CP2) and MA:SS (CP4), respectively, showed higher Ca2+ ions tolerance
by an order of magnitude. For example, Ca2+ ion tolerance values obtained for CP1, CP2, and CP4 were 190 ppm, 165 ppm, and N250 ppm
respectively, compared to 60 ppm obtained for HP1. An excellent example of improved Ca2+ ions tolerance of AA-based copolymers is also presented in Fig. 9 (HP4, CP5). As shown in Fig. 9, partly replacing acrylic
acid in HP4 with 2-acrylamido-2-methylpropane sulfonic acid, results
in an ~four fold increase in Ca ions compatibility of CP5. In Fig. 10, the
morphology of the Ca-inhibitor solid obtained from the interaction of
Ca2+ with CP3 is shown. The morphology was similar with precipitates
formed with the rest of the polymers tested and showed that it
consisted of plate-like nanoparticles.
The XRD spectra of the Ca-inhibitor salts, presented in Fig. 11,
showed mostly amorphous character as may be seen from the lack of
clear reections and the low number of counts obtained. The amorphous character of the calcium-polymer salts may be attributed partly
to the unreacted polymer and partly to the small size of the crystallites
formed.
3. Summary
The results of the present work have shown that acrylic (AA) and
maleic acid (MA) homopolymers are effective inhibitors for the formation of gypsum and calcium carbonate scale deposits. Both the presence
of functional groups and the MW of the polymers played a signicant
role. The presence of freely ionizable, sterically less inhibited carboxyl
groups showed higher inhibitory efciency. The MW of the polymers
tested did not affect signicantly their anti-scaling efciency. However,
higher efciency was obtained for the lower MW polymers, a fact attributed to the stronger interaction with the surface of the calcareous

[1] L. Brecevic, D. Kralj, Croat. Chem. Acta 80 (2007) 467.


[2] W.W. Frenier, M. Ziauddin, Formation, Removal, and Inhibition of Inorganic Scale in
the Oileld Environment, Society of Petroleum Engineers, Richardson, Texas, 2008.
[3] D. Hasson, H. Shemer, A. Sher, Ind. Eng. Chem. Res. 50 (2011) 7601.
[4] G.H. Nancollas, Ceram. Trans. 1A (1998) 1.
[5] M. Lahav, L. Leiserovitz, J. Phys. D. Appl. Phys. 26 (1993) B22.
[6] A. Martinod, M. Euvrard, A. Foissy, A. Neville, Desalination 220 (2008) 345.
[7] L.F. Greenlee, F. Testa, D.F. Lawler, B.D. Freeman, Water Res. 44 (2010) 2957.
[8] J.-H. Bang, Y.N. Jang, W. Kim, K.S. Soy, C.W. Jean, S.C. Choe, S.-W. Lee, S.-J. Park, M.G.
Lee, Chem. Eng. J. 174 (2011) 413.
[9] J.D. Rodriguez-Blanco, S. Shaw, P. Bots, T. Roncal-Herrero, L.G. Benning, J. Alloys
Compd. 536S (2012) S477.
[10] F. Li, J. Liub, G. Yanga, Z. Pana, X. Ni, H. Xua, Q. Huang, J. Crystal, Growth 374 (2013) 31.
[11] C.T. Benatti, C.R. Granhen-Tavares, E. Lenzi, J. Environ. Manag. 90 (2009) 504.
[12] P.G. Klepetsanis, P.G. Koutsoukos, J. Cryst. Growth (1989) 98,480.
[13] Z. Amjad, J.P. Hooley, J. Colloid Interface Sci. 111 (1986) 496.
[14] G. Halsz, B. Gyre, I.M. Jnosi, K.G. Szab, T. Tl, J. Appl. Phys. (2007) 751092.
[15] F. Rahman, Desalination 319 (2013) 79.
[16] B.C. McCool, A. Rahardianto, J. Faria, K. Kovac, D. Lara, Y. Cohen, Desalination 261
(2010) 240.
[17] H.C. Simpson, H.R. Hutchinson, Desalination 2 (1967) 308.
[18] J.S. Gill, Desalination 124 (1999) 43.
[19] J.M. Donohue, In Encyclopedia of Physical Science and Technology, Third edition
Academic Press, 2003. 671697.
[20] J.K. Fink, Petroleum Engineer's Guide to Oil Field Chemicals and Fluids, Elsevier,
Amsterdam, 2012.
[21] M. Oner, in: Z. Amjad (Ed.), The Science and Technology of Industrial Water Treatment, CRC Press, Boca Raton, FL, 2010, pp. 2138.
[22] Z.J. Henneman, G.H. Nancollas, F.H. Ebettino, R. Graham, G. Russel, R.J. Phipps,
J. Biomed. Mater. Res. 85A (2008) 993.
[23] Z. Amjad, Tenside Surf. Det. 48 (2011) 53.
[24] Z. Amjad, Tenside Surf. Det. 43 (2006) 242.
[25] Z. Amjad, Water Treat. 9 (1994) 47.
[26] Z. Amjad, Desalin. Water Treat. 36 (2011) 270.
[27] Z. Amjad, R.W. Zuhl, J.A.T. Wholever, in: Z. Amjad (Ed.), in Advances in Crystal
Growth Technologies, Kluwer Academic\Plenum, New York, 2002, pp. 7184.
[28] J. Lyklema, Fundamentals of Interface and Colloid Science, Vol. II, Academic Press,
San Diego CA, 1995.
[29] E. Akyol, A. Bozkurt, M. ner, Polym. Adv. Technol. 17 (2006) 58.
[30] M.J. Jenkins, K.L. Harisson, Polym. Adv. Technol. 17 (2006) 474.
[31] E.K. Giannimaras, P.G. Koutsoukos, Langmuir 4 (1988) 855.
[32] H. Hull, A.G. Turnbull, Geochim. Cosmochim. Acta 37 (1973) 683.
[33] K. Kamiya, S. Sakka, K. Terrada, Mater. Res. Bull. 12 (1977) 1905.
[34] H. Ell, H. Roques, Desalination 137 (2001) 177.
[35] W.D. Carlson, Rev. Mineral. 11 (1984) 191.
[36] N. Vdovi, D. Kralj, Colloids Surf. A Physicochem. Eng. Asp. 161 (2000) 499.
[37] T.R. Soror, Open Corros. J. 2 (2009) 45.
[38] D.H. Troup, J.A. Richardson, Werkst. Korros. 29 (1978) 312.
[39] P. Shakkthivel, D. Ramesh, R. Sathiyamoorthi, T. Vasudevan, J. Appl. Polym. Sci. 96
(2005) 1451.
[40] P. Shakkthivel, R. Sathiyamoorthi, T. Vasudevan, Desalination 164 (2004) 111.
[41] D. Kuila, G.A. Blay, R.E. Borjas, S. Hughes, P. Maddox, K. Rice, W. Stansbury, N. Laurel,
J. Appl. Polym. Sci. 73 (1999) 1097.
[42] S. Mishra, I. Patil, Chem. Eng. Technol. 25 (2002) 573.
[43] Z. Amjad (Ed.), Science and Technology of Industrial Water Treatment, IWA
Publishers, 2010.
[44] Chapter 5 in. R. Zuhl, Z. Amjad, The Science and Technology of Industrial Water
Treatment, CRC Press, Boca Raton, 2010.
[45] G.A. Parks, in: W. Stumm (Ed.), Equilibrium Concepts in Natural Water Systems, Advanced Chemistry Series, 67, American Chemical Society, Washington DC, 1967,
pp. 121160.

Das könnte Ihnen auch gefallen