Sie sind auf Seite 1von 31

D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S.

81

4 ION EXCHANGE IN WATER TREATMENT

This chapter was extracted from Ullmann´s Encyclopedia of Industrial Chemistry, for
further details you should inspect the original literature. Also the internet can be
consulted if further special information is required. Bayer´s specialty chemical division
(Lewatit) provides a lot of information on it´s homepage www.ionexchange.com.

4.1 Introduction
Definition and Principles. In ion exchange, ions of a given charge (either cations or
anions) in a solution are adsorbed on a solid material (the ion exchanger) and are replaced
by equivalent quantities of other ions of the same charge released by the solid.
The ion exchanger may be a salt, acid, or base in solid form, that is insoluble in water but
hydrated. Exchange reactions take place in the water, retained by the ion exchanger; this
is generally termed swelling water or gel water. The water content of the apparently dry
material may constitute more than 50% of its total mass.
Figure (1) shows the partial structure of a cation exchanger; each positive or negative ion
is surrounded by water molecules.

+ +
Fig.1: Structure of a cation exchanger that exchanges H for Na ions; swelling water is
represented in the insert (top left).

Ion exchange forms the basis of a large number of chemical processes which can be
divided into three main categories: substitution, separation, and removal of ions.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 82

1) Substitution. A valuable ion (e.g., copper) can be recovered from solution and
replaced by a valueless one. Similarly, a toxic ion (e.g., cyanide) can be removed
from solution and replaced by a nontoxic ion.
2) Separation. A solution containing a number of different ions passes through a
column containing beads of an ion-exchange resin. The ions are separated and
emerge in order of their increasing affinity for the resin.
+
3) Removal. By using a combination of a cation resin (in the H form) and an anion
– + –
resin (in the OH form), all ions are removed and replaced by water (H OH ). The
solution is thus demineralized.

Historical Aspects. The discovery of ion exchange dates from the middle of
the nineteenth century when THOMSON [1] and WAY [2] noticed that
ammonium sulfate was transformed into calcium sulfate after percolation
through a tube filled with soil.
In 1905, GANS [3] softened water for the first time by passing it through a
column of sodium aluminosilicate that could be regenerated with sodium
chloride solution. In 1935, LIEBKNECHT [4] and SMIT [5] discovered that certain
types of coal could be sulfonated to give a chemically and mechanically stable
cation exchanger. In addition, ADAMS and HOLMES [6] produced the first
synthetic cation and anion exchangers by polycondensation of phenol with
formaldehyde and a polyamine, respectively. Demineralization then became
possible. At present, aluminosilicates and phenol–formaldehyde resins are
reserved for special applications and sulfonated coal has been replaced by
sulfonated polystyrene.

4.2 Structures of Ion-Exchange Resins


An ion exchanger consists of the polymer matrix and the functional groups that interact
with the ions. In this chapter we only deal with organic ion exchangers; inorganic ion
exchangers are of minor importance and are primarily layer silicates and zeolites.

4.2.1 Polymer Matrices


Polystyrene Matrix The polymerization of styrene
(vinylbenzene) yields linear polystyrene. Linear
polystyrene is a clear moldable plastic and is soluble in
certain solvents (e.g., styrene or toluene). If a proportion
of divinylbenzene is mixed with styrene, the resultant
polymer becomes cross-linked and is then completely
insoluble.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 83

In the manufacture of ion-exchange resins, polymerization generally occurs in suspension.


Monomer droplets are formed in water and, upon completion of the polymerization
process, become hard spherical beads of the polymer.

Polyacrylic Matrix. Matrices for ion exchangers can also be obtained by polymerizing an
acrylate, a methacrylate, or an acrylonitrile, any of which can be cross-linked with
divinylbenzene, DVB;

Other Types of matrix. Other types of matrix include


1) phenol – formaldehyde resins which show interesting adsorption properties; and
2) polyalkylamine resins, obtained from polyamines by condensation with
epichlorohydrin, which gives an anion exchanger directly in a single step.

4.2.2 Functional Groups

Cation-Exchange Resins
Cation-exchange resins in current use can be separated into two classes according to their
active groups:
1) strongly acidic (sulfonic groups) and
2) weakly acidic (carboxylic groups).

Strongly Acidic Cation-Exchange Resins. Chemically inert polystyrene beads are treated
with concentrated sulfuric or chlorosulfonic acid to give cross-linked polystyrene 3-sulfonic
acid. This material is the most widely used cation-exchange resin and is strongly acidic .

Examples: Amberlite IR 120, Dowex HCR, Duolite C 20, Lewatit S 100.


D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 84

Weakly Acidic Carboxylic Cation-Exchange Resins. The weakly acidic resins are
almost always obtained by hydrolysis of polymethylacrylate or polyacrylonitrile to give a
poly(acrylic acid) matrix.

Examples: Amberlite IRC 76, Duolite C 433, Relite CC.

Anion-Exchange Resins
Polystyrene Materials. Cross-linked polystyrene is the basis for generating anion exchange
resins carrying amine groups. These amine groups determine the degree of basicity of the
exchanger. The anion exchangers listed below are arranged in order of decreasing basicity :

where R can be
+ –
–CH2N (CH3)3Cl e.g., Duolite A 101 (type 1 resin)
+ –
–CH2N (CH3)2CH2CH2OHCl e.g., Duolite A 102 (type 2 resin)
–CH2N(CH3)2 e.g., Duolite A 378
–CH2NHCH3
–CH2NH2 e.g., Duolite A 365

Resins with quaternary ammonium groups are strongly basic. Those with benzyltrimethyl-
ammonium groups are known as type 1 and are the most strongly basic, whereas those with
benzyldimethylethanolammonium groups are known as type 2 and are slightly less basic.

Type 1 resins are used when total removal of anions, even those of weak acids (including
silica), is essential.

Resins whose active group is an amine are generally denoted as weakly basic, although
their basicity may vary considerably. Tertiary amines are sometimes called medium base
or intermediate base resins, whereas primary amines are very weakly basic and are rarely
used.

The most widely used weakly basic resins contain tertiary amine groups and adsorb any
strong acids present in the solution to be treated but do not affect neutral salts or weak
acids.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 85

Other Types of Ion-Exchange Resins


By using polymerization and activation methods analogous to those described above, a
wide variety of functional groups can be grafted onto a given polymer. Some of these
groups can be used for selective uptake of ions, principally metals (Table 1).
The thiol group forms very stable bonds with certain metals, particularly mercury. The
iminodiacetic, aminophosphonic, and amidoxime groups form metal complexes whose
stability depends mainly on the pH of the solution. Selective adsorption of certain metals
can thus be achieved by varying pH. These types of material are known as chelating or
complexing resins.

Tab. 1: Principal active groups of ion exchangers used for selective uptake of metals

4.2.3 Adsorbent Resins and Inert Polymers


Strictly speaking, adsorbent resins are not ion exchangers but resemble them very closely.
They have a high porosity and are used for the adsorption of nonionic or weakly ionized
species as a complement to ion exchange. In water treatment the most widely used types
are Inert adsorbents (uncharged), which are macroporous copolymers of styrene and
divinylbenzene with a very high degree of cross-linking and a large surface-to-volume
ratio. These resins are used to remove organic, weakly ionized, or nonionic substances,
such as phenols, chlorinated solvents, antibiotics, and complexing agents, from aqueous
or organic solutions.

4.3 Properties
4.3.1 Degree of Cross-Linking and Porosity
An increase in the degree of cross-linking (i.e., the weight percentage of DVB related to
the total amount of monomer prior to polymerization) produces harder, less elastic resins.
Resins with higher degrees of cross-linking show more resistance to oxidizing conditions
that tend to de-crosslink the polymer. Above 10–12% DVB, however, the structure
becomes too hard and dense. Finally, the rate of exchange increases in proportion to the
mobility of the ions inside the exchanger bead: if the structure is too dense, ionic motion
is slowed down, thus reducing the operating capacity of the resin.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 86

Macroporous resins are made by mixing the monomers with a compound (e.g., heptane,
saturated fatty acids, C4– C10 alcohols or polyalcohols, or low molecular mass linear
polystyrene) which expands the resin. The substance does not itself polymerize and, thus,
although it acts as a solvent for the monomers, it causes the polymer to precipitate from the
liquid.

Channels are formed inside the beads, producing an artificially high porosity. Resins
containing such channels are described as macroporous, wherease other resins with
natural porosity are known as gel resins (Fig. (2)).

Fig. 2: Arrangement of structural units in gel (A) and macroporous (B) resins

Macroporous resins have a higher degree of cross-linking than gel resins to strengthen the
matrix and compensate for voids left by the added solvent. The porosity and mechanical
strength of the resin can be modified by varying the degree of cross-linking or the amount
of solvent added. Therefore, various macroporous resins are available, with different
moisture-holding capacities and internal structures.
The pore diameter is ca. 100 nm in a macroporous resin and ca. 1 nm in a gel resin.
Exchange is thus faster in a macroporous resin.

Macroporous resins are highly resistant to physical stress and generally withstand osmotic
shock very well. Finally, macroporous resins are used when reversible uptake of large
molecules is necessary, without fouling the resin.

4.3.2 Exchange Capacity


Total Capacity. The total exchange capacity of a resin, expressed in equivalents per unit
weight (or per unit volume), represents the number of active sites available. The capacity is
expressed in equivalents (eq) per kilogram of dry resin (the weight capacity Cp) or equivalents
per liter of wet settled resin (the volume capacity CV). Total capacity values for some of the
most common resins are given in Table 2.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 87

Tab. 2: Typical capacities of ion-exchange resins *

Operating Capacity. The operating capacity is defined as the proportion of total capacity
used during the exchange process. It can amount to a large or small proportion of the total
capacity and depends on a number of process variables including
1) concentration and type of ions to be absorbed ;
2) rate of percolation;
3) temperature;
4) depth of resin bed; and
5) type, concentration, and quantity of regenerant.

In a packed column, reaction between the ions in solution and those in the resin occurs
over a well-defined region of the resin bed known as the reaction zone.
The longer the column, the deeper is the reaction zone and the greater is the operating
capacity of the resin. Figure (3) illustrates this: the top of the column contains the
completely exhausted resins (a), whereas the reaction zone (b) contains partially
exhausted resin. The time at which the lowest point of this zone reaches the bottom of the
column (i.e., when the adsorbed ions break through the bottom of the bed) is generally
taken to be the time at which the service phase is complete. A proportion of the resin is
still not exhausted at the time of breakthrough.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 88

Fig. 3: Reaction zone in a resin


column during percolation
a) Exhausted resin ;
b) Reaction zone ;
c) Regenerated resin

In practice, strongly acidic and strongly basic resins are never 100 % regenerated at the
beginning of a cycle. The operating capacity thus represents the difference between the
available capacity at the beginning of a cycle and that remaining at the end point. The
most important factor is the amount of regenerant used to convert the resin to the
regenerated form required at the beginning of the service cycle (c).

4.3.3 Stability and Service Life


Because ion-exchange resins must give several years of service, their stability over long
periods of time is of prime importance.

Chemical Stability of the Matrix. Industrially available resins have a degree of cross-
linking high enough to make them insoluble.
Highly oxidizing conditions (presence of chlorine or chromic acid) can attack the matrix
and destroy cross-linking. A sulfonated polystyrene cation-exchange resin with 8% DVB
cross-linking withstands 0.2 mg/kg of chlorine at ambient temperature for several years
and is also completely stable at 120 °C in the absence of oxidants. However, 1 mg/kg of
chlorine oxidizes the polymer at a rate dependent on temperature; this breaks down the
cross-linking, releases sulfonated organic compounds and causes the resin to swell until it
softens, resulting in excessive head loss [13]. When oxidizing agents are present, highly
cross-linked resins with a greater resistance to oxidation, such as the macroporous resins,
should be used.

Degradation products from a cation-exchange resin may foul anion resins [14], [15]. This
is particularly critical in processes designed to produce ultrapure water (Fig. (4)).
Oxidants break the cross-links to produce soluble, short-chain oligomers that can be
measured as the total organic carbon (TOC) in the treated water (Fig. (4)A). Under
normal conditions of water treatment, resins can operate continuously for many years
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 89

(sometimes up to 20 years) without deterioration of their physical and chemical


properties.

Fig. 4: Suitability of resins for producing ultrapure water in mixed beds [14] Resins were
tested at ambient temperature during first chlorine exposure by using an influent with 0.30
ppm active chlorine dosed as NaOCl and 0.02–0.03 ppm TOC.
Resins: a) Gel cation–porous gel anion; b) Gel cation–standard gel anion; c) Macroporous cation–porous
gel anion; d) Macroporous cation–standard gel anion; e) Macroporous cation–macroporous anion; f)
Macroporous cation–developmental anion; g) Gel cation–developmental anion

Thermal Stability of Active Groups [16] . The sulfonic group of cation-exchange resins is
extremely stable. Anion-exchange resins, on the other hand, are temperature-sensitive. When
heated, Hofmann degradation may transform quaternary ammonium groups (strongly basic)
into tertiary amines (weakly basic) or even destroy the active group completely. Because this
reaction occurs under alkaline conditions, anion exchangers are more stable in the form of a
salt than as a base.
Strongly basic type 1 resins are the most stable–the Hofmann degradation reaction becomes
significant only above 50 °C (Fig. (5)). At ambient temperature, these resins can last for five
to seven years or more.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 90

Type 2 resins are more liable to undergo Hofmann degradation because the ethanol
group weakens the bond:

Osmotic Stability. During ion exchange, the configuration around each active group in
the resin changes: the adsorbed ion generally has a different size and, more important, a
different hydration layer than the displaced ion. The resin bead may, therefore, swell or
contract appreciably during the reaction. The stresses to which the resin is subjected
during these volume changes are known as osmotic forces. They are very intense and can
produce local pressures of several thousand kilopascals—much greater than purely
mechanical stress.
Resins for industrial use must be able to withstand hundreds of cycles of exhaustion and
regeneration. The nature and, hence, the strength of osmotic shock vary according to the
ionic species in solution and their concentration.

Resistance to Drying. Repeated drying and rewetting produce stresses analogous to those
due to osmotic shock and can lead to fragmentation of most gel resins. Resins must,
therefore, be kept permanently moist.

4.3.4 Particle Size


For industrial use, particle size is a compromise between the speed of the exchange
reaction (which is greater with small beads) and high flow rates (which require coarse
particles to minimize the head loss). Standard resins contain particles with diameters from
0.3 to 1.2 mm, but coarser or finer grades are available.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 91

4.3.5 Moisture Content


Ion-exchange resins carry both fixed and mobile ions which are always surrounded by
water molecules located in the interior of the resin beads. The water retention capacity
governs the kinetics, exchange capacity, and mechanical strength of ion-exchange resins.
Figure 5 shows how the moisture content varies with the proportion of DVB for gel-type
sulfonic polystyrene resins.

Fig. 5: Variation of moisture content (A) and total capacity (B) with the degree of cross-
linking in a sulfonated polystyrene resin in sodium form

4.4 Ion-Exchange Reactions

4.4.1 Cation Exchange


General cation exchange is used widely to remove undesirable ions from a solution
without changing the total ionic concentration or pH. The resin can be used in many
ionic forms, but the sodium form is usually preferred because the resin has a relatively
low affinity for sodium, which facilitates the adsorption of other metals. Furthermore,
sodium chloride is an inexpensive regenerant.

The following reaction is used to treat wine (R denotes the resin):


– + + – – + + –
R Na + K (HTartrate) → R K + Na (HTartrate)
The reaction used in water softening is
– + 2+ – 2+ +
2 R Na + Ca (HCO3–)2 → (R )2Ca + 2 Na HCO3–
In each case, the resin is regenerated by reversal of the reaction with sodium chloride
solution.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 92

Hydrogen Exchange in Strongly Acidic Resins. The replacement of metallic ions with
hydrogen ions leads to a reduction of the total dissolved solids in solution and the production
of free acid:
– + + – – + + –
R H + Na Cl → R Na + H Cl
This reaction is used as the first stage in the demineralization of water and other solutions. It
is sometimes called salt splitting. Regeneration is carried out with a mineral acid.

Hydrogen Exchange in Weakly Acidic Resins. Carboxylic resins are such weak acids that
they ionize only slightly under acid conditions. However, they have such a high affinity for
divalent metals that in the presence of these metals, they are forced to ionize and therefore
remain active under slightly acidic conditions down to ca. pH 4.5.
2 RCOOH + Mg(OH)2 → (RCOO)2Mg + 2 H2O
2 RCOOH + Na2CO3 → 2 RCOONa + H2O + CO2
RCOOH + NaCl No reaction
2 RCOOH + CaCl2  (RCOO)2Ca + 2 HCl
+
However, H from carboxylic resins cannot remove significant amounts of metals from
solutions of mineral acid salts because the acid produced quickly lowers the pH and prevents
further ion exchange. In this case, ion exchange is controlled by the basicity of the anion in
solution.

Because the resin has a very high affinity for divalent ions (the effect of chelation), but only
moderate affinity for monovalent ions, it has a high capacity for removing calcium and
magnesium from bicarbonate solution but takes up only a small amount of sodium. This
occurs because sufficient carbonic acid is formed to suppress the exchange of monovalent
ions :
RCOOH + NaHCO3 → RCOONa + H2O + CO2
2 RCOOH + Ca(HCO3)2 → (RCOO)2Ca + 2 H2O + 2 CO2
Carboxylic exchangers (weakly acidic) are selective: they preferentially remove divalent or
trivalent cations until competition arises from alkaline anions present in solution.

4.4.2 Anion Exchange


General Anion Exchange. The most widely used resin for general anion exchange is a
strongly basic exchanger in chloride form :
+ – + + + –
R Cl + Na NO3– → R NO3– + Na Cl
This process is used to remove natural organic acids (e.g., humic acid) and nitrate from
water:
+ – + n + n + –
n R Cl + (Na ) Humate – → (R ) Humate – + n Na Cl
n n
+ – + – + – + –
R Cl + Na [Au(CN)2] → R [Au(CN2] + Na Cl
The resins are regenerated by reversal of the reactions with sodium chloride solution.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 93

Acid Absorption in Strongly Basic Resins is the most widely used form of anion
exchange. When it follows hydrogen exchange in strongly acidic resins (Section 4.4.1.
Cation Exchange ), it completes the demineralizing process :
+ – + – + –
R OH + H Cl → R Cl + H2O
+ – +
R OH + CO2 → R HCO3–

Regeneration of the bicarbonate form of the resin requires two equivalents of OH ions

per equivalent of HCO3– taken up, because half the OH neutralizes the bicarbonate and
converts it to carbonate. The same applies to silica.

Acid Absorption in Weakly Basic Resins. Weakly basic resins, in which the active
groups are usually amines, do not have a true hydroxide form. They ionize only under
acidic conditions:
+ – + –
RN(CH3)2 + H Cl → RN H(CH3)2Cl
Under alkaline conditions, they are not in the ionic form and therfore do not exchange
anions (neutral salts are not "split" ):
RN(CH3)2 + NaCl No reaction
+
RN(CH3)2 + CO2 + H2O RN H(CH3)2 + HCO3–
RN(CH3)2 + SiO2 No reaction

Weakly basic exchangers can be regenerated with ammonia or sodium carbonate:


RN(CH3)2 · HCl + NH3 → RN(CH3)2 + NH4Cl
RN(CH3)2 · HCl + Na2CO3 → RN(CH3)2 + NaCl + NaHCO3
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 94

4.4.3 Cation and Anion Exchange in Water Treatment


Figure 6 summarizes the way in which the various forms of ion exchange described in the
previous sections. The composition of raw water is described in detail in Section 4.7.3. Water
Analysis .

Fig. 6: Summary of the kinds of ion exchange used in water treatment


SAC = strongly acidic cation exchangers ; SBA = strongly basic anion exchangers ;
WAC = weakly acidic cation exchangers ; WBA = weakly basic anion exchangers
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 95

4.5 Ion-Exchange Equilibria

4.5.1. Dissociation and pK Value


Dissociation of the acid group in a cation-exchange resin is described by the equilibrium
reaction
. – +
R H ⇔ R +H

where R is the co-ion fixed in the matrix structure. The acidity of the resin is defined by its
degree of dissociation at equilibrium

where K is the equilibrium constant; the quantities in square brackets represent


oncentrations, and underlines indicate the resin phase rather than the aqueous phase.

Fig. 7 shows the titration curves of sstrong and a weak acidic ion exchange resin.

Fig. 7: Titration curves of cation-exchange resins


a) Duolite C 20, sulfonic acid resin ; b) Duolite C 433 carboxylic acid resin

In both cases, the total capacity of the resin can be read from the titration curves and
corresponds to the amount of sodium hydroxide that produces the sharp rise in pH. In Figure
7, Duolite C 20 has a total capacity of 2 meq/mL, whereas C 433 has a capacity >4 meq/mL.
Similar considerations apply to anion-exchange resins.

4.5.2 Mono – Monovalent Exchange

The law of mass action applied to the reaction


– + + – + +
R H + Na R Na + H gives
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 96

+ +
where [Na ] and [H ] are the equivalent concentrations in the liquid phase and γNa and γH
are the corresponding activity coefficients. The concentrations and coefficients for the
resin are indicated by underlines. Thus,

The activity coefficients may be assumed constant, so that

Na
(the + signs are omitted for simplicity). The parameter K H
is known as the selectivity
+ +
coefficient for the Na /H exchange.

Table 4 lists the relative selectivities of sulfonic resins for mono- and divalent cations, and
Table 5 gives the selectivities of strongly basic type 1 and type 2 resins for monovalent
anions. Values increase with the degree of cross-linking and tend to 1 as the cross-linking
tends to zero. Data are only approximate and merely demonstrate the scale of selectivities.

Tab. 4: Relative selectivities of sulfonic resins for cations


Cation Degree of cross-linking, % DVB
4 8 12 16
Monovalent
H* 1.0 1.0 1.0 1.0
Li 0.90 0.85 .81 0.74
Na 1.3 1.5 1.7 1.9
NH4 1.6 1.95 2.3 2.5
K 1.75 2.5 3.05 3.35
Rb 1.9 2.6 3.1 3.4
Cs 2.0 2.7 3.2 3.45
Cu 3.2 5.3 9.5 14.5
Ag 6.0 7.6 12.0 17.0
Divalent
Mn 2.2 2.35 2.5 2.7
Mg 2.4 2.5 2.6 2.8
Fe 2.4 2.55 2.7 2.9
Cu 2.7 2.9 3.1 3.6
Cd 2.8 2.95 3.3 3.95
Ni 2.85 3.0 3.1 3.25
Ca 3.4 3.9 4.6 5.8
Sr 3.85 4.95 6.25 8.1
Hg 5.1 7.2 9.7 14.0
Pb 5.4 7.5 10.1 14.5
Ba 6.15 8.7 11.6 16.5

* Reference value.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 97

Tab. 5: Relative selectivities of quaternary ammonium exchangers for monovalent anions


Anion Resin
Type 1 Type 2
Hydroxide * 1.0 1.0
Benzenesulfonate >500 75
Salicylate 450 65
Iodide 175 17
Phenolate 110 27
Bisulfate 85 15
Chlorate 74 12
Nitrate 65 8
Bromide 50 6
Cyanide 28 3
Bisulfite 27 3
Bromate 27 3
Nitrite 24 3
Chloride 22 2.3
Bicarbonate 6.0 1.2
Iodate 5.5 0.5
Formate 4.6 0.5
Acetate 3.2 0.5
Propionate 2.6 0.3
Fluoride 1.6 0.3
.

4.5.3 Mono – Divalent Exchange (Water Softening)

In the exchange reaction


– + 2+ – 2+ +
2 R Na + Ca ⇔ (R )2Ca + 2 Na
the law of mass action gives

Ca
K Na =
[c Ca ] ∗ [cNa ] ² where [c] are concentrations of adsorbed species and [cNa] and
[c Na ] ² ∗ [cCa ]
[cCa] are ion concentrations in solution. And if X is the fraction of ions of a certain kind
(here Ca or Na) then one can obtain:

The behavior is best illustrated in Fig. 8. The resin takes up more calcium as the total
cation concentration of the solution decreases. Figure 8 shows the equilibrium curves for
a given resin at various total concentrations of the solution.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 98

Fig. 8: Mono – divalent equilibrium curves for Na+ – Ca2+ solutions of different total concentration
+ 2+
Total concentration of Na and Ca , eq/L: a) 0.005 ; b) 0.01 ; c) 0.1 ; d) 1 ; e) 5

Each point on a curve corresponds to the equivalent fraction of calcium in the resin at
2+
equilibrium, i.e., the curve gives the proportion of active sites in the resin in Ca form as
a function of the proportion of calcium in solution. These curves are called ion-exchange
isotherms.

Water Softening. Hardness of natural waters (Ca- or Mg-ions) can be removed by


ion exchange. Let´s consider the following example:
For water with a total salinity of 5 meq/L (conc 0.005 mol/l), Figure 8 shows that
2+
water containing only 5 % Ca (XCa = 0.05 or [Ca] = 0.25 meq/L) is in equilibrium
with a resin loaded with 95 % calcium (X Ca = 0.95). In other words, the resin is
capable of removing calcium from the water even if the concentration accounts for
only 5 % of total cations (XCa = 0.05), as long as the resin contains more than 5 %
sodium ( X Ca < 95 %). At high concentrations, the affinity of the resin for calcium
over sodium decreases until sodium becomes favored. This is the effect that makes
regeneration of the resin possible.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 99

Water Softening

4. 6. Exchange Kinetics

6.1. Principles
Fig. 9: Successive equilibria in the water-softening cycle

Mass action equations apply only to systems in equilibrium. In industrial practice where a
solution flows through the resin, equilibrium is not necessarily reached and the results are
influenced by kinetic considerations.
In fully ionized systems, the rate-determining step of ion exchange is the diffusion of the
mobile ions toward, from, and in the resin phase, rather than the chemical reaction between
fixed ions of the resin and mobile counterions.

Fig. 10: Diffusion through a film and inside a particle


D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 100

+ +
Figure 10 illustrates the uptake of Na ions by a bead of H -form resin. A static layer of
solution, known as the Nernst film, surrounds the bead. This film is defined such that it is
unaffected by convection (i.e., flow) around the bead; ion transport takes place by
diffusion only. Strong convection (i.e., high flow rate) decreases film thickness. The ion
concentration is practically constant outside the Nernst film, and a concentration gradient
occurs within it (Fig. 10).

Diffusion through the film and in the solid phase occurs at different rates and two steps
may be rate-determining:
1) Diffusion of ions within the resin (particle diffusion)
2) Diffusion in the Nernst film (film diffusion)

The slower step controls the overall ion-exchange rate.

Film Diffusion
At concentrations up to 10 meq/L and flow rates up to 120 m/h used in water treatment,
diffusion rates through the resin mass are much greater than through the surrounding film.
The film thus controls the rate of exchange, and the process exhibits film-controlled
kinetics. When sodium chloride solution passes through a column of resin, originally in
+
H form, the concentration of the ion under consideration in the effluent shows variations
of the form shown in Figure 11.

Fig. 11: Film kinetics: exhaustion curves and volumes passed at maximum leakage .

If the flow rate is slow enough, equilibrium is established as the solution reaches a new layer
of the resin. The concentration in the effluent is represented by the curve OFP. At F, the
breakthrough point, the concentration reaches its maximum permitted value and flow is
usually stopped. If it were continued up to P where the concentration in the effluent water
equals that of the raw solution, the resin would be 100% exhausted.

In industrial practice, flow is stopped when the concentration of the ion under consideration in
the effluent reaches a small fraction (e.g., 1%) of the concentration in raw solution. At F, a
volume A of the solution has flown through the column. The capacity used is given by the
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 101

area OFA’Z, which differs very little from the area OFPZ representing the total capacity of
the resin.

If the exhaustion time is too short (i.e., if resin volume is small and flow rate per unit volume
is high) like in a small lab column, the operating capacity is less than the available capacity.

Particle Diffusion
As the concentration of ions in solution increases, the mass-transfer rate through the film
rises until it exceeds the diffusion rate through the resin beads. Diffusion through the resin
then becomes the controlling factor, and the system is said to exhibit particle-controlled
kinetics. This condition occurs mainly during regeneration of resins with solutions having
concentrations between 1 and 3 N.
Breakthrough curves are similar to those in Figure 11. Because of the high concentration
gradient through the resin, the whole process is much faster than the exhaustion stage.
Virtually complete equilibrium can be achieved in 15 min, but if a shorter regeneration time
is used, operating capacity can be significantly reduced.

4.6 Practical Results of Ion-Exchange Equilibrium and Kinetics


As shown in Chapter 4.6 Exchange Kinetics, under normal conditions the exchange
process is always incomplete in both the service and the regeneration stages, which means
that the total capacity of the resin can never be fully used.

4.6.1 Operating Capacity, Regeneration Efficiency, and


Regenerant Usage
Calculation of operating capacity must take into account the following:
raw water analysis
required quality of treated water (acceptable leakage)
service flow rate (l/min, m³/h)
temperature of water to be treated
type and amount of regenerant
regeneration flow rate
regenerant temperature
required duration of cycle

In practice, regeneration is not performed to reach a completely 100% exchanged fresh


ion exchanger, mostly because an enormous amount of regenerant would be required to
remove all the ions from the resin that were taken up during the previous cycle. The
variable that has the greatest effect on operating capacity is, therefore, the amount of
regenerant used.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 102

Regeneration efficiency is defined as the ratio of the operating capacity (eq) to the amount
of regenerant (eq) used and is always <1. With most strongly acidic or strongly basic
resins more than 60% of their total capacity is seldom used.

Let´s consider the following example: The total capacity of a strongly basic resin is 1.2
eq/L. A basic operating capacity of 0.46 eq/L can be obtained by using 60 g of caustic
soda per liter of resin (1.5 eq/L). In this case, the regeneration efficiency (0.46/1.5) barely
exceeds 30% (see Fig. 12). By doubling the amount of regenerant, a basic operating
capacity of 0.615 eq/L can be achieved: this is only 1/3 greater than before, and
regeneration efficiency falls to 21%. Thus, such a resin is seldom operated at more than
50% of its total capacity.

0,46

60

Fig. 12: Operating capacity of a strongly basic resin as a function of regenerant dosage
(caustic soda=NaOH); Numbers on the curves indicate percentage of regenerant efficiency.

4.6.3 Water Analysis


Calculations for designing water treatment plants that use ion-exchange resins involve some
simple concepts related to the composition of the raw water.

Units. Ion concentration is usually measured in equivalents per liter.


Water Composition. For ion-exchange calculations, the exact composition of the solution to
be treated must be known. The composition of a typical water is given in Fig.13. The total
concentration of all the anions and cations in solution is known as the total dissolved solids
+ 2+ 2+
(TDS). Ions normally encountered in water treatment are the cations Na , Ca , Mg , and the
– – – –
anions OH , CO32 , HCO3–, Cl , NO3–, SO42 .

+ 2+ 2+
Other ions may be present (K , NH4+, Mn , Fe ), but their concentration in natural water is
usually very low. For ionic equilibrium, the total cation and anion concentrations measured as
equivalents per liter must be equal.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 103

Fig. 13: Composition of a typical water

2+ 2+ 2+ 2+
Both Ca and Mg ions (and, possibly, Fe and Mn ) are classified as ions producing
hardness, and their concentration gives the total hardness (TH) measured in milliequivalents
2+ 2+ +
per liter. In water softening, Ca and Mg are exchanged for Na .

Fouling with Organic Material. The most common organic constituents in natural waters are
the high molecular mass carboxylic acids, humic and fulvic acid. These large molecules enter
the ion-exchange resins, and are trapped in the most highly cross-linked regions, their chains
becoming tangled with the resin matrix. This eventually reduces the capacity of anion
exchangers. Rinsing becomes more difficult, and problems arise with the quality of treated
water because the carboxylic character of the organic matter means that, as the pH varies,
caustic soda is initially taken up and then leached out.
Although qualitative and quantitative assessment of organic materials is difficult, their
concentration is usually expressed in milligrams of potassium permanganate required per
liter of water to oxidize them under the given conditions. The fouling factor N is the
quantity of organic matter (in milligrams per liter of KMnO4) divided by the total anion
concentration (in milliequivalents per liter).

4.6.4 Calculations in the Design of Ion-Exchange Plants for


Water Purification
General Method. Resin manufacturers provide standard charts and curves for each type of
resin, enabling calculations to be made of the volume of each resin, amount of regenerant,
operating capacity, and leakage.
2+ 2+
Weakly acidic carboxylic resins have a very high capacity for divalent ions (Ca and Mg ):
they remove the temporary hardness.
Regeneration is carried out with an amount of acid calculated to be slightly in excess of the
design capacity, with a regenerant ratio of the order of 105 – 110 %. These resins are,
therefore, very efficient.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 104

Strongly Acidic Sulfonic Resins. In hydrogen exchange, capacity depends on the type of
acid used for regeneration: hydrochloric acid is the most efficient because sulfuric acid is not
completely dissociated at the concentration normally used. In addition (this also applies to
carboxylic resins), if the resin has taken up a lot of calcium, the sulfuric acid must be highly
diluted to avoid precipitation of calcium sulfate during regeneration.

Strongly basic resins remove strong and weak acids. Because the uptake of silica is poorer
than that of other anions, it is the first to "leak." The regenerant level thus depends on the
acceptable silica leakage. In addition, because silica has a tendency to polymerize on the
resin, regenerating with hot sodium hydroxide is sometimes worthwhile. However, only type
1 materials withstand high temperature.
In practice, regenerant requirements are from 150 to ca. 500 %, so the efficiency of these
resins is only moderate.

Calculation of Resin Volume. From the calculated operating capacities, the volume of each
type of resin can be determined as a function of the operating time chosen between two
regenerations:

where
3 3
V =resin volume, m Q=flow rate, m /h
3
t =operating time, h C =operating capacity of the resin, eq/m of resin
3
S =”salinity”, total ions to be adsorbed by the resin, eq/m of water

Specific Flow Rate and Cycle Duration. The specific flow rate should not be too low
because a very even flow must always be obtained; irregularities in flow may cause
"channeling" in the resin bed, thereby impairing (hindering) complete exchange. Too high a
rate produces an excessive head loss. The specific flow rate Q/V should be between 4 and 40
BV/h.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 105

4.6.5 Description of the Ion-Exchange Cycle


Figure 14 shows a vertical section through a coflow ion-exchange column. The solution to be
treated is introduced into the column through a distributor. It passes through the resin and
emerges through a collector or collection system. In a conventional plant, the resin fills only
half the available space so that the bed can be decompacted by an upward flow of water. This
expands the resin and removes suspended matter and fragments of resin accumulated during
the previous cycle.

Fig. 14: An industrial ion-exchange column


a) Distributor ; b) Resin ; c) Collector

The ion-exchange cycle is divided into four stages:


1) exhaustion (or service),
2) decompaction (or backwash),
3) regeneration, and
4) rinsing.

Exhaustion (or Service). The solution to be treated passes through the resin bed and exhausts
it. As soon as the quantity of ions taken up reaches the operating capacity (i.e., the
breakthrough point) and leakage reaches a predetermined limiting value, the service stage is
stopped.
Backwash. After the service stage, the resin is decompacted for about 15 min by an upward
current of water. This treatment also removes any particles deposited on the surface of the
bed, together with any fragments of resin. The bed is then allowed to settle.
Regeneration. The regenerant solution is introduced, usually at a concentration of a few
percent, and slowly percolates through the bed. This takes about 15–60 min.
Rinsing. The regenerant is then displaced by water at a low flow rate until the resin bed
contains no more than traces of regenerant. This displacement or slow rinse stage is followed
by a rapid rinse stage at a higher flow rate to remove the last traces of regenerant. The
concentration of the residual regenerant is measured at the end of the operation and, as soon
as this falls to an appropriate limiting value, the next cycle begins. When water is being
demineralized, the electrolytic conductivity of the effluent is measured.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 106

Problems with Regeneration


Counterflow Regeneration [24], [25]. In the traditional method, the resin is regenerated by
flow in the same direction as that during the service stage (coflow). Figure 15 illustrates the
exhaustion of the bed in this case. Since the bed is never completely regenerated, the bottom
of the bed is only partially regenerated, which results in high leakage.
The leakage problem can be overcome by regenerating in the reverse direction (i.e., from
bottom to top). The lowest resin layers reach equilibrium with fresh, uncontaminated acid and
are therefore completely regenerated. This minimizes permanent leakage during the next
cycle.

Fig. 15: Distribution of sodium ions in ion exchange bed that was regenerated in co-
current flow.

The difference between coflow and counterflow regeneration is illustrated in Figure 16,
and their effects on treated water quality are shown in Figure 17.

Fig. 17: Effects of coflow and counterflow regeneration on the condition of resin beds during
loading and regeneration; dark areas indicate 100 % exhausted resin ; white areas, regenerated resin.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 107

Fig. 18: Effects of coflow and counterflow regeneration on ionic leakage

Mixed Beds. When beads of a cation- and an anion-exchange resin (RC and RA , respectively)
are mixed intimately, the normal equilibrium for each resin is displaced:
+ + – – + –
RC–H + Na Cl + RA+OH → RC–Na + RA+Cl + H2O
This reaction produces water. The exchange process is no longer reversible and continues
until completion. Mixed-bed resins in well-designed equipment reduce the concentration of
dissolved salts to 0.01 mg/L, giving a conductivity of 0.055 µS/cm at 25 °C, equal to that of
completely pure water. Regeneration is carried out after the two resins have been separated.
The advantages of mixed beds are twofold: (1) they produce water of excellent quality in a
single stage, and (2) final rinsing is very rapid because the regenerants neutralize each other.
However, the operating capacities of the resins are quite low, and the end of the cycle
(breakthrough) is very abrupt.

4.6.6 Choice of Resin


General Selection Criteria. To ensure efficient operation of an ion-exchange installation,
minimize running costs, and obtain a maximum resin lifetime, the resins must be selected
carefully. Before individual resins are selected, the following questions should be answered:
1) Is the solution to be treated aggressive? Does it contain oxidants?
2) What is the operating temperature?
3) Is the flow velocity high or moderate?
4) Is a high operating capacity required?
6) Is the expected leakage very low?
7) Are the resins used in a fixed, packed, or fluidized-bed unit or in a continuous
system?
8) Is an effluent total organic carbon (TOC) strictly limited?
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 108

The answers to these questions provide useful hints for resin selection:

1) In the presence of oxidants, a resin with a high degree of cross-linking should be used.
2) High operating temperature is critical for the choice of strong base resins: above 40
°C, only type 1 resins should be used.
3) For a high flow rate, physically stable resins and a relatively coarse particle size
should be used.
4) If a high capacity is required, gel-type resins are preferred over macroporous types;
5) For low leakage, resins with high cross-linking (i.e., high selectivity) associated with
high regenerant dosage will give the lowest concentration of residuals.
6) Concentrated regenerant also mean stronger osmotic shocks;
7) For low effluent TOC, special resin grades must be used that give rise to a minimum
of leachables matter.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 109

4.7 Literature:

[1] H. S. Thompson: "On the adsorbent power of soils," J. R. Agric. Soc. Engl. 11 (1850) 68.
[2] J. T. Way, J. R. Agric. Soc. Engl. 11 (1850) 313.
[3] R. Gans, Jb. Preuss. Geol. Landesamt 26 (1905) 126.
[4] O. Liebknecht, US 2 191 060, 1940.
[5] P. Smit, US 2 191 063, 1940.
[6] B. A. Adams, E. L. Holmes, J. Soc. Chem. Ind. London 54 (1935) 1.
[7] G. F. D'Alelio, US 2 366 007, 1945.
[8] C. H. Mc Burney, US 2 591 573, 1952.
[9] The Permutit Co. Ltd., GB 849 112, 860 695, 1960 (J. R. Millar).
[10] Farben Fabriken Bayer., DEI 045 102, 1957.
[11] Rohm and Haas Co., GB 932 125, 932 126, 1959.
[12] R. E. Anderson: "Fundamentals of column ion-exchange in fully-ionised system," 18th
Annual Liberty Bell Corrosion Course (USA), 1980.
[13] S. Fisher, G. Otten: "Sloughage of organic materials from field-decrosslinked sulfonic
acid cation exchange resins" Proceedings of the 42nd International Water Conference,
Pittsburgh, Pennsylvania, 1981.
[14] B. Hoffman, M. Kasahara, M. Gavaghan: "The effects of chlorine on mixed beds in
ultrapure water systems," Sixth annual semiconductor pure water conference, 1987.
[15] D. C. Auerswald: "Effects of cation resin leachables on condensate polisher system
performance," Ion Exchange for Industry, Ellis Horwood Ltd, Chichester 1988, p. 11.
[16] R. Kunin: Thermal stability of anion exchange resins. Amber-Hi-Lites 139, 1974
(published by Rohm and Haas, Philadelphia).
[17] R. E. Anderson : "Estimation of ion exchange process limits by selectivity calculations,"
AIChE Symp. Ser. 152 (1973) no. 71, 236.
[18] R. E. Anderson : "Basics of column ion exchange with strong acid and strong base
resins," Third Annual Convention Am. FESAAC, Mexico, 1982.
[19] F. G. Helfferich : Ion Exchange, McGraw Hill, New York 1964.
[20] F. G. Helfferich : Ion exchange Kinetics – Evolution of a Theory, Mass Transfer and
Kinetics of Ion Exchange, NATO ASI Series, Series E N° 71, Martins Nijhoff, The Hague
1983, p. 157.
[21] J. T. McNulty, M. Eumann, C. A. Bevan, V. C. Tan : "Anion exchange resin kinetic
testing: an indispensable diagnostic tool for condensate polisher troubleshooting,"
Proceedings of the 47th International Water Conference, Pittsburg, Pennsylvania, 1986.
[24] J. J. Wolff : "Une technique moderne d'utilisation des échangeurs d'ions: la régénération
à contre-courant," L'Eau et l'Industrie (F) 10 (1976) 67.
[25] E. W. Jackson, J. H. Smith : "Make up treatment – Countercurrent regeneration
experience in the United Kingdom," Proceedings of the 38th International Water
Conference, Pittsburgh, Pennsylvania, 1977.
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 110

4.8 Questions about this chapter:

1) Bei dem Einsatz von Ionenaustauschern wird im wesentlichen zwischen 3


Trennaufgaben unterschieden: substitution, separation and removal; geben Sie
Anwendungen der drei Verfahren auf die Behandlung einer Lösung von NiSO4;
2) Die meisten polymeren Ionentauscher werden aus welchen Polymeren hergestellt?
3) In Ionentauscherharzen sind die Polymere i.A. vernetzt, dabei wird oft der Begriff DVB
genannt; was ist DVB, wozu dient die Vernetzung?
4) Welche funktionellen Gruppen tragen Kationenaustauscher, welche Anionentauscher?
Stark saure, bzw. stark basische Ionentauscher tragen welche Gruppen?
5) Wodurch unterscheiden sich makroporöse bzw. gelförmige Ionentauscher? Wann
werden welche Typen bevorzugt?
6) Warum sind schwach saure Ionenaustauscher nicht über den gesamten pH-Bereich
einsetzbar?
7) Skizzieren Sie den Konzentrationsverlauf in einer Ionentauschersäule zu verschiedenen
Zeiten, u.a. wenn der Ionentauscher halb beladen ist und unmittelbar vor dem
Durchbruch.
8) Ionentauscherharze können auf unterschiedliche Weise beschädigt werden, nennen Sie
Beispiele. Was versteht man unter osmotischer Stabilität?
9) Die folgenden stark bzw. schwach sauren Ionentauscher werden mit Lösungen in
Kontakt gebracht. Vervollständigen Sie die Reaktionsgleichungen (nur nennenswert
ablaufende Reaktionen):
Schwach sauer: RCOOH + Na+2CO2-3 →
RCOOH + Na+2SO2-4 →
RCOOH + Ca2+Cl-2 →
RCOONa + Mg2+(NO3-)2 →
RCOONa + Zn2+SO2-4 →
Stark sauer: RSO3H + Na+2CO2-3 →
RSO3H + Na+2SO2-4 →
RSO3H + Cu2+Cl-2 →
RSO3Na + Ni2+Cl-2 →
10) Die folgenden stark bzw. schwach basischen Ionentauscher werden mit Lösungen in
Kontakt gebracht. Vervollständigen Sie die Reaktionsgleichungen (nur nennenswert
ablaufende Reaktionen):
+ – + –
Stark basisch: R OH + H Cl →
+ –
R OH + CO2 →
+ –
R Cl + NO3- →
+ –
Schwach basisch: RN(CH3)2 + H Cl →
RN(CH3)2 + Na+Cl- (neutral) →
RN(CH3)2 + CO2 + H2O →
RN(CH3)2 + SiO2 →
11) Woran erkennt man bei einem Ionentauscherversuch den Einfluß der sog.
Filmdiffusion? Unter welchen Bedingungen tritt dieser vor allem auf? Wann spielt
die innere Diffusion in den Partikeln eine Rolle?
D:\Eigene Dateien\abwasser\wasseraufbereitung2\skriptum\ionexchange.doc S. 111

12) Zeichnen Sie Durchbruchskurven für den Fall großen und kleinen Einflusses der
Filmdiffusion, ab welchem Zeitpunkt bricht man den Beladevorgang (service) des
Ionentauschers ab?
13) Wie werden Ionentauscher regeneriert? Welche Lösungen werden verwendet, was
kann man zu Gleichstrom gegenüber der Gegenstromregeneration sagen? (vgl. auch
Abb. 17)
14) Warumm macht es keinen Sinn, beim Regenerieren sehr hohe Regenerationsraten
von nahe 100% anzustreben, wie hoch ist in etwa der Regeneriermittelbedarf und
typische Regenerationsraten in der Praxis?
15) Wie kann man einwertige von zweiwertigen Ionen trennen, wie sieht der komplette
Belade- und Regenerierzyklus aus? (vgl. Abb. 9)

Das könnte Ihnen auch gefallen