You are on page 1of 103

Introduction to Electrodeionization

Welcome to Introduction to EDI. This section is dedicated to those who are just
learning about Continuous Electrodeionization also known in generic terms as
Electrodeionization (EDI). We will provide you with a basic understanding of EDI and why
it is important. It will also better prepare you for the content in the Intermediate EDI section
Introduction to EDI gives you the basics. This section helps you obtain information on
EDI fundamentals, definitions and how EDI works. It also enables you to see the benefits of
EDI, understand how to design an EDI system and watchouts during design and operation to
maintain your system.
This section is broken down into the following topics:

Advantages of EDI
Old vs. New Technologies
Development History
EDI or Ion Exchange?
Continuous Operation
Where EDI works
Comparing Old and New Technology
Benefits of EDI
Salt Movement in an Electric Field
Ion Exchange Membranes
Mechanisms of Ion Removal

Advantages of EDI
EDI utilizes chemical-free regeneration. This means a safer workplace because there
is no need to store or handle hazardous acid and caustic. There are fewer regulation concerns
due to the absence of these corrosive chemicals and there are no waste neutralization or
disposal issues.
EDI is a continuous process. The ion exchange resins are being continuously being
regenerated by the DC electric field. There is no breakthrough" of ions as happens in
conventional ion exchange operations, therefore the quality of the water remains at a constant
high level of purity. The electric field also provides a bacteriostatic environment inside of the
EDI cell, inhibiting the growth of bacteria and other organisms.
EDI has significantly lower operating costs than conventional ion exchange
processes. Only a relatively small amount of electric power is needed to provide high purity
water. The lack of acid and caustic regeneration means less operator attention and lower
labor costs. Capital costs can also expected to be lower, especially because no chemical
storage, pumping and neutralization equipment is required.
EDI has a significantly smaller footprint than conventional ion exchange processes.
This means less plant space will be required to provide the same quantity of water.

When to Consider Using EDI


EDI may be considered to be a competitive alternate process to:
Regenerable Mixed Bed Deionization
No acid or caustic bulk storage, pumping, waste neutralization or disposal issues.
Lower operating cost due lower manpower requirements as well as the lack of chemical
regeneration Smaller footprint.
Service Mixed Bed Deionization (off-site regenerated rental vessels)
No ionic breakthrough resulting in a constant high quality of water. No rental vessels,
associated freight or monthly demurrage charges.
Second pass of RO
Eliminates the need for a second bank of RO membranes and associated plumbing,
pumping and control equipment.
Typical applications for EDI
Providing USP (United States Pharmacopeia) grade water
Purified Water
Feed to WFI (Water For Injection) stills
Steam Generation / Boiler Feed
Microelectronics / Semiconductor makeup and rinsewaters
For high quality makeup demineralized water
General Industry
Surface finishing
Chemical manufacturing
Hospital & University Central Systems
Advantages of EDI - Pharmaceutical Applications
EDI runs in a bacteriostatic service mode. The electric field minimizes bacteria
growth in the resin beds. Some EDI devices hot water sanitizable, resulting in more effective
sanitization, faster rinseup and more easily validated systems. RO/EDI easily meets Stage 1
conductivity specification, while conventional 2-Pass RO (RO/RO) does not, due to
CO2 and/or NH3.
Advantages of EDI - Steam Generation Applications
EDIs continuous electrochemical regeneration provides a constant water quality with
no breakthrough" of bands of ions. EDI is a low maintenance process requiring little
operator attention. EDI is also ideal for remote locations where you might be dependent on
delivery of chemicals or DI tanks. EDI can provide low product water conductivity, of much
better quality then would be possible with 2-pass RO alone

Advantages of EDI - Microelectronics Applications


EDI provides high quality water, low in particles, partiall due to the fact that there is
no resin attrition from backwashing or osmotic shock, as would be the case with conventional
ion exchange processes. EDI also provide water with a lower TOC (Total Organic Carbon)
content because the electric field will remove charged organic molecules. Boron removal is
better on average than mixed bed deionization, unless the mixed bed columns are regenerated
very frequently.
Perceived Limitations of EDI
Leaks
Leaks have been completely eliminated in some module designs.
Sensitivity to chlorine
EDI is just as sensitive to the chlorine as thin-film reverse osmosis (RO) processes. This
is easily addressed with proper pretreatment system design.
Sensitivity to hardness
Most EDI devices have a 1 ppm hardness limit which is easily addressed with proper
pretreatment system design.
Not suitable for high CO2 loads
No ion exchange process is cost-effective for removal of large amounts of CO2.
Removal Mechanisms
While both ion exchange and EDI use ion exchange resins, the removal mechanisms
are quite different. Conventional ion exchange utilizes chemically regenerated ion exchange
resins which function in a capture (exhaustion cycle) and discharge (regeneration cycle)
mode. This results in a breakthrough of ions at the end of the service cycle and a rinseout of
regenerant at the beginning of the next service cycle. Capacity & selectivity are most
important resin properties in this mode of operation.
EDI (continuous electrodeionization) utilizes a reaction/transport mechanism to
remove ions (through resin under influence of DC field). This requires continuous path of
like-charge resin beads. The transport is largely across the surface of the resin beads.
Transport through resin bead (particle diffusion) can be limiting.

Water Treatment - Old vs. New Technologies

Conventional water treatment systems rely on chemically-regenerated ion exchange


resins to remove dissolved solids. Regeneration chemicals are costly, hazardous and, even
though they are neutralized prior to releasing to streams and rivers, add a significant amount
of dissolved solids to the waterways. These systems typically use cation exchange vessels
followed by anion exchange to handle the bulk of the demineralization. Mixed bed
exchangers are used in many cases to polish the treated water to a resistivity of 18
megohms. The waste regenerants from these systems are usually combined, neutralized and
released to the environment. The ion exchange systems are usually supplied in duplicate, to
allow one system to provide water while the other one is being regenerated. Regenerations of
ion exchangers typically takes several hours, require bulk storage and pumping facilities for
regenerant chemicals, and usually require a waste neutralization tank.

State-of-the-art water treatment systems utilize reverse osmosis (RO) membranes to


do the bulk of the demineralization. RO systems do not require chemical regeneration and
also remove many types of total organic carbon (TOC) which will pass through ion exchange
resins. (In some cases, the ion exchange resins actually contribute to the TOC content in the
water). The polishing of the RO product water is carried out by continuous
electrodeionization (EDI) which is capable of producing water in excess of 18 megohm
resistivity. Duplexing of large EDI systems is not required, because a single EDI module may
be taken off line for maintenance or repair while the remaining modules operate at a slight
increase in flow rate to maintain the required flow through the system. EDI modules can be
sized to operate from a fraction of a gpm up to about 50 gpm. The EDI process is a

continuous process, utilizes no chemicals for regeneration, does not pollute the environment
and requires a fraction of the operator attention necessary for conventional ion exchange
systems. EDI systems typically run between 90 and 95% water recovery. The reject stream is
usually of better quality than the feed to the RO system, enabling the reject stream to be
completely re-used by pumping it back to the pretreatment section of the RO system. RO/EDI
systems may achieve overall water recoveries of greater than 90% by recycling a significant
portion of the RO reject stream and all of the EDI reject stream back to the front end of the
treatment system.
Development History
EDI theory and practice have been advanced by a large number of researchers
throughout the world. It is believed that EDI was first described in a publication by scientists
at Argonne Labs in January 1955 as a method for removal of trace radioactive materials from
water (Walters, et. al.). One of the earliest known patents describing a EDI device and process
was awarded in 1957 (Kollsman). It is thought that the first pilot device incorporating mixed
resins was developed by Permutit Company in the United Kingdom in the late 1950's for the
Harwell Atomic Energy Authority, as described in a paper (Gittens and Watts) and in more
than one patent (Kressman; Tye). One of the first detailed theoretical discussions of EDI was
written in December 1959 (Glueckauf). In April 1971, a Czechoslovakian researcher reported
results of his experimental and theoretical work that advanced the theory of ionic transport
within a EDI device (Mateka). Layered bed devices were described in the patent literature in
the early 1980's (Kunz).
EDI devices and systems were first fully commercialized in early 1987 by a division
of Millipore that is now part of United States Filter Corporation (Ganzi et. al., 1987). Since
then, the theory and practice of EDI has advanced worldwide, and commercial EDI devices
are now manufactured by a number of companies (Towe et. al.; Parsi et. al.; Rychen et. al.,
Stewart and Darbouret). There are now several thousand EDI systems in commercial
operation for the production of high purity water at capacities ranging from less than 0.1 to
more than 250 m3/h. This includes EDI systems that have been in continuous operation for
nearly ten years, producing makeup water for high pressure boilers.
Continuous Electrodeionization or Ion Exchange?
Companies are constantly working to reduce operating costs, improve efficiency and
eliminate the use of hazardous chemicals in the workplace. Such goals have caused an
increase in the use of continuous electrodeionization technology to produce high-purity water.
Continuous electrodeionization (EDI) uses a combination of ion exchange resins and
membranes, and direct current to continuously deionize the water without regeneration
chemicals.
The principle behind ion exchange is that polymer resin beads are chemically
structured to provide either positively or negatively charged fixed functional groups that

attract and remove certain contaminant ions from the water. Conventional ion exchange
technology can remove dissolved inorganics such as minerals and salts and some dissolved
organics. It does not remove particles, colloids, bacteria or pyrogens.
Cationic resins remove positively charged ions such as calcium, magnesium and
sodium, replacing them with hydrogen (H+) ions. Anionic resins remove such negatively
charged ions as chloride, nitrate and silica and replace them with hydroxide (OH -) ions. The
hydrogen and hydroxide ions then unite to form water molecules.
When the water passes through a tank containing a mixture of both cation and anion
exchange resins, the process is called mixed-bed ion exchange. Mixed-bed systems can
produce very high-quality water with resistivities up to 18.2 megohms-cm.
Over time, however, the resin beads become saturated with contaminant ions and
become less effective at treating the water. Also, the high-purity water flowing past these
saturated resins may actually extract trace amounts of contaminant ions by "chromatographic
effects," causing a decline in water quality. The exhausted resins must be chemically
regenerated off-line before reuse. During regeneration, the cation resin beads are restored to
their hydrogen form by treating them with hydrochloric acid (HCl) or sulfuric acid (H 2SO4).
The anion resin beads are restored to their hydroxide form by treating them with caustic
(NaOH).
A Continuous Operation
Like conventional ion exchange, continuous electrodeionization removes dissolved,
ionizable materials such as salts, acids and bases. It can also remove weakly ionized materials
such as dissolved silica, carbon dioxide and some organics. Contaminants too large to pass
through the ion exchange membrane -- such as large particles and large organic molecules -are not removed.
Continuous electrodeionization modules consist of mixed-bed resins sandwiched
between alternating anion and cation membranes. These membranes are actually ion
exchange resins manufactured in sheet form. Resin compartments in this "sandwich"
construction alternate between diluting and concentrating compartments. Compartment sets
are called cell pairs and form the basic element in a module.
In the module, direct current is applied to the anode (positive electrode) on one end of
the module, and to the cathode (negative electrode) on the other end. This electric potential
drives the ions captured by the ion exchange resins through the membrane.
Because the resins in the module are continuously regenerated by the electric current,
they do not become exhausted. This continuous electro-regeneration enables continuous
electrodeionization systems to produce multi-megohm water without the need for chemical
regeneration or downtime.
Commercially available continuous electrodeionization modules are normally plateand-frame devices with varying numbers of cell pairs to accommodate flow rates from 0.5

gallons per minute (gpm) to 1,000 gpm. Customized systems can produce even higher flow
rates.
Where it Works
Continuous electrodeionization can be chosen for new projects that require ultra-high
purity water or stringent wastewater discharge requirements. It can also be a cost-effective
way to upgrade existing ion exchange systems.
Applications include creating process water for the biotechnology and food and
beverage industries, and providing high-quality rinse water for electronics, surface finishing
and optical-glass processes. For electronics applications, continuous electrodeionization can
reduce total organic carbon (TOC) levels in the water.
Because it is able to meet United States Pharmacopoeia (USP) purified water
specifications -- including those for bacteria and pyrogens -- continuous electrodeionization
systems are frequently used in the pharmaceutical industry, hospitals, university research
facilities and dialysis centers.
Electrodeionization (EDI) is a process that removes ionizable species from liquids
using electrically active media and an electrical potential to effect ion transport. The
electrically active media in EDI devices may function to alternately collect and discharge
ionizable species, or to facilitate the transport of ions continuously by ionic or electronic
substitution mechanisms. EDI devices may comprise media of permanent or temporary
charge, and may be operated batchwise, intermittently, or continuously.
The continuous electrodeionization (EDI) process, a subset of EDI, is distinguished
from the EDI collection/discharge processes such as electrochemical ion exchange (EIX) or
capacitive deionization (CapDI), in that EDI performance is determined by the ionic transport
properties of the active media, not the ionic capacity of the media. EDI devices typically
contain semi-permeable ion-exchange membranes and permanently charged media such as
ion-exchange resin. The EDI process is essentially a hybrid of two well-known separation
processes - ion exchange deionization and electrodialysis, and is sometimes referred to as
filled-cell electrodialysis.
This section will focus on EDI devices, and more specifically on the use of such
devices in conjunction with reverse osmosis to produce water of sufficient quality to feed to
high pressure boilers.
Benefits of CEDI
Conventional ion exchange is a viable treatment option for applications where high-flow and
high-conductivity requirements are not critical. In other situations, however, cost
comparisons between ion exchange and continuous deionization may reveal potential
economic advantages.

The biggest difference is that continuous electrodeionization eliminates all chemical


regeneration and waste neutralization steps. While capital equipment costs may be higher
with continuous deionization systems, operating expenses are lower because there are no
regeneration chemicals, and labor or maintenance costs are less.
Electrical requirements are nominal. A typical system uses one kilowatt-hour (KwH)
of electricity to deionize 1,000 gallons, based on a feed conductivity of 50 micromhos/cm and
a product water resistivity of 10 megohm-cm.
Continuous electrodeionization systems do not require duplexing (two separate
treatment units) which can increase the cost, complexity and size of the system.
Both conventional ion exchange and continuous electrodeionization may require
pretreated feedwater to prevent scale formation and plugging by colloids and particles.
Pretreatment is also required to reduce high levels of free chlorine and organic foulants. The
type of pretreatment required is determined by product water quality requirements. For most
high-purity water needs, a reverse osmosis (RO) pretreatment step is sufficient.
With RO pretreatment, continuous electrodeionization systems can achieve better than
99.5 percent salt removal, reduce the levels of individual ionic species to parts-per-billion or
even parts-per-trillion levels, and produce high-purity water with resistivities of 10 to 18
megohm-cm (or 0.1 to 0.055 micromho/cm conductivity).
Continuously regenerating the ion exchange resins also removes the possibility of
exhausted or improperly regenerated resins contaminating the product water.
Continuous electrodeionization systems typically convert 80 to 95 percent of the
feedwater into product water. The waste stream can be discharged without treatment or
recycled back to the RO pretreatment unit.
It can also reduce facilities costs because waste neutralization equipment and
hazardous fumes ventilation equipment are not required. The elimination of regeneration
chemicals can help improve workplace health and safety, as well as prevent corrosion from
hydrochloric acid fumes. The costs of meeting EPA and OSHA requirements and "right to
know" laws are also reduced.
There are significant tangible cost benefits associated with the elimination of
regeneration. The costs of regeneration labor and chemicals are replaced with a small amount
of electrical consumption. A typical EDI system will use approximately 1 kW-hr of electricity
to deionize 1000 gallons from a feed conductivity of 50 microsiemen /cm to 0.1 S/cm

product conductivity. Since the EDI concentrate (or reject) stream contains only the feed
water contaminants at 5-20 times higher concentration, it can usually be discharged without
treatment, or used for another process. Thus facility costs can also be reduced since waste
neutralization equipment and ventilation for hazardous fumes are not necessary.
There are also less tangible cost reductions, which are harder to quantify, but usually
favor the use of EDI systems. By eliminating hazardous chemicals wherever possible,
workplace health and safety conditions can be improved. With today's increasing regulatory
influence on the workplace, the storage, use, neutralization, and disposal of hazardous
chemicals result in hidden costs associated with monitoring and paperwork to conform to
safety and environmental requirements (such as the US EPA and OSHA laws). In addition,
the fumes, particularly from acid, often cause corrosive structural damage to facilities and
equipment.
For the most part the elimination of regenerant chemicals is considered advantageous,
but the chemicals do offer at least one benefit. In conventional demineralizers, acid and
caustic is typically applied to the ion exchange resins at concentrations of 2-8% by weight. At
these concentrations the chemicals not only regenerate the resins but clean them as well. The
electrochemical regeneration that occurs in a EDI device does not provide the same level of
resin cleaning. Therefore proper pretreatment is even more important with a EDI device, in
order to prevent fouling or scaling. This is one of the reasons that RO pretreatment is
normally required upstream of a EDI system. In general, the feed water requirements for EDI
systems are more stringent than for chemically regenerated demineralizers.
In summary:

EDI does not require chemicals (as does DI resin regeneration)


EDI does not require shutdowns
EDI modules are the smallest and lightest per unit flow on the market, EDI systems
are compact as well.
Consistent, continuous ultrapure water
Requires little energy
Economic use of capital, no operating expenses, just the cost of power.
Reduced facility requirements

Minimal operator supervision required

How does it Work

First, lets start with a container of water with salts (cations and anions) in solution.

A typical EDI device contains alternating semipermeable anion and cation ion-exchange
membranes. The spaces between the membranes are configured to create liquid flow
compartments with inlets and outlets. A transverse DC electrical field is applied by an
external power source using electrodes at the ends of the membranes and compartments.
When the compartments are subjected to an electric field, ions in the liquid are attracted to
their respective counter-electrodes. The result is that the compartments bounded by the anion
membrane facing the anode and the cation membrane facing the cathode become depleted of
ions and are thus called diluting compartments. The compartments bounded by the anion
membrane facing the cathode and cation membrane facing the anode will then ?trap? ions
that have transferred in from the diluting compartments. Since the concentration of ions in
these compartments increases relative to the feed, they are called concentrating
compartments, and the water flowing through them is referred to as the concentrate stream
(or sometimes, the reject stream).
Ion Exchange Membranes
Now let's add some ion exchange membranes to direct the ions into different flow channels as
shown in the animation below. The red membranes are cation-selective membranes and the
blue membranes are anion-selective membranes. The negatively-charged anions (e.g., Cl-) are
attracted to the anode (+) and repelled by the cathode (-). The anions pass through the anion-

selective membrane and into the adjacent concentrate stream where they are blocked by the
cation-selective membrane on the far side of the chamber, and are thus trapped and carried
away by the carrier water in the concentrate stream. The positively-charged cations (e.g., Na+)
in the purifying stream are attracted to the cathode (-) and repelled by the anode (+). The
cations pass through the cation-selective membrane and into the adjacent concentrate stream
where they are blocked by the anion-selective membrane and are carried away.
In the concentrate stream, electrical neutrality is maintained. Transported ions from the two
directions neutralize one another's charge. The current draw from the power supply is
proportional to the number of ions moved. Both the "split" water (H+ and OH-) and the
intended ions are transported and add to the current demand.

Theory of Operation
In an EDI device, the space within the ion depleting compartments (and in some cases in the
ion concentrating compartments) is filled with electrically active media such as ion exchange
resin. The ion-exchange resin enhances the transport of ions and can also participate as a
substrate for electrochemical reactions, such as splitting of water into hydrogen (H+) and
hydroxyl (OH-) ions. Different media configurations are possible, such as intimately mixed
anion and cation exchange resins (mixed bed or MB) or separate sections of ion-exchange
resin, each section substantially comprised of resins of the same polarity: e.g., either anion or
cation resin (layered bed or LB and single bed or SB).
The electrodeionization process uses a combination of ion-selective membranes and ionexchange resins sandwiched between two electrodes (anode (+) and cathode (-)) under a DC
voltage potential to remove ions from RO-pretreated water.
Ion-selective membranes operate using the same principle and materials as ion- exchange
resins, and they are used to transport specific ions away from their counter ions. Anionselective membranes are permeable to anions but not to cations; cation-selective membranes
are permeable to cations but not to anions. The membranes are not water-permeable.
By spacing alternating layers of anion-selective and cation-selective membranes within a
plate-and-frame module, a "stack" of parallel purifying and concentrating compartments are
created. The ion-selective membranes are fixed to an inert polymer frame, which is filled with
mixed ion-exchange resins to form the purifying chambers. The screens between the purifying
chambers form the concentrating chambers.
This basic repeating element of the EDI, called a "cell-pair", is illustrated in Figure 1. The
"stack" of cell-pairs is positioned between the two electrodes, which supply the DC potential
to the module. Under the influence of the applied DC voltage potential, ions are transported

across the membranes from the purifying chambers into the concentrating chambers. Thus, as
water moves through the purifying chambers, it becomes free of ions. This stream is the pure
water product stream.
Most commercial EDI devices comprise alternating cation- and anion-permeable membranes
with spaces in between configured to create liquid flow compartments with inlets and outlets.
The compartments bound by a positively charged anion exchange membrane (AEM) facing
the positively charged anode (+) and a negatively charged cation exchange membrane (CEM)
facing the negatively charged cathode (-) are diluting or purifying compartments. The
compartments bound by an anion membrane facing the cathode and a cation membrane facing
the anode are concentrating compartments. To facilitate ion transfer in low ionic strength
solutions, the dilute compartments, and sometimes the concentrate compartments, are filled
with ion exchange resins. A transverse DC electrical field is applied by an external power
source using electrodes at the bounds of the compartments such that ions in the liquid are
attracted to their respective counter electrodes. The result is that the diluting compartments are
depleted of ions and the concentrating compartments are concentrated with ions. Figure 1 is a
representation of the process showing two diluting compartments and one concentrating
compartment.

FIGURE 1.
NOTES:
CEM = Cation exchange membrane.
AEM = Anion exchange membrane.
The product stream may also be referred to as the dilute stream.

The reject stream may also be referred to as the concentrate stream.


To facilitate ion transfer in low ionic strength solutions, the dilute compartments, and
sometimes the concentrate compartments, are filled with ion exchange resins. A transverse DC
electrical field is applied by an external power source using electrodes at the bounds of the
compartments such that ions in the liquid are attracted to their respective counter electrodes.
The result is that the diluting compartments are depleted of ions and the concentrating
compartments are concentrated with ions. The figure above is a representation of the process
showing two diluting compartments and one concentrating compartment.
Mechanism of
Ion Removal
There are two distinct operating regimes for EDI devices: enhanced transfer and
electroregeneration (Ganzi, 1988). In the enhanced transfer regime, the resins within the
device remain in the salt forms. In low conductivity solutions the ion exchange resin is orders
of magnitude more conductive than the solution, and acts as a medium for transport of ions
across the compartments to the surface of the ion exchange membranes. This mode of ion
removal is only applicable in devices that allow simultaneous removal of both anions and
cations, in order to maintain electroneutrality.
The second operating regime for EDI devices is known as the electroregeneration regime.
This regime is characterized by the continuous regeneration of resins by hydrogen and
hydroxide ions from the electrically-induced dissociation of water. This dissociation
preferentially occurs at bipolar interfaces in the ion-depleting compartment where localized
conditions of low solute concentrations are most likely to occur (Simons). The two primary
types of bipolar interfaces in EDI devices are resin/resin and resin/membrane. The optimum
location for water splitting depends on the configuration of the resin filler. For mixed-bed
devices water splitting at both types of interface can result in effective resin regeneration,
while in layered bed devices water is dissociated primarily at the resin/membrane interface
(Ganzi et. al., 1997).
Regenerating the resins to their H+ and OH- forms allows EDI devices to remove weakly
ionized compounds such as carbonic and silicic acids, and to remove weakly ionized organic
compounds. This mode of ion removal occurs in all EDI devices that produce ultrapure water.
Figure 13-1 is a representation of the process showing two diluting compartments, which
illustrates the transport of ions and electrochemical regeneration of ion exchange resins in one
type of EDI cell.
To construct a large scale EDI device, many of these cells are assembled together and fed in
parallel as shown in Figure 2.

Figure 2
The animation below illustrates the removal of ions and the splitting of water molecules:
Mechanism of Ion Removal
There are two distinct operating regimes for EDI devices: enhanced transfer and
electroregeneration (Ganzi, 1988). In the enhanced transfer regime, the resins within the
device remain in the salt forms. In low conductivity solutions the ion exchange resin is orders
of magnitude more conductive than the solution, and acts as a medium for transport of ions
across the compartments to the surface of the ion exchange membranes. This mode of ion
removal is only applicable in devices that allow simultaneous removal of both anions and
cations, in order to maintain electroneutrality.
The second operating regime for EDI devices is known as the electroregeneration regime.
This regime is characterized by the continuous regeneration of resins by hydrogen and
hydroxide ions from the electrically-induced dissociation of water. This dissociation
preferentially occurs at bipolar interfaces in the ion-depleting compartment where localized
conditions of low solute concentrations are most likely to occur (Simons). The two primary
types of bipolar interfaces in EDI devices are resin/resin and resin/membrane. The optimum
location for water splitting depends on the configuration of the resin filler. For mixed-bed
devices water splitting at both types of interface can result in effective resin regeneration,
while in layered bed devices water is dissociated primarily at the resin/membrane interface
(Ganzi et. al., 1997).
Regenerating the resins to their H+ and OH- forms allows EDI devices to remove weakly
ionized compounds such as carbonic and silicic acids, and to remove weakly ionized organic
compounds. This mode of ion removal occurs in all EDI devices that produce ultrapure water.
Figure 13-1 is a representation of the process showing two diluting compartments, which
illustrates the transport of ions and electrochemical regeneration of ion exchange resins in
one type of EDI cell.

To construct a large scale EDI device, many of these cells are assembled together and fed in
parallel as shown in Figure 2.

Figure 2
The animation below illustrates the removal of ions and the splitting of water molecules:

There are various types of EDI devices. Due to the nature of the process however,
manufacturers to date have taken only two design approaches. These are plate and frame and

spiral wound. The plate and frame design was the first to emerge and is similar to a plate and
frame filter press or heat exchanger in construction. Alternating diluting and concentrating
cells are stacked between the electrodes and sandwiched together with some type of closing
mechanism. Increasing the number of cell pairs (one dilute and one concentrate cell)
increases the capacity of the unit. The main advantage of this type of construction is the ease
of assembly.
In contrast, spiral wound EDI is a bit more complicated and hence more difficult to assemble.
We will not attempt to explain the details of spiral EDI construction, however the basic
principles of deionization are similar to plate and frame configuration.
Currently, the majority of EDI devices on the market use plate and frame construction. Plate
and frame devices can be broken down into two major subsets. These are thin cell and thick
cell, designated as such based on the thickness of the diluting compartments. For the purposes
of this discussion, thin cell will encompass devices with a diluting cell thickness of 2-3 mm
and thick cell will encompass 8-10 mm.

There are various types of EDI devices. Due to the nature of the process however,
manufacturers to date have taken only two design approaches. These are plate and frame and
spiral wound. The plate and frame design was the first to emerge and is similar to a plate and
frame filter press or heat exchanger in construction. Alternating diluting and concentrating
cells are stacked between the electrodes and sandwiched together with some type of closing
mechanism. Increasing the number of cell pairs (one dilute and one concentrate cell)
increases the capacity of the unit. The main advantage of this type of construction is the ease
of assembly.
In contrast, spiral wound EDI is a bit more complicated and hence more difficult to assemble.
We will not attempt to explain the details of spiral EDI construction, however the basic
principles of deionization are similar to plate and frame configuration.
Currently, the majority of EDI devices on the market use plate and frame construction. Plate
and frame devices can be broken down into two major subsets. These are thin cell and thick
cell, designated as such based on the thickness of the diluting compartments. For the purposes
of this discussion, thin cell will encompass devices with a diluting cell thickness of 2-3 mm
and thick cell will encompass 8-10 mm.
Inter
media
te
Index
The Intermediate section is comprised of the following EDI topics:
EDI construction

EDI Module Construction


Electrode Reactions and Material Selection

Electroactive Media used in EDI Devices

Ion Exchange Resin Selection


Ion Exchange Membrane Selection
Mixed bed Resin Filler (EDI-MB) - Intermembrane Spacing
Mixed bed Resin Filler (EDI-MB) - Resin Packing
Layered Bed Resin Filler (EDI-LB)
Separate Bed Resin Filler (EDI-SB)

Thin Cell and Thick Cell EDI

Thin Cell EDI


Thick Cell EDI

DC Current and Voltage

Faraday's Law
Current Efficiency and E-factor
Ohm's Law and Module Resistance

Removal of Ionic Species

Series removal of species


Carbon Dioxide Removal
Silica Removal
Water splitting and Module Resistance

Gas Generation in the Electrode Compartment


CEDI
Const
ructio
n
This description of module construction will first discuss the overall device and then the
individual cell. Commercially available devices are produced in two main configurations:
plate-and-frame or spiral wound. The plate-type devices are similar in concept to a plate-andframe heat exchanger, with multiple fluid compartments sandwiched between a set of
endplates (and electrodes) that are held in compression by bolts or threaded rods. The
compartments alternate between diluting and concentrating, and are hydraulically in parallel
but electrically in series. An exploded view of a typical plate-and-frame EDI device is shown
in Figure 1.

Figure 1
Plate-and-frame EDI device
The spiral EDI devices are analogous to a spiral wound membrane element, but with the
membrane, resins, and spacers wound spirally around a center electrode rather than a
permeate tube. Spiral wound devices must be installed inside a pressure vessel, while plateand-frame devices incorporate some means of sealing on the individual fluid compartments,
essentially making each a pressure vessel. Spiral wound devices are somewhat more difficult
to assemble than plate-and-frame units. A typical spiral device is shown in Figure 2.

Figure 2
Spiral-wound EDI device
The cells themselves can be classified at either thin cell or thick cell. Thin-cell devices are
those with a spacing of approximately 1.5-3.5 mm between the ion exchange membranes in

the diluting compartments, while thick cell devices typically use intermembrane spacing of 810 mm. Both plate-and-frame and spiral wound configurations are suitable for either thin or
thick cell construction. As will be shown below, thin-cell devices allow the use of intimately
mixed anion and cation exchange resins in the product compartments, while thick cells work
best with separate regions that contain primarily resins of the same polarity.
Flow Spacers
All commercial EDI devices use ion exchange resin in the diluting compartments, and
therefore require a component to contain the resin. This 'resin spacer' consists of an inlet port,
an inlet distributor, the resin compartment, an outlet distributor, and an outlet port. It is
necessary to provide a means of sealing the ion exchange membrane against the spacer to
form the sides of the resin compartment. Some designs will also include additional ports to
allow slurrying the resin in and out of the cell. A typical dilute spacer for a plate-and-frame
EDI device is shown in Figure 3.

Figure 3
Dilute spacer from thick cell, layered bed EDI module
All EDI devices will also require flow compartments for the concentrate and electrode
streams as well. The two types most commonly used are either a flow-through screen or a
resin compartment. The flow-through screen is similar to a sheet-flow electrodialysis spacer.
It generally consists of a woven plastic mesh screen (also like an RO feed spacer) that
incorporates some sort of sealing mechanism, such as a rubber gasket interpenetrated in the
perimeter of the screen.
The use of screen-type concentrate spacers is quite common in EDI devices, as they are fairly

inexpensive and relatively easy to fabricate. Their major disadvantage is that they are not
conductive. Since the makeup water feeding the concentrate compartments is normally RO
permeate (to avoid scaling and fouling), the concentrate stream is not very conductive, in
spite of the ions transferring into the concentrate from the diluting compartments. For
example, a EDI system fed RO permeate with a conductivity of 5 S/cm and operating at
90% water recovery would typically have a concentrate outlet conductivity of about 50
S/cm. This is low enough to limit the amount of current that can be passed through the
module (see discussion of resistance, below). Several manufacturers recommend injection of
salt into the concentrate to raise the conductivity to 300 S/cm or more.
An alternative to the use of screen-type spacers for the concentrate and electrode
compartments is to use a resin-filled compartment similar to the ones used for the diluting
compartments. By employing a conductive filler, the use of salt injection can be avoided. It
has been found that injection of a salt solution into a resin-filled concentrate compartment has
little effect on module resistance.
Use of the same size compartments for both diluting and concentrating compartments would
typically require a concentrate recirculation pump to maintain adequate velocity in the
concentrate while limiting the water sent to drain, to provide high water recovery. An
alternative approach is to make the concentrate resin compartments thinner than the dilute
resin compartments, to avoid the use of a concentrate recirculation pump.
Electrode Reactions and Material Selection
At the cathode, or negatively charged electrode, electrons are transferred from the external
circuit to ions in the solution by the following reaction:
H2O+ e--> H2 + OHEq. 6
Therefore an electrode that is stable in the presence of base and hydrogen is required. The
most common cathode material for EDI devices is stainless steel.
At the anode, or positively charged electrode, electrons are transferred from ions in solution
to the external circuit by one or more of the following reactions:
H2O -> O2 + H+ + eEq. 7
Cl- -> Cl2 + eEq. 8
Commonly used anode materials include iridium-coated titanium and platinum-coated
titanium.
Gases are evolved by the reactions at both the cathode and anode. These must be removed to
prevent masking the surface of the electrode, which would result in a voltage drop and reduce
the voltage applied to the cells. Removal of the gases is accomplished by maintaining a flow

of water across the surface of the electrodes during operation. This requires the use of a flow
compartment adjacent to the electrode. Such compartments could be either gasketed screentype spacers or resin-filled compartments.
Since copper wire is commonly used to conduct electric current to the electrodes, the junction
of the copper wire and the non-copper electrode may be subject to corrosion, particularly if it
is damp. If possible, it is best to have a projection of the electrode material that passes
through the end plate of the module to an external connection that can be kept clean and dry
(Schweitzer).

CEDI
Construction
This description of module construction will first discuss the overall device and then the individual
cell. Commercially available devices are produced in two main configurations: plate-and-frame or
spiral wound. The plate-type devices are similar in concept to a plate-and-frame heat exchanger,

with multiple fluid compartments sandwiched between a set of endplates (and electrodes) that are
held in compression by bolts or threaded rods. The compartments alternate between diluting and
concentrating, and are hydraulically in parallel but electrically in series. An exploded view of a
typical plate-and-frame EDI device is shown in Figure 1.

Figure 1
Plate-and-frame EDI device
The spiral EDI devices are analogous to a spiral wound membrane element, but with the
membrane, resins, and spacers wound spirally around a center electrode rather than a permeate
tube. Spiral wound devices must be installed inside a pressure vessel, while plate-and-frame
devices incorporate some means of sealing on the individual fluid compartments, essentially
making each a pressure vessel. Spiral wound devices are somewhat more difficult to assemble than
plate-and-frame units. A typical spiral device is shown in Figure 2.

Figure 2
Spiral-wound EDI device
The cells themselves can be classified at either thin cell or thick cell. Thin-cell devices are those
with a spacing of approximately 1.5-3.5 mm between the ion exchange membranes in the diluting
compartments, while thick cell devices typically use intermembrane spacing of 8-10 mm. Both
plate-and-frame and spiral wound configurations are suitable for either thin or thick cell
construction. As will be shown below, thin-cell devices allow the use of intimately mixed anion
and cation exchange resins in the product compartments, while thick cells work best with separate
regions that contain primarily resins of the same polarity.
Flow Spacers
All commercial EDI devices use ion exchange resin in the diluting compartments, and therefore
require a component to contain the resin. This 'resin spacer' consists of an inlet port, an inlet
distributor, the resin compartment, an outlet distributor, and an outlet port. It is necessary to
provide a means of sealing the ion exchange membrane against the spacer to form the sides of the
resin compartment. Some designs will also include additional ports to allow slurrying the resin in
and out of the cell. A typical dilute spacer for a plate-and-frame EDI device is shown in Figure 3.

Figure 3
Dilute spacer from thick cell, layered bed EDI module
All EDI devices will also require flow compartments for the concentrate and electrode streams as
well. The two types most commonly used are either a flow-through screen or a resin compartment.
The flow-through screen is similar to a sheet-flow electrodialysis spacer. It generally consists of a
woven plastic mesh screen (also like an RO feed spacer) that incorporates some sort of sealing

mechanism, such as a rubber gasket interpenetrated in the perimeter of the screen.


The use of screen-type concentrate spacers is quite common in EDI devices, as they are fairly
inexpensive and relatively easy to fabricate. Their major disadvantage is that they are not
conductive. Since the makeup water feeding the concentrate compartments is normally RO
permeate (to avoid scaling and fouling), the concentrate stream is not very conductive, in spite of
the ions transferring into the concentrate from the diluting compartments. For example, a EDI
system fed RO permeate with a conductivity of 5 S/cm and operating at 90% water recovery
would typically have a concentrate outlet conductivity of about 50 S/cm. This is low enough to
limit the amount of current that can be passed through the module (see discussion of resistance,
below). Several manufacturers recommend injection of salt into the concentrate to raise the
conductivity to 300 S/cm or more.
An alternative to the use of screen-type spacers for the concentrate and electrode compartments is
to use a resin-filled compartment similar to the ones used for the diluting compartments. By
employing a conductive filler, the use of salt injection can be avoided. It has been found that
injection of a salt solution into a resin-filled concentrate compartment has little effect on module
resistance.
Use of the same size compartments for both diluting and concentrating compartments would
typically require a concentrate recirculation pump to maintain adequate velocity in the concentrate
while limiting the water sent to drain, to provide high water recovery. An alternative approach is to
make the concentrate resin compartments thinner than the dilute resin compartments, to avoid the
use of a concentrate recirculation pump.
Electrode Reactions and Material Selection
At the cathode, or negatively charged electrode, electrons are transferred from the external circuit
to ions in the solution by the following reaction:
H2O+ e--> H2 + OHEq. 6
Therefore an electrode that is stable in the presence of base and hydrogen is required. The most
common cathode material for EDI devices is stainless steel.
At the anode, or positively charged electrode, electrons are transferred from ions in solution to the
external circuit by one or more of the following reactions:
H2O -> O2 + H+ + eEq. 7
Cl- -> Cl2 + eEq. 8
Commonly used anode materials include iridium-coated titanium and platinum-coated titanium.
Gases are evolved by the reactions at both the cathode and anode. These must be removed to
prevent masking the surface of the electrode, which would result in a voltage drop and reduce the

voltage applied to the cells. Removal of the gases is accomplished by maintaining a flow of water
across the surface of the electrodes during operation. This requires the use of a flow compartment
adjacent to the electrode. Such compartments could be either gasketed screen-type spacers or resinfilled compartments.
Since copper wire is commonly used to conduct electric current to the electrodes, the junction of
the copper wire and the non-copper electrode may be subject to corrosion, particularly if it is damp.
If possible, it is best to have a projection of the electrode material that passes through the end plate
of the module to an external connection that can be kept clean and dry (Schweitzer).
Electroactive
Media used
in CEDI
Devices
Ion Exchange Resin Selection
Ion exchange resins function much differently in EDI devices than in a conventional
demineralizer, or even than in a collection/discharge type EDI device. In EDI, the ability of
the resin filler to rapidly transport ions to the surface of the ion exchange membranes is much
more important than the ion exchange capacity of the resin. The resins are therefore not
optimized for capacity, but for other properties that influence transport, such as water
retention and selectivity.
Membrane/resin combinations must also be carefully chosen to selectively catalyze the
electrochemical splitting of water at various locations within the EDI device, as mentioned
previously. Considerable research has gone into optimization of resin fillers for EDI devices,
mostly by the manufacturers of the EDI devices rather than the manufacturers of the ion
exchange resins.
Ion Exchange Membrane Selection
Ion exchange membranes are different from the many types of filtration membranes in that
they are essentially impermeable to water. They combine the ability to act as a separation
wall between two solutions (the diluting and concentrating streams) with the chemical and
electrochemical properties of ion exchange resin beads. Ion exchange membranes are
selectively permeable, as they will allow the passage of counter ions while excluding co-ions.
When placed in a water solution and an electric field, a cation membrane will permit the
passage of cations only, while an anion membrane will allow the passage of anions only. An
in-depth discussion of the theory and properties of permselective membranes is available
elsewhere.
There are two main types of commercially available ion exchange membranes, heterogeneous
and homogeneous. Homogeneous membranes consist of thin films of continuous ion
exchange material, typically on a fabric support. These are essentially equivalent to an ion
exchange resin bead, only in the form of a thin sheet. Heterogeneous membranes consist of

small ion exchanger particles embedded in an inert binder, with or without any support.
Some of the more important properties of ion exchange membranes used in EDI devices
include the following:

Low water permeability


Low electrical resistance
High permselectivity
High strength
Resistance to contraction or expansion
Resistance to high and low pH

Ion exchange membranes that were developed for electrodialysis may not have sufficient
mechanical strength and handling properties for use in assembly of EDI devices, so most
manufacturers have developed special ion exchange membranes that are optimized for their
EDI devices. Extruded heterogeneous membranes based on a polyolefin binder have become
very popular for this application. They are relatively low in cost, offer flexibility in
formulation, and have been shown to be fouling resistant.
Mixed bed Resin Filler (EDI-MB) - Intermembrane Spacing
The first commercial EDI devices used mixed-bed ion exchange resin as a conductive media
in the diluting compartments. For devices using a mixed-bed resin filler, one of the most
important design constraints is the distance between the ion exchange membranes. In order
for the resin to transport an ion to the membrane, there must be a continuous path of the
appropriate type of ion exchange resin, i.e. cation resin for transfer of cations and anion resin
for transfer of anions. For simple cubic packing and equal quantities of equal diameter anion
and cation beads, the probability of a direct conductive path can be related to the number of
resin beads between the membranes by Equation 1.

Eq. 1
This shows that the probability of a direct conductive path decreases as the intermembrane
spacing increases. The effect of intermembrane spacing on salt removal in a EDI-MB device
has also been demonstrated experimentally, as shown in Table 1.
Cell
Thickness,
mm
1.0
2.3
4.7
7.2

Salt
Feed, Product, Velocity,
Removal,
S/cm S/cm cm/sec
%
99.8
600
1.2
0.86
99.9
600
0.6
0.86
94.3
600
34
0.86
71.7
600
170
0.86

Table 1
Relationship between cell thickness and performance for a EDI-MB device
Mixed bed Resin Filler (EDI-MB) - Resin Packing
It has also been shown that the performance of a EDI-MB device can be improved
significantly by the use of uniform particle size ion exchange resins instead of conventional
resins, which have a Gaussian distribution of bead sizes. The uniform beads allow a higher
packing density, approaching a hexagonal close-packed structure. The effect of packing
density on salt removal is illustrated by the data in Table 2.
Product,
Product,
MegOhm-cm
Feed uS/cm
MegOhm-cm
non-uniform
uniform beads
beads
145
0.4
0.7
87
0.8
1.5
65
1.5
4.2
41
3.4
10.5
Table 2
Resin particle size distribution and performance for a EDI-MB deviceLayered Bed
Resin Filler (EDI-LB)
In the late 1980s and early 1990s there was considerable activity in the development of
layered bed (EDI-LB) devices. In this configuration the media comprise separate, sometimes
alternating layers (or in one variation, clusters) of ion-exchange resin, each layer containing
mainly one type of resin: e.g., either anion or cation resin. Liquid to be deionized flows
sequentially through the layers of resins.
For EDI-LB devices there is essentially no "enhanced transfer" regime and less limitation on
the intermembrane spacing. This is because transfer of only one type (polarity) ion is
enhanced at any given time. In order to maintain electroneutrality, the ion that is transferred
out is replaced by a co-ion resulting from splitting of water. This is illustrated in Figure 5.
One of the main design constraints is the choice of ion exchange resin, which must catalyze
the water splitting reaction at the resin/membrane interface. Resin selection must also ensure
that the electrical resistance of the layers is similar, so that the DC current is fairly evenly
distributed through the cell instead of preferentially passing through a single type of layer. It
is likely that the use of uniform particle size resins will offer some benefit to the performance
of thick-cell layered-bed devices, but that the difference will not be as dramatic as it is for a
thin-cell mixed-bed.

One of the main advantages to the use of thicker cells is that it greatly reduces the amount of
ion exchange membrane used to construct the device, which significantly reduces the
assembly cost (both materials and labor). The tradeoff is that the performance for salt
removal is lower than for thin cell devices, due to the higher flow per unit membrane area and
greater distance that ions need to travel across the cell to reach the ion exchange membrane.
The EDI-LB module performance is more sensitive to increases in feed water concentration
and to decreases in feed water temperature. However, this is less important now than when
EDI was first commercialized, due to improvements in reverse osmosis and gas transfer
membranes that have reduced the typical ionic load on the EDI device. The performance of
thick-cell EDI devices is sufficient for their use in most ultrapure water applications, given
proper system design.
The other significant advantage of thick-cell devices is that the thicker resin chambers are
considerably stronger than thin spacers. They also offer more flexibility in the design of the
intercompartment sealing, such as the use of grooves and O-ring seals. This allows
construction of modules without external leaks and with higher pressure rating. The only
commercial EDI devices that are capable of operating continuously at 7 bar (100 psig) are
thick-cell type. Even the spiral-wound devices in a pressure vessel are limited to 4 bar (60
psig) or less.
Separate Bed Resin Filler (EDI-SB)
Another electrodeionization device uses completely separate compartments for the cation and
anion resins, and is somewhat analogous to a two-bed demineralizer. The cation exchange
resin is placed in a compartment between a cation membrane and the anode, with the resin in

direct contact with the electrode. The anion exchange resin is between an anion membrane
and the cathode. The two ion exchange membranes create a concentrate compartment at the
center of the cell. This configuration is shown in Figure 5.

Figure 5
Removal mechanism in thick-cell, separate-bed EDI cell
Instead of splitting water at a resin/membrane or resin/resin interface, this process obtains the
hydrogen (H+) or hydroxyl (OH-) ions needed to regenerate the resin from the electrode
reactions; hydrogen ions being generated at the anode and hydroxyl ions at the cathode
Since the resins are in the electrode compartments, the O2, H2, and Cl2 gas that is created
remains in the product water, which may require an additional gas removal process step. It is
possible that the electrode reaction could produce enough chlorine to reduce the life of the
ion exchange resin, depending upon the amount of chloride in the feed water.
It has been shown that the salt removal by EDI-SB device with 10 mm intermembrane
spacing, is not nearly as good as for a EDI-MB device with 2.5 mm spacing. But the main
disadvantage of the EDI-SB device is that it requires a set of electrodes for each cell. Since
the electrodes are by far the most costly component of a EDI device, this approach is only
cost effective for low flow rate applications where a single cell is sufficient. There have been
some attempts to produce a multi-cell device using bipolar ion exchange membranes, but
these have not been commercialized due to the short life of the bipolar membranes
Thick
and Thin
Cell
CEDI

Thin Cell EDI


The first commercial EDI devices were thin cell with mixed bed ion exchange resin in the
diluting cells. Although they have been modified over the years to improve performance, the
basic principles have remained constant and the technology has proven to be effective and
reliable.
Thin cell, mixed bed electrodeionization devices require a much greater area of ion exchange
membrane per unit volume of water processed, and are therefore not as cost effective as thick
cell devices. The following discussion will attempt to explain the subtle difference in removal
mechanism between the two.
In thin cell, mixed bed EDI, two distinct zones are created inside the diluting compartments.
Strongly ionized substances are removed first and then weakly ionized substances are
removed as the water continues down through the flow path. We refer to these zones as
enhanced transport and electro regeneration respectively.
In the production of ultrapure water, the feed to a EDI device is pretreated with reverse
osmosis. This water contains low amounts of dissolved, ionized solids and some weakly
ionized substances such as carbon dioxide and silica. Because of the low load, the device is
able to remove most of the strongly ionized substances in the enhanced transport zone. Here,
the ion exchange resin simply acts as a conductor to speed the passage of ions from the dilute
compartment through the respective membrane, and into the concentrating compartment. This
is because the ion exchange resin is several orders of magnitude more conductive than the
water. This is shown in the top portion of the resin bed in Figure 3.
After most of the strongly ionized substances have been removed at the top of the cell,
conductance of the diluting cell is maintained by the ion exchange resins. At locations where
the minimum thermodynamic overvoltage for water splitting is applied, the concentrations of
hydrogen (H+) and hydroxyl (OH-) ions are increased. This is shown in the figure in the
electro-regeneration zone. The water decomposition reaction is catalyzed by conditions at the
resin/resin or membrane/resin interfaces of dissimilar polarities. Here, liberated H+ and OHions convert the resins into the regenerated state where weakly ionized substances can react,
become ionized and be moved into the concentrating stream.

Thick Cell EDI


Thick cell EDI devices arrived commercially in 1996, and several different types are now
available. Besides dilute cell thickness, another thing that differentiates these devices from
thin cell EDI is the fact that the dilute cells can use separate resins or a combination of
separate resins and mixed bed resins.
Thick diluting channels can be a detriment to performance using mixed bed resin filler due to
a lower chance of obtaining a continuous path between membranes. Water splitting can still
occur at dissimilar resin/resin and resin/membrane interfaces but much of the split H+ and
OH- ions will recombine when they encounter the counter ions traveling in the opposite
direction. The basis for ion removal is different in devices that use separate resins in the
dilute cells. Because a single type of resin is present at any given point between membranes,
the transfer of co-ions is not possible. Therefore, water must decompose to provide H+ and
OH- ions for transfer to take place while maintaining electrical neutrality. Therefore, current
passage and water splitting are critical for both weak and strong ion removal.
Current
and
Voltage
Faraday's Law
In a continuous electrodeionization device the DC current is the driving force for the removal
of ions, while the applied DC voltage is the means of obtaining the required current.
Faraday's Law states that the electric charge required to liberate one gram-equivalent of a
substance by electrolysis is 96,487 coulombs (a coulomb is the amount of electric charge that
crosses a surface in one second when a steady current of one ampere is flowing across the
surface). In both electrodialysis and electrodeionization, Faraday's Law is used to relate the
transfer of salts through the membranes and the amount of current flowing through the

membranes. A common form of this relationship is given in Equation 2:

Eq. 2
This shows that the amount of DC current required is directly proportional to the flow rate
through the diluting compartments and the amount of ionic equivalents to be removed, and
inversely proportional to the current efficiency.
Current Efficiency and E-factor
Current efficiency can be defined as the ratio of the theoretical minimum current predicted by
Faraday's law (at 100% efficiency) to the actual current applied to the electrodes of the
device, as shown in Equation 3:

Eq. 3
In a EDI device, current that does not cause the transfer of salt will cause water (HOH) to
split into hydrogen (H+) and hydroxyl (OH-) ions, allowing electrochemical regeneration of
the ion exchange resins within the device. For example in a module that is operating at 25%
current efficiency and drawing 4 amps of DC current, 1 amp is causing the transfer of salt and
3 amps are causing water splitting that is unrelated to ion transfer. Cross-leakage and back
diffusion can also cause some current loss, but these are normally small compared to the
water splitting.
In order to produce high purity water (over 1 megohm-cm resistivity) with a EDI system, it is
generally necessary to feed the system with low TDS water such as RO permeate (normally
less than 0.0005 equivalents/liter) and to operate at a current efficiency of less than 35%. For
optimal removal of weakly ionized solutes such as silica and boron, current efficiencies as
low as 5% are sometimes employed.
Some authors prefer to use the term E-factor. This is defined as the ratio of the applied
current to the theoretical current, and is therefore the reciprocal of the current efficiency:

Eq. 4
Ohm's Law and Module Resistance
Ohm's law states that the direct current flowing in an electric circuit is directly proportional to
the voltage applied, and inversely proportional to the resistance of the element:

Eq. 5
Most manufacturers of EDI devices limit the applied voltage to 600 VDC, in order to avoid

the need for the more expensive wiring construction that is required for higher voltages.
Given such a voltage limitation, the electrical resistance of the module therefore controls how
much current can be passed through the cells. Since the DC current determines how much
water can be processed for a given product quality (or what the quality will be for a given
flow rate), it is important to optimize the electrical resistance of the module.
The overall resistance of the EDI module can be affected by the following:

Resistance of the anion-selective membranes


Resistance of the cation-selective membranes
Resistance of the ion exchange resins
Resistance of the concentrate stream
Resistance of the anolyte
Resistance of the catholyte
Feed water temperature
Ionic composition of the feed water

In addition to proper selection of resin and membranes, there are several methods that reduce
the electrical resistance of the cell and therefore allow greater passage of DC current. The
first technique used to accomplish this was to increase the water recovery and therefore the
amount of salt in the concentrate compartments. This is generally done by incorporating a
feed-and-bleed arrangement, using a pump to recirculate the concentrate stream and ensure
adequate flow distribution, while decreasing the flow rate of the bleed that is sent to drain.
An alternative method of reducing the cell resistance is to inject a conductive salt such as
NaCl into the feed to the concentrate compartments using a dosing pump. There are several
possible drawbacks to this method. Increasing the TDS may prevent reclaiming the
concentrate stream for other uses, and may increase the possibility of salt bridging and stray
DC currents. If the concentrate is used to feed the electrode compartment, this can also lead
to generation of chlorine gas at the anode.
A third method is to incorporate resin filler into the concentrate (and in some cases, electrode)
compartments, which eliminates the need for injection of a conductive salt. It has also been
seen that the resin helps ions transfer away from the surface of the concentrate side of the ion
exchange membrane. This reduces the ion concentration in the boundary layer, reducing the
driving force for back-diffusion and improving salt removal.
Ion
Removal
Ionic species are not all removed by the EDI process with equal efficiency. This fact impacts
the quality and purity of the product water.
Easy ions are removed first.
The ions with the strongest charge, the smallest mass, and the highest adsorption to the resins

are removed with the highest efficiency. These typically include: H+, OH-, Na+, Cl-, Ca+2, and
SO4-2 (and similar ions).
In the first section of the EDI module, these ions are removed preferentially to other ions. The
relative quantity of these ions affect the removal of the other ions. The pH approaches 7.0 in
this section since the H+ and 0H- ions become balanced.
The first section of the EDI module is known as the "working bed".

Moderately ionized and polarizable ions are removed next (e.g., CO2).

Carbon Dioxide Removal


This graph shows the effect of CO2 in relatively pure deionized water. CO2 is the next most
common EDI feedwater constituent. CO2 has complex chemistry depending on the local
concentration of protons, and is considered moderately ionized:
CO2 + H2O => H2CO3 => H+ + HCO3- => 2H+ + CO3-2
Eq. 8
Since the pH is forced to be near 7.0 in this section, most of the CO2 is forced into the
bicarbonate (HCO3-) form. Bicarbonate is weakly adsorbed by the anion resin, so cannot
compete with "easy" ions such as Cl- and SO4-2.
In the second section ("Polishing Section") of the EDI module, CO2 (in all of its forms) is
removed preferentially to weaker ions. The amount of CO2 and HCO3- in the EDI feed
strongly effects the final resistivity of the product water and the efficiency of silica and boron
removal.
It is found that as long as CO2(in all forms) is less than 5 mg/l, high quality ultrapure water
can be achieved. If the CO2 concentration is greater than 10 mg/l, it can interfere with the
total removal of ions and strongly impacts the EDI product quality and the silica removal.

Weakly ionized species are removed last (e.g., dissolved silica and boron).

Silica Removal
Since species such as molecular silica are very weakly ionized, and difficult to adsorb on ionexchange resin, they are the most difficult to remove using any deionization process.
If all of the "easy" ions are removed, and all of the CO2 is removed, the EDI module can
focus its force on removing these weakly ionized species. The residence time available in this
third section of the module is important. The longer the residence time available in the
module, the higher the removal efficiency. A long third-section residence time can be
achieved by minimizing the conductivity of the RO product (the quantity of "easy" ions to be
removed) and minimizing the quantity of CO2 in the RO product.
The second and third sections of the EDI module are known as the "polishing bed".
Silica is one of the more important minerals to remove from water for power generation and
semiconductor applications. It also one of the most difficult to remove.

Silica chemistry is complex. On the most basic level, silica comes in "colloidal" and
"reactive" forms. silica level in feedwater depends upon the geology of the region and
whether the source is surface water or well water. Silica in raw water will range from less
than 2 ppm to over 100 ppm.
Physical processes such as reverse osmosis (RO) will remove colloidal silica. EDI will only
remove reactive silica.
Removal of reactive silica depends upon its charge. silica has little, if any, charge at neutral
pH near 7 since the pK1 of silicic acid is 9.8 This makes it difficult to exchange with ion
exchange resin, or to remove with RO or EDI. Raising the pH to above 9.8 helps with the
driving force.
Silica scaling is also an issue. The solubility of silica at pH 6-8 is only 120 ppm at 25oC. this
means that 30 ppm of silica in an RO feedstream with 75% recovery will begin to scale.
There are two prevention techniques for silica scaling. One is the use of an antiscalant in the
RO process, which will delay the precipitation of solid silica. The other is raising the pH,
which increases the solubility limit of silica. At pH 10 silica is soluble up to 310 ppm. Of
course, high pH will cause hardness scaling if the feed is not adequately softened.
It is important to maintain the silica content of the EDI feed stream to under 0.5 ppm as silica
in order to:

Avoid scaling in the EDI concentrate stream


Minimize silica levels in the product water

Typical commercial RO modules will reject silica at only twice the passage of chloride ion.
Most spiral RO module manufacturers claim 99.7% rejection for individual high quality
elements. 99.0% - 99.5% is a reasonable silica rejection for a well designed RO system.
With a 20 ppm silica feed and 75% recovery, a 99.0% rejection element will maintain the RO
product at 0.5 ppm silica.
For higher levels of silica in the feed, the RO system should be designed with higher quality
RO elements and/or lower recovery. Using 99.7% rejection elements and 65% recovery, the
RO feed can approach 90 ppm and still maintain 0.5 ppm silica in the effluent.
Water splitting and Module Resistance
As described above, water splitting is critical for the removal of weakly ionized species like
silica, carbon dioxide, and boron. It is also critical for the removal of strong ions in thick cell,
separate bed EDI. The amount of water splitting can be quantified using Faraday's law to
compare the theoretical amount of current needed to transfer a given amount of ionized
species out of one electrochemical cell to the actual applied current through that cell.
Faraday's law states that

Eq. 9

Where:
I = theoretical current, amps
Eq = number of equivalents transferred per cell
t = time, seconds
F = the Faraday constant = 96,500 coulombs/equivalent
If we define the current efficiency as the theoretical current given by Faraday's law, divided
by the actual applied current, we obtain the following equation:

Eq. 10
Where:
h = current efficiency, %
Ia = applied current, amps
To make this equation easier to use, we can substitute flow rate and feed concentration for the
equivalents removed per unit time per cell so that:

Eq. 11
Where:
TDS = total dissolved solids, mg/l as CaCO3
Q/n = product flow rate per cell, l/min/cell
3.22 = conversion factor
In Eq. 11 we show TDS as the total feed load to the cell but the total feed load also includes
weakly ionized species, such as carbon dioxide and silica, which are also removed. To take
these into account, Ionpure uses a term called Feed Conductivity Equivalent (FCE) and ECell uses a term called Total Exchangeable Anions (TEA). To calculate true current
efficiency, TEA can be substituted directly for TDS in Eq. 11. However, to calculate current
efficiency using FCE the conversion factor changes as shown:

Eq. 12
We can make a couple of important conclusions from Eq. 10 and 11. First, thick cell EDI,
which operates at higher flow per cell than thin cell EDI, will require higher current to
maintain the same current efficiency. Second, current efficiency can be lowered, and hence
water splitting increased, by increasing the applied current at a constant flow rate and feed
concentration. Current efficiency is especially important with regard to weak ion removal. It
is not uncommon to operate below 10% current efficiency for improved removal of silica,
boron, etc.
E-Cell (General Electric corp.) uses a term, which is a variation of current efficiency. They

actually refer to it as the E-factor, which is the inverse of the fractional current efficiency.
Therefore, increasing the E-factor is analogous to lowering the current efficiency.
In any case, increasing the current requires either increasing the voltage or lowering the
module electrical resistance. Increasing the voltage imparts several drawbacks including
greater power consumption and increased safety risk. So reducing the module resistance
becomes paramount in EDI module design.
Because all EDI devices use ion exchange resin in the diluting cells, the concentrating stream
is really the limiting resistance in a module. For modules that don't have resin in the
concentrate, the only way to increase the conductivity is to increase the conductivity of the
water. This is done by increasing the recovery or by direct injection of a salt, such as sodium
chloride, to the concentrate. In addition, many manufacturers use recirculation of the
concentrate stream along with high recovery or salt injection to provide a less variable
concentration along the length of the module.
There are several drawbacks to concentrate recirculation. Concentrate recirculation requires
the use of a pump and additional ancillary equipment for control such as motor starters and
throttling valves. This adds complexity to the system design and increases overall cost. In
systems with fluctuating operating conditions, operation or adjustment of the pump can make
the process significantly more labor intensive. Also, the power required to operate the pump
can be a large portion of the total power consumption of the system. A typical industrial EDI
system with concentrate recirculation would consume about 1 to 2 kilowatt-hours per
thousand gallons of product water (kWh/kgal), where about 0.5 kWh/kgal is for the
recirculation pump alone.
Salt injection also has several drawbacks. Increasing the salt concentration in the EDI
concentrate stream:

Limits the ability to recover that stream


Increases the concentration gradient between the diluting and concentrating cells
facilitating co-ion back diffusion if the membranes are not ideally permselective
Increases the possibility of salt bridging and electrical shorting, and

Leads to the formation of chlorine at the anode when fed with recirculated concentrate
water.
Gas
Generation
in CEDI
Cells
Chlorine Gas Production

Normally the reaction at the anode will result in the formation of oxygen, but with sufficient
voltage and with chloride ion present, chlorine gas can be formed. In particular this is an
issue with electrodeionization systems which inject sodium chloride into the concentrate and
electrode streams in order to reduce module resistance and improve performance. Because

many EDI devices feed the reject water to the electrode chambers, concentrate salt injection
increases the chloride concentration at the electrode. As shown below in Eq. 13, chloride at
the anode is converted to chlorine, which is a strong oxidant and a toxic gas that can damage
ion exchange membranes and create a safety hazard for those near the device.
2Cl- Cl2 + 2e-

Eq. 13
To investigate the extent of chlorine formation, an E-Cell MK-1 module was operated at 600
volts DC and 3.0 amps. The temperature of the feed water was varied between 15 and 25oC
such that the resistance of the module changed proportionally. Then the concentration of
sodium chloride was varied at the inlet of the concentrate cells to overcome this change in
resistance and maintain 3.0 amps. Free chlorine was measured in the electrode product water.
As shown in Figure 7, a significant amount of free chlorine was detected. There was also a
large off-gassing of chlorine evidenced by the odor near the device. Concentrations in the
range shown in Figure 7 can cause significant damage to ion exchange membranes or resins
in contact with the electrode stream. In particular this is an issue with electrodeionization
systems which inject sodium chloride into the concentrate and electrode streams in order to
reduce module resistance and improve performance.

Figure 7. Chlorine Generation at the Electrodes in an E-Cell Module


Other systems do not require brine injection because they use an ion-exchange resin filler in
the concentrate compartments in order to minimize the electrical resistance of the module.
This in itself limits the possibility of chlorine generation at the anode by reducing the
chloride concentration. Because the resin is orders of magnitude more conductive than typical
RO product water, it removes the conductivity of the water as a contributing factor in the

overall resistance. This alleviates the need for salt injection or recirculation. In addition,
chlorine production is virtually eliminated. An additional advantage of the this design is that
with elimination of salt injection and chlorine production, the EDI concentrate will generally
be better quality than the raw water and can often be recycled.
To show the effect of concentrate resin filling, two modules were tested with and without salt
injection under similar conditions. Shown in Figure 8 is the resistance per cell pair for a
module with concentrate resin filling compared to a module without concentrate resin filling,
as a function of temperature. The electrical resistance of the module with a resin filled
concentrate was lower by two orders of magnitude.

Figure 8. The Effect of Concentrate Resin Filling on Module Resistance


Hydrogen Gas Production
People often express concern about the hydrogen gas produced at the cathode, since it is
known that under certain conditions hydrogen can be explosive. However, the amount of gas
that is produced by an EDI module is so small that it does not present a safety hazard when
the EDI system is installed in an area with normal ventilation. Codes require that buildings
have multiple air changes every day, one air change being a turnover of air equivalent to the
building's internal volume. The number of air changes varies depending upon the building's
use and the local codes, but a widely accepted value is half an air change per hour.
To show how little risk is presented by the electrode gases, let's perform an example
calculation assuming that an EDI module is installed in a 4m x 4m x 4m office which has the
normal ventilation of half an air change per hour. This is equivalent to an air flow of 32 m3/h.
Assume the EDI module is operating with a DC current of 10 amps, therefore producing
hydrogen at the rate of 74.6 ml/min (equivalent to 0.0045 m3/h). If all the hydrogen gas
leaves the water and enters the room air, then the resulting concentration of hydrogen in the

air would be about 0.014% (v/v), or about 141 ppm. This is well below the explosive limit of
a hydrogen/air mixture, which is 4.2% v/v at STP, and also well below the concentration at
which asphyxiation would occur. Even at the maximum current output of an EDI DC power
supply there is no safety risk as long as the ventilation meets typical building code
requirements.
Advanced
Index
The advanced section is comprised of the following EDI topics:
Pretreatment to EDI systems - EDI Feedwater Requirements
Critical contaminants that adversely affect the EDI process include hardness (calcium,
magnesium), organics (TOC), particulates and suspended solids (SDI), active metals (iron,
manganese), oxidants (chlorine, ozone), and carbon dioxide (CO2). The pretreatment process
designed for the RO/EDI system should remove these contaminants from the feed water as
much as possible. Proper pretreatment design will greatly enhance EDI performance.
Suggested water treatment strategies are listed throughout this section. The main areas that
affect EDI operation and performance are:

Dechlorination
CO2 - Carbon Dioxide
Particulates, Metals and TOC

EDI Process Considerations

EDI system Safeguards


EDI Process Design
Feed Water Hardness
Design to prevent EDI scaling
Avoiding a common mistake
Water Recovery
Product Water Quality
Reclaim of EDI Reject Water
Boiler Makeup Water (Power Plant)
Other EDI "Watchouts" Feed Water
Other EDI "Watchouts" Hydraulic
Feed Water Conductivity Equivalent (FCE)
DC Volts and Amps
Definition of EDI Recovery
Calculating Reject Flow
Suggested Test Kits

Factors Affecting EDI Performance

Voltage

Current
Ionic Species
Temperature
Flow
Feed Conductivity
Effects of contaminants

Optimizing EDI Performance

Voltage Driving Force


Current density
Ionic Balance and pH

Troubleshooting EDI Systems

Some Very Important Points to Remember!


Requirements for EDI troubleshooting
Questions To Ask
EDI Troubleshooting: Data Required
EDI Troubleshooting: Low Product Quality
Mass Balance Calculation
More going in than coming out
EDI Troubleshooting: Resistance Increase
EDI Troubleshooting: Low Flow
Flowcharts
Oxidants
Scale
Organic Fouling
Biofilm
Excessive Feed Pressure
High Temperature
Particulates

EDI Cleaning and Sanitization

Purpose of cleaning
When to Clean?
Chemicals Used on EDI
For Any Cleaning/Sanitization
Typical Cleaning System
General Considerations

Multiple Cleanings
HCl Cleaning Procedure
Brine/Caustic Cleaning
Sanitization: Definitions
Caustic Cleaning/Sanitization
Sodium Percarbonate Cleaning/Sanitization
Multi-Agent Cleaning ("3 Step")
Peracetic Acid Sanitization
Hot Water Sanitization
Hot Water Sanitization "watch-outs"
Which Cleaning for High d/p?
Which Cleaning for High Module Resistance?
Action for Quality Decline with no Other Symptoms?

Shutdown/Preservation
Pretreatment Dechlorination
The main cause of EDI failure is oxidation of the module, either by chlorine. Oxidation
decrosslinking also causes the resins to disintegrate which results in an increased pressure
drop across the module. For this very reason, dechlorination is extremely important to put the
proper measures in place in the pretreatment to both the RO and EDI system. In addition to
chlorine, chloramine or ozone attacks ion-exchange resins and membrane and cause
decrosslinking, which results in reduced capacity. If Chloramines are present, further methods
of removal may be required. Oxidation will increase apparent TOC, and the byproducts
cause fouling of the anionic resin and membrane, reducing the ion-transfer kinetics.

The ideal concentration level for oxidants is zero. RO membranes are more resistant to
chlorine that EDI modules so do not measure the decrease of RO rejection as an indication
that there is no chlorine present!
Rule #1: There should be no detectable chlorine in the EDI feed water. Oxidative attack
on a EDI module is not recoverable by cleaning!
There are three principal means of achieving the above requirement:

Granular activated carbon


Injection of sodium bisulfite or sodium sulfite
UV dechlorination

Recommendation: Activated carbon dechlorination is preferred over chemical


dechlorination because it is both more reliable and also less likely to cause biological fouling
problems. UV dechlorination is a relatively new process, which may have difficulty
achieving complete chlorine removal, especially on chlorinated feed water.
Feed water chlorine can be measured with a test kit. The smallest increment for this test kit is
0.02 mg/l.
A) Dechlorination by activated carbon
It is the responsibility of the system supplier to ensure that the activated
carbon filter is sized properly for complete removal of the feed water chlorine,
taking into account the concentration of chlorine in the raw feed water.
Ionpure would usually design dechlorination carbon filters for a minimum
empty bed contact time of 6 minutes.
B) Dechlorination by chemical injection (reducing agent)
It is the responsibility of the system supplier to ensure that the chemical
injection system is sized properly for complete removal of the feed water
chlorine, and includes adequate instrumentation and control safeguard to
ensure that the EDI system is not fed water containing chlorine should there be
a change in the feed water conditions or an upset in the operation of the
chemical injection system.
In systems that contain injection of multiple pretreatment chemicals (for
example antiscalant and sodium sulfite or sodium bisulfite and sodium
hydroxide) it is recommended that there be a separate injection point and static
mixer for each chemical.
C) Ultraviolet Dechlorination
It is the responsibility of the system supplier to ensure that the UV
dechlorination system is sized properly for complete removal of the feed water
chlorine, taking into account the concentration of chlorine and chloramine in
the raw feed water.
SUMMARY

There must be no detectable chlorine in EDI feed water!


Chlorine damage is irreversible
Chlorine is a common cause of EDI failure

Take Cl2 readings at RO inlet and EDI inlet


RO permeate can contain more free chlorine than RO feed

Suggested Test Kit: Use Cl2 Hach Model CN-70 (#1454200)


Pretreatment
- Hardness
Hardness
Hardness can cause scaling in the reverse osmosis and EDI units, If this occurs, it will take
place in the concentrate chambers at the high-pH surfaces of the anionic membrane. This
formation of scale which in turn increases the electrical resistance of the module, as well as
the pressure drop in the concentrate, and current efficiency is reduced. Minimization of inlet
hardness will lengthen the time between cleanings.
All EDI devices tolerate a maximum feed hardness of 1 ppm as CaCO3 and some
manufacturers recommend as little as 0.1 ppm as CaCO3. In some instances, feed hardness
limits the EDI system recovery. If the feedwater source has little to no hardness, a softener
may or may not be required. If a softener is required, there are two main methods of
removing the hardness.
RULE#1 - Reduce the hardness feeding the EDI module to less that 1 ppm (the lower
the better)
Reduction Methods
1) Pre-RO Softener - This design protect the RO from scaling as well as the EDI system.
The downside is that more salt is used than post-RO softening.

2) Post RO Softener - This design has advantage of being able to use a smaller softener and
less salt vs. Pre-RO softening.

3) Antiscalent Injection - This method involved the dosing of chemicals but is not
recommended due to higher operator attention and more room for error.
SUMMARY

All EDI devices sensitive to hardness Most tolerate maximum 1 ppm as CaCO3
Some recommend 0.1 ppm as CaCO3
Hardness may limit recovery
LSI not good predictor of scale potential
Hardness scale increases module electrical resistance, reduces product water quality,
increased reject pressure drop and cleaning frequency

Suggested Test Kit - Hardness Hach Model HA-71A (#145201)


Pretreatment
- Carbon
Dioxide
Carbon Dioxide and EDI
Carbon dioxide is typically exists in a dissolved state in the feedwater. Measures to reduce the CO2
in the EDI feed should be taken because CO2 is in fact an ionic load on the EDI module. Carbon
dioxide must be measured and ADDED to the Ionic load in the RO permeate.
Tip: Each ppm of CO2 as CO2 =2.66 S/cm OR ppm as CO2 x 2.66 = S/cm
Reduction Methods
1) Caustic Injection before the RO
One way to remove dissolved CO2 is to raise the pH prior to the RO to convert the CO2 to
carbonate. The carbonate is then rejected by the RO membrane and sent down the drain. Usually a
small pH increase to 8-8.5 is required. See the chart below for the relationship of bicarbonate
alkalinity and CO2 on pH.

The higher the pH, the higher the bicarbonate and lower free carbon dioxide. This method is used
more often due to the minimal cost incurred by adding a chemical injection system. For end users
that do not wish to injection chemicals, see below.
2) Degasification
The other method of reducing dissolved carbon dioxide is through degasification. There are two
main methods, 1) forced draft, and 2) membrane degasification. One manufacturer of membrane
degassing modules is Membrana with their Liqui-Cel product.

SUMMARY

Reduce the dissolved CO2 feeding the EDI system as much as possible

Suggested CO2 Test Kit - Hach Model CA-23 (#143633), smallest increment 1.25 mg/l
Pretreatment
- Other
Particulates, Metals and TOC
Organics (TOC) adsorb to the surfaces of the resins and membranes. This causes fouling of
the active sites. Fouled resins and membranes are inefficient at removing and transporting
ions. Stack electrical resistance will increase.
Particulate matter (SDI), colloids, and suspended solids cause plugging and fouling of the
membranes and the resin in the chambers. Plugging of resin increases the pressure drop
across the module.
Iron and other active metals (e.g., Mn) may catalyze resin oxidation, and may strongly and
permanently adsorb to, and reduce the capacity of, the internal resins and membranes. This
happens even in sub-ppm concentrations.
Process Considerations
Feed Water Conductivity Equivalent (FCE)
The concept of EDI feed water conductivity equivalent arose from the need for a simple field
method of estimating the ionic load on a EDI device. Certainly the best way to determine the
ionic load is to perform a complete water analysis and determine the concentration of all
ionized and ionizable constituents, but in some cases this is not practical. What had
frequently been substituted for a complete water analysis was a simple measurement of the
conductivity of the EDI feed water. This could introduce considerable error, as a conductivity
measurement does not detect the full amount of weakly ionized species such as carbon
dioxide (CO2) and silica (SiO2). For example, 10 S/cm water could contain 4 ppm of NaCl
or 60 ppm CO2. Just as this would have a huge impact on the service cycle of a conventional
demineralizer, it can have a major impact on sizing a EDI system. For this reason we have
developed the concept of EDI feed conductivity equivalent, which attempts to take into
account weak ions such as CO2 and silica.
Calculating FCE

Measure EDI feed water conductivity (S/cm)


Measure the ppm CO2
ppm as CO2 x 2.79 = S/cm
Measure the ppm SiO2
ppm SiO2 x 1.94 = S/cm
Add measured conductivity to CO2 & SiO2 S/cm

In the field, the CO2 concentration in the EDI feed water can be measured using Hach test kit
Model CA-23 (#143601). The smallest increment for this test kit is 1.25 mg/l. The EDI feed

water conductivity can be measured with a hand-held conductivity meter such as the Myron L
Model 4P or with the permeate conductivity meter on the RO system.
Example FCE Calculation

Measured conductivity = 6 S/cm


Measured CO2 = 5 ppm as CO2
Measured SiO2 = 0.5 ppm as SiO2
FCE = 6 + (5 x 2.79) + (0.5 x 1.94) = 20.9 S/cm

DC Volts and Amps

Voltage (potential) causes current to flow


Current causes transfer of salt, regeneration of resin
Amount of current required proportional to product water flow, amount of salt being
removed
Use Faradays law to calculate current required

Calculating Amps Required

Faradays Law
o I = 1.31 (Q)(FCE)/(# cells)(eff)
o I = DC current, amps (per module)
o Q = product flow rate, liters/min/module
FCE = feed conductivity equivalent, S/cm
eff = current efficiency, % (assume 10% current efficiency)

Example DC Current Calculation

Product flow = 50 lpm


Feed = 5 S/cm + 3.75 ppm CO2
FCE = 5 + (3.75 x 2.79) = 15.5 S/cm
I = 1.31(50 lpm)(15 mS/cm)/(24 cells)(10%)
= 4.2 amps per module

Setting DC Amperage

Use constant current power supply


Estimate amps required per module
Set amperage to calculated value
Power supply adjusts voltage to maintain current
Important to record (trend) voltage & amperage

Definition of EDI Recovery

% Recovery (R) = (QP)(100)/(QP + QR)

QP = Product (Dilute) flow rate


QR = Reject (Concentrate) flow rate

Calculating Reject Flow

Reject = ((100 - R)/R) x product flow


o where R is the percent recovery
Example: Product flow = 50 m3/h
o Recovery = 90%
o Reject = ((100-90)/90) x 50
o

= 5.6 m3/h

Silica &
Boron
Silica and Boron are two of the most difficult ions to remove in a feedwater due to their
weakly ionized nature. Both are usually are one of the last to be removed and the first ions to
pass through an EDI device. In addition, if the feed silica is high, silica scale can form in the
EDI module which in turn increases the module resistance and affects the removal efficiency
of the electric field. As a result, product quality decreases. Below are some guidelines when
silica and boron are an issue.
Water Recovery
EDI systems that include ion exchange softening can usually operate at 95% recovery.
EDI systems that DO NOT include ion exchange softening must operate at 90% recovery or
below.
NOTE: In some cases the recovery may be limited by the concentration of silica in the feed
water, where the maximum recovery can be calculated as follows:

Maximum recovery, % = 100 (4 x ppm feed silica)

In the field, the silica concentration in the EDI feed water can be measured using a silica test
kit.
Suggested Test Kit:

Silica Hach Model SI-7 (#2255000)


o Smallest increment 0.02 ppm
o Optional (for power plant applications)

Boron - Field samples sent to lab or use Boron analyzer

Mechanical & Control


Basic Flow Path
Ok, lets start with the basics, what does the flow path through an EDI module look like?

The Feed (RO Permeate) is split into two inlet streams, 1) Dilute Feed and 2) Concentrate Feed. There
are two outlet streams, 1) Product and 2) Reject.
To obtain the best EDI systems have to offer (consistent and continuous chemical-free operation),
proper pretreatment design and EDI system design is a MUST! Shortcuts should not be taken!
EDI system Safeguards - There are several key safeguards that all EDI system should have in place to
protect the EDI modules from improper operation. Some are required, some are highly recommended.

Required safeguard
o Low flow alarm for product and reject flow - shown below as FE for Flow Element and
FAL (Flow Alarm Low)
o RO (or EDI feed pump) interlock so EDI runs only when RO or feed pump is in
operation - Shown on DC power controller as an interlock.

Optional safeguard
o Product divert valve (Sends low product quality to drain)
o Low feed quality alarm (Protects EDI from high hardness or chlorine/ORP), shown as
AE for Analysis Element and AL for Alarm Low

Water Recovery
EDI systems that include ion exchange softening can usually operate at 95% recovery.
EDI systems that DO NOT include ion exchange softening must operate at 90% recovery or below.
NOTE: In some cases the recovery may be limited by the concentration of silica in the feed water,
where the maximum recovery can be calculated as follows:

Maximum recovery, % = 100 (4 x ppm feed silica)

In the field, the silica concentration in the EDI feed water can be measured using a silica test kit.
Suggested Test Kit: Use Silica Test Kit 0.02ppm (20ppb) Hach Model #2255000
Reclaim of EDI Reject Water
There is normally no problem using the EDI reject water for other purposes elsewhere in the plant as
long as it is sent to an atmospherically vented tank to allow venting of hydrogen gas.
There are two potential issues with the recycling the EDI reject stream to the feed of the RO system:
carbon dioxide (CO2) and hydrogen (H2).
1) CO2 in EDI Reject Water:
Feed water CO2 will pass through the reverse osmosis system, be removed by the EDI
module and transferred to the EDI reject stream. Recycling the EDI reject to the feed of
the RO system will typically result in about a 3X increase in the EDI inlet CO2
concentration, which will affect the EDI performance. This must be taken into account
when performing the EDI performance projection. The use of degasification or caustic
injection pre-RO can prevent the 3-fold increase in CO2 concentration.
H2 in EDI Reject Water:
To prevent the buildup of hydrogen the EDI reject must be returned to an
atmospherically vented tank. Do not return the EDI reject directly into a pressurized line.
Summary - EDI Reject Reclaim

Reject contains electrode gases & CO2


Returning reject to RO feed increases EDI feed CO2 by up to 3X (if no degasification or NaOH
injection before RO)
May affect EDI product water quality

Other EDI "Watchouts" Hydraulic

Water hammer
Simultaneous operation of product and divert valves
Should open one before closing other
Pressure differential between compartments
o

Keep pressure slightly higher on product side (2-5 psi recommended)

Factors Affecting Performance


Voltage
Voltage is the driving force, which pulls the impurity ions from the feed streams into the
concentrate streams. The local voltage gradients also cause H2O to split into H+ and OH- ions.
The constant formation and high local concentration of these ions allows the state of the
resins in the polishing section within the EDI module to be in the full hydrogen and hydroxyl
form, and fully able to remove weakly ionized species such as CO2 and silica. These also
prevent the growth of bacteria within the EDI module. Excess H+ and OH- ions are pulled
from the feed streams into the concentrate streams, which also compete with any impurity
ions for transport sites.
Optimum Voltage
The optimum voltage depends first on the number of cells in the module. Normal operating
voltage range is approximately 5 to 8 Volts/cell. The optimum voltage also depends on:
1. Temperature
2. Concentrate conductivity
3. Concentrate flow rate (recovery).
Quality vs. Voltage
There is an optimum voltage to achieve the highest quality water. At voltages lower than this,
the driving force is inadequate to move the ions across the purifying chamber resin bed and
then across the membranes before the product stream exits the module. At voltages higher
than optimum, the overvoltage creates excess water splitting and therefore excess current, and
also causes ion polarization and thus back diffusion, which lowers product water resistivity.
Within the range set for each module type, the optimum voltage depends on the ion load and
on the water recovery rate. Higher ion loads in the feed and higher recovery rates lead to a
higher ion concentration in the concentrate chambers, which lowers the resistance of the
module the lower stack resistance leads to a lower optimum voltage.

Current
Typical current draw for an EDI module at nominal voltage is 2-4 amps with a feed
conductivity of 4-10 uS/cm. Current may be as low as 1 amp or less. Current at high feed
conductivities (20-30 uS/cm) will lead to currents as high as 8 amps, or higher.
Fundamentally, current is proportional to the total number of ions moved. These ions include
the impurity ions in the RO permeate, such as Na+ and Cl-, and the ions caused by water
splitting, H+ and OH-. The water splitting rate depends on the local voltage gradients, so that
higher voltages across the resin chambers leads to higher quantities of H+ and OH- to be
moved.
A portion of the current is then directly proportional to the ion content of the feed (TDS, or
uS/cm). The other portion of the current, proportional to water splitting, increases nonlinearly with overvoltage. The current efficiency is the fraction of the total current that is
required to move the impurity ions in the EDI feed.
If the module current is higher than expected, it could be because the voltage is set higher
than optimum, and excess water splitting results in the excess current.
Current also depends on the concentrate flow, and therefore on the module water recovery.
The nominal concentrate flow is 10% of the feed flow. If the concentrate flow is lower than
recommended, the concentrate will be more conductive and the current will increase.
Steady State Operation
Normally, an EDI module will start up with high quality product water. This is because the
EDI module has excess mixed bed ion-exchange resins in it in the H+ and OH- form, in the
polishing section.
However, after operating conditions have changed, a module will take between 8 and 24
hours to reach a new steady state. The true steady state is defined as reaching a mass balance
on the ions entering and leaving the module. At steady state, the kinetics of ion migration
match the ion feed rate. Steady state for trace ions such as silica may take as long as 2-4
weeks.
If voltage is lowered or ion load is raised, the internal ion-exchange resins will begin to
adsorb the excess ions. In this condition, fewer ions leave the module than enter. Eventually a
new steady state is reached. During this time, the working ion front progresses in the resin
bed from near the bottom of the module upward.
If voltage is raised or ion load is lowered, the resins will lose some of their excess ions to the
concentrate stream, and more ions will be exiting the module than enter it. During this time,
the location of the working ion front grows closer to the inlet of the module. This latter is the
mechanism of the regeneration procedure.
An ion balance done on the module(s) during operation is a valuable tool in determining if
the EDI system is operating at steady state.

At Steady State: Total Ions Out = Total Ions In


Module Filling with Ions: Total Ions Out < Total Ions In
Module Recovering from Overload: Total Ions Out> Total Ions In
Ionic Species
The ability for EDI to remove ions from a stream depends in part on the properties of the
ionic species. In a standard resin bed, the adsorption strength and kinetics depend on the ionic
size, the degree of hydration, and on the type of resin.
In EDI, the ionic charge is even more important since this is the driving force to move the
ions along the resin surfaces to the membrane, and through it.
Ionic Size
The following ionic sizes are the effective sizes in aqueous solution at 25oC. These sizes
include full hydration. The larger the effective size the slower the diffusion rate larger ions
are removed by EDI less well. The larger the effective size, the more distributed the charge
and the less well adsorbed by the resin.
Ionic Radius

Cations

Anions

<3.0

NH4+,K+

CL-, NO3-

3.5

OH-, F-

4.0-4.5

Na+

6.0
8.0-9.0

Li+, Ca+2, Fe+2


H+ Mg+2, Fe+3

SO4-2, CO3-2

Ionic Charge
The higher the ionic charge the stronger the applied voltage will pull the ion through the
membrane. This is counterbalanced by higher degrees of hydration and larger, heavier
molecules which slow diffusion.
Selectivity Coefficients of Ions for Resin
The table below shows the selectivity of different ions for resin. This is a measure of their
adsorption strength to the resin. Strong adsorption means low leakage through a resin bed or a
EDI module.

Cation
Li+
H+
Mg+2

Selectivity
Selectivity
Anion
Coefficient
Coefficient
0.8
HSiO31.0
F0.1
1.2
HCO30.5

Na+
Ca+2
NH4+
K+

1.6
1.85
2.0
2.3

OHClNO3I-

0.6
1.0
3.3
7.3

Easy Ions (Na+, Cl-, Ca+2, H+, and OH-)


Sodium (Na+), Chloride (Cl-), Calcium (Ca+2), Hydrogen ion (H+), and Hydroxyl ion (OH-)
are considered easy ions for EDI. All of these ions are adsorbed well by the resin and have a
charge that is definite and difficult to polarize. These ions are fairly easy to remove in the
working section of the EDI module.
Large, Weakly-charged Ions (Carbon Dioxide, Silica, Boron)
Carbon Dioxide (CO2), Silica (SiO2), and Boron (H3BO3) all have weak anionic charge under
normal operation and pH. Because of this, they are weakly adsorbed to the resins and the
applied voltage has little driving force.
To effectively remove these ions, other system strategies are used.

Minimize the ionic content of the feed


Minimize the CO2 in the feed
Maximize the removal of silica and boron by the RO

If the total ion load to the EDI is lowered, the working section of the module will be small,
and the polishing section relatively larger. The larger polishing section will aid the removal of
the hard-to-remove ions.
CO2 can be removed by the RO if the pH of the feed is raised (the pK1 of carbonic acid
(H2CO3) is 6.35). Hence, with moderately high pH, bicarbonate ion can be removed, Of
course, hard cations (Ca+2, Mg+2) must be removed first to operate the RO at high pH. CO2
can be removed as a gas after the RO (see Liqui-Cel), which is helped with low pH (so the
CO2 is not in ionic form). The pK1 of silicic acid (H2SiO3) is 9.77. The pK1 of orthoboric acid
(H3BO3) is 9.28. only with pH > 10 can silica and boron be removed effectively. This is the
theory behind the HEROtm process.
Raising the pH of the EDI feed is counter-productive. Since Na+ and OH- ions are "easy ions
to remove, the addition of NaOH before the EDI simply raises the ion load for the working
section of the EDI, and the pH returns to 7.0 by the end of the working section. Further, the
size of the polishing section is now smaller.
Temperature
Pressure Drop vs. Temperature
Pressure drop depends on temperature mostly due to the effect on the viscosity of water. The
table below shows the absolute viscosity of water (cP) at temperatures of interest, and the
relative viscosity (based at 25oC). Pressure drop will increase or decrease proportionally to

the viscosity. Note that at 5oC the viscosity of water is 70% higher than at 25oC.

Temperature

Relative
Viscosity

Absolute
Viscosity (cP)

5oC (41oF)

+70%

1.51

15oC (59oF)

+28%

1.14

20oC (68oF)

+12%

1.00

25oC (77oF)

0.89

30oC (86oF)

-10%

0.80

35oC (95oF)

-19%

0.72

Stack Resistance vs. Temperature


As temperature increases, the resistance of the stack will decrease. At a given voltage, current
will increase. One cause of this phenomenon is increased ionic activity at higher
temperatures. All other things being equal, the stack resistance will change about 2% per 1oC.
The quality optimization will depend on other factors (below), and so the optimum setting of
the voltage will change with temperature.
Quality vs. Temperature (re-optimization of operating conditions)
There is an optimum temperature for operation.
As temperature increases to 35oC, product quality will generally increase since ions are more
mobile and move more easily. Higher than this, quality will lessen due to increases in ionic
leakage. This is caused by lower adsorption of the ions to the internal ion exchange resins.
In addition, the actual ionic resistivity, uncompensated for temperature, will increase and
there is less accuracy in the reading (see section below).

A lower voltage is required at higher temperatures to move the ions into the
concentrate.

As temperature decreases toward 15oC, product quality may lessen. Some of this is due to
errors in resistivity measurement temperature compensation; some improvement is due to the
greater adsorption of the ions to the internal ion exchange resin. As temperature decreases
further, the activation of diffusion through the membrane will become larger and quality will
decrease.

At low temperatures, a higher voltage will be needed to continue to split water


effectively, and move sluggish ions faster.

Resistivity Measurement Correction with Temperature


Resistivity measurements change strongly with temperature, and are normally corrected to a
standard temperature (25oC). Impurity ions in water have a higher electrical conductance at

higher temperatures because the ions are more mobile. Similarly, Ultrapure water has a lower
electrical resistance as temperature is raised because water dissociates into H+ and OH- more.

The correction for temperature in the meters is large, and is subject to errors. A high
quality resistivity meter is recommended.

The correction for conductivity with temperature for tap water and RO permeate is about
2%/oC. The correction for resistivity with temperature for ultrapure water is 5-7%/oC. In both
cases the temperature correction is large and large errors may be introduced. Accurate
temperature compensation becomes more important as working temperatures are different
from 25oC
Hot Dl water is the most difficult to measure accurately.
Temperature, C

Uncompensated
Resistivity,
Megohm.cm

15

31.8

25

18.2

35

11.1

At low temperatures, a higher voltage will be needed to continue to split water


effectively, and move sluggish ions faster.

Flow
Pressure Drop vs. Flow
There are two or three module pressure drops to consider, depending upon module
construction:
1. Feed to Product
2. Concentrate Inlet to Outlet
3. Electrolyte Inlet to Outlet (not present on all systems)
The pressure drop will increase on each of these streams as the flow to each is increased.
Pressure drop is defined here as measured near the inlet and outlet fittings of the module.
Electrolyte Pressure Drop: at 0.05 gpm (10 lph) the pressure drop will be approximately 20
psi (1.4 bar). If pressure drop rises above this, then the inlet may be fouled or blocked with
debris. The inlet water must be finely filtered. This flow should be independent of the size of
the module and the number of cells since there is only one anode/cathode per module.
Concentrate Pressure Drop: The concentrate flow will be different for each design, for each
operation, and for each EDI model. Typically, the concentrate outlet be set between 5%
and10% of the EDI product flow.
If the concentrate pressure drop increases during operation, it may need cleaning or it may

have debris in the concentrate inlet. The inlet water must be finely filtered (RO effluent).
Feed-Product Pressure Drop: The Feed-Product pressure drop increases with flow. The
pressure drop is close to being linear (first-order) with flow; that is, twice the flow will cause
twice the pressure drop.
For a new module, typical pressure drop will be as low as 10 psi (0.7 bar) at the low end of
the flow range or as high as 60 psi (3.4 bar) at the high end of the flow range

Pressure drop will change dramatically as the water temperature varies from 25oC.
There can be substantial pressure drop in manifold lines. valves, flowmeters, solenoid
valves, elbows, and tees.

Effect of Outlet Pressure on Quality and Internal Leakage


Plate-and-frame modules are sealed with internal gaskets, and there will be some internal
leakage. In an EDI module, if the concentrate leaks into the product stream then the product
resistivity will suffer.

Product Outlet pressure must be greater than Concentrate Outlet pressure.

To ensure that internal leakage does not impact product quality, the product outlet must have
a higher pressure than the concentrate or electrolyte streams outlets. This way, any internal
leakage will not add ions to the product stream.
For the simplest, easiest system, there should be no back-pressure applied to the concentrate
stream outlet. In systems with valves to manually control the concentrate backpressure, the
result is often complication and operator error.
To send the concentrate stream to the inlet of the RO, the outlet is ideally first plumbed into
an ambient break tank then pumped independently into the RO inlet pretreatment line.
When this is done, the EDI module can approach 99% recovery.
Feed Conductivity
Product Quality (at design and maximum flow):
The product quality depends on the ability of the module to remove ions from the purifying
chamber before they exit the module. More feed ions will result in lower product quality. This
is true for both general ionic conductivity (NaCl) and weak ions (silica, boron, and
bicarbonate).
Additional ions add a load that has two results. The first is that the depth of the working bed
within the EDI module lengthenscausing the polishing bed to shorten. The first quality
deterioration occurs in lower removal of weakly charged species.

Reducing feed conductivity helps improve silica and CO2 removal.

The second result is that the module current draw increases as feed conductivity increases.
Moving more ions takes more electrons. The current increase is not linear because the current

also moves water, which has been split.

Increased feed conductivity increases current draw.


Optimizatio
n of CEDI
Systems

Voltage driving force


There is an optimum voltage for each operating condition. One may apply too much or too
little voltage for a specific condition. A typical voltage range is given for each modulethe
optimum should be in this range.
If the voltage is too low, then the driving force is too low and not enough ions can be moved
from the feed stream into the concentrate stream. In addition, water is not being split
effectively and the ion exchange resins in the polishing bed section may not be free of
impurity ions and fully capable of trapping and moving silica for example.
When the voltage is initially lowered, the ion-exchange bed within the module will begin to
fill with ions until it reaches a steady state. During this time, more ions will be entering than
leaving the module. The symptom is that the concentrate will not have as many ions as
normal. Steady state may take 24 hours to achieveduring this time, product resistance will
decline slowly.
If the voltage is too high, then too much water is split and the voltage driving force becomes
ineffective. A symptom of this is excess gas production in the electrolyte stream, and excess
current draw. Excess voltage also causes a phenomenon called "concentration backdiffusion". In this state, ions are forced to diffuse from the concentrate streams into adjacent
dilute chambers to achieve electrical neutrality.
When the voltage is initially raised, the ion-exchange bed within the module will begin to
discharge ions until it reaches steady state. During this time, more ions will be exiting than
entering the module. The symptom is high conductivity in the concentrate stream. Steady
state may take 24 hours to achieve. During this time, product resistance will improve slowly.
Current densities
The current density in the working section of the bed is high, caused by movement of the
primary ions in the feed. There is relatively high electrical resistance in the concentrate
stream since the water there is RO-quality water with 4-30 uS/cm.
In the polishing section of the bed, the concentrate stream is full of the ions it picked up in the
working bed. At 90% recovery, the conductivity will be about 10 times the feed
concentration. It can therefore be in the range 40-300 uS/cm. The voltage drop is now higher
across the resins in the purifying chambers (where there are very few feed ions remaining).
The net result is a higher rate of splitting water and higher concentration and transfer of
protons and hydroxyls in this region. This results in better polishing, better removal of carbon

dioxide and silica, and higher product resistivity.


The product quality is optimized if the module is in balance and does not have extreme
internal current densities. The quality of the continuous regeneration of the polishing portion
of the bed is critical for achieving the highest resistivities.
Ionic balance and pH
Electrical neutrality must be maintained on an ionic level. It is not possible for more cations
to diffuse than anions.
Because of this, counter ions matter highly. If the ions in the feed are made up of a highly
mobile cation and a slow anion, then the EDI kinetics will adjust to the rate of the slowest
ions. In addition, the mobile protons (H+) and hydroxyls (OH-) will play a role in adjusting
the ionic balance. If there is a large mismatch in the ions in the feed, then there will be large
pH shifts between the product and concentrate streams. The quality will not be optimized.
pH therefore also influences quality highly. At lower pH, excess protons will diffuse as the
counter ions to the feed anions. The feed cations will not be as effectively removed.
At higher pH, there will not be protons to act as fast counter-cations. However, the carbon
dioxide will be more highly charged (carbonate) and therefore more mobile. The silica, too, is
more charged and mobile.
The recommended pH for optimal operation is 7.0 with minimal CO2 present.
Affecting the location of the "ionic front"
As said above, the location of the ionic front (the interface between the "working bed" and
the "polishing bed" sections of the EDI module) is important to the quality of the product.
For the highest resistivity water, and the lowest silica, the variables must be set to maximize
the depth of the polishing bed.

The ionic load must be minimized.


The product flow rate should be at the high end of the given range.
The voltage should be at the optimum (not too high or too low).
The concentrate outlet flow should be correct (e.g., 90% recovery) to effectively
remove ions from the surfaces of the membrane in the concentrate chambers. This
will also use more of the applied voltage across the critical dilute chambers.
The load of carbon dioxide should be minimized.
The pH should be at 7.0.

To save energy, if lower quality water is sufficient for the application, one can extend the
depth of the working bed and limit the depth of the polishing bed. This is achieved by:

Lowering the voltage.


Lowering the concentrate flow (higher recovery)this lowers the resistance of the
stack. This can also be achieved with concentrate recirculation or salt injection. Note:
the risk is hardness scaling in the concentrate.

Gas Production of the EDI Cell


Theory
At the anode, or positively charged electrode, electrons are transferred from ions in solution
to the external circuit (the anode accepts electrons - oxidation) by one or more of the
following reactions:
H2O + O2 + H+ + e- (E0 = 1.23 V) (1)
Cl- + Cl2 + e- (E0 = 1.37 V) (2)
At the cathode, or negatively charged electrode, electrons are transferred from the external
circuit to ions in the solution (the cathode donates electrons - reduction) by the following
reaction:
H2O+ e- + H2 + OH- (3)
These reactions are common to all electrodeionization devices, they are not unique to EDILX. What can differ somewhat from one device to another is the relative proportions of
reactions (1) and (2). This is affected by the concentration of chlorine in the anolyte, as
discussed below.
Estimating the amount of gas produced
Faradays Law states that the electric charge required to liberate one gram-equivalent of a
substance by electrolysis is 96,487 coulombs (a coulomb is the amount of electric charge that
crosses a surface in one second when a steady current of one ampere is flowing across the
surface). In both electrodialysis and electrodeionization, Faradays Law is used to relate the
transfer of salts through the membranes and the amount of current flowing through the
membranes. It can also be used to estimate the amount of gas produced at the electrodes.
Based on our EDI-LX test work we, assume only reactions 1 & 3 occur.
For reaction (1), 96,487 amp-sec will produce mole O2 gas & 1 mole H+ ion
For reaction (3), 96,487 amp-sec will produce mole H2 gas & 1 mole OHion
Using the ideal gas law and a gas constant of 0.08206 (l-atm)/(mole-K) the volume of a
mole of gas is calculated for STP (298K and 1 atm):

This allows us to determine the volume of gas produced per amp of current passed through

the electrodes.
For reaction 1: (1/4 mole O2/ 96,487 amp-sec) x 60 sec/min x 24 liters/mole x 1000 ml/liter =
3.73 ml O2/amp-min
For reaction 1: (1/2 mole H2/ 96,487 amp-sec) x 60 sec/min x 24 liters/mole x 1000 ml/liter =
7.46 ml H2/amp-min
Therefore, approximately 11 ml/min gas (3.7 ml/min O2 + 7.5 ml/min H2) are produced per
amp of current, at standard conditions of 25C and 1 atm.
Removal of gases
Some EDI systems are designed so that there is no separate electrode stream and the water
that flushes the surface of the electrodes mixes in with the rest of the concentrate as it exits
the module (the electrode compartments are connected hydraulically in parallel with the
concentrate compartments). The concentrate outlet stream therefore contains the products of
electrode reactions as well as the salts removed from the EDI feed water.
Chlorine
Normally the reaction at the anode will result in the formation of oxygen, but with sufficient
voltage and with chloride ion present, chlorine gas can be formed. In particular this is an
issue with electrodeionization systems which inject sodium chloride into the concentrate and
electrode streams in order to reduce module resistance and improve performance.
Some systems do not require brine injection because they use an ion-exchange resin filler in
the concentrate compartments in order to minimize the electrical resistance of the module.
This in itself limits the possibility of chlorine generation at the anode by reducing the
chloride concentration. But in addition, chlorine production is virtually eliminated by the use
of a resin filler in the electrode compartments. An additional advantage of the USFilter design
is that with elimination of salt injection and chlorine production, the EDI concentrate will
generally be better quality than the raw water and can often be recycled.
Hydrogen
People often express concern about the hydrogen gas produced at the cathode, since it is
known that under certain conditions hydrogen can be explosive. However, the amount of gas
that is produced by a EDI module is so small that it does not present a safety hazard when the
EDI system is installed in an area with normal ventilation. Codes require that buildings have
multiple air changes every day, one air change being a turnover of air equivalent to the
building's internal volume. The number of air changes varies depending upon the building's
use and the local codes, but a widely accepted value is half an air change per hour.
To show how little risk is presented by the electrode gases, let's perform an example
calculation assuming that a EDI module is installed in a 4m x 4m x 4m office which has the
normal ventilation of half an air change per hour. This is equivalent to an air flow of 32 m3/h.
Assume the EDI module is operating with a DC current of 10 amps, therefore producing
hydrogen at the rate of 74.6 ml/min (equivalent to 0.0045 m3/h). If all the hydrogen gas
leaves the water and enters the room air, then the resulting concentration of hydrogen in the

air would be about 0.014% (v/v), or about 141 ppm. This is well below the explosive limit of
a hydrogen/air mixture, which is 4.2% v/v at STP, and also well below the concentration at
which asphyxiation would occur. Even at the maximum current output of an EDI DC power
supply there is no safety risk as long as the ventilation meets typical building code
requirements.
Troubleshooting
Very Important Points to Remember!

EDI modules are subject to plugging and scaling like an RO system


EDI modules are subject to fouling and oxidation like a DI system
Data collection is critical - troubleshooting is nearly impossible without some
historical data
Most EDI operational problems are caused by improper pretreatment or out-of-limit
conditions of the feed water

Required for EDI troubleshooting

CO2 test kit


Conductivity meter
Volt-ohm-meter
Ohms Law V=IR
Faradays Law
Tape measure, torque wrench
Imagination
Open mind

Why an open mind?

"It can't be chlorine because the on-line chlorine analyzer always reads zero ppm Cl2".
Test kit shows 1 ppm free chlorine
"It can't be hardness scaling because the system has series softeners followed by RO".
Test kit shows raw water hardness 8 gpg, roughing softener effluent 6 gpg, polishing
softener effluent 1 gpg

Questions To Ask

Has there been a power interruption?


Have any modifications been made to system?
Have there been any pretreatment upsets (softener, bisulfite feed)?
Have any media filters rebedded?
Have there been noticeable changes in feed water?
Has there been a change in water source?

Has the system been out of service for extended period?


Has there been DC power on without water flow?

EDI Troubleshooting: Data Required

Present & historical data


DC volts, DC amps, temperature
Flow rates and inlet/outlet pressures of all streams
EDI feed water S/cm and CO2
In some cases:
o Hardness
o Silica
Reject conductivity
Free chlorine

Data Required: Why

Ohms Law: ohms = volts / amps


DC amps affects product quality
Increase in resistance can indicate scaling
Temperature affects resistance
Flow rates and Delta Ps can indicate increased internal restriction
Due to scale, biofilm, particulates, etc.
Drop in flow is same as increase in DP
Feed S/cm and CO2 affect product quality
Change in feed CO2 often undetected
Hardness and silica (feed & reject) are useful to assess potential for scale
Free chlorine is important if product flow has declined (possible resin oxidation)
Must use Cl2 test kit that is sensitive to 0.02 ppm

EDI Troubleshooting: Low Product Quality

Has product flow changed?


Check product (dilute) : concentrate DP
Has feed S/cm or CO2 changed?
Has DC current declined?
Has the power supply reached maximum voltage?

Mass Balance Calculation

Predict concentrations in reject stream Compare predicted and actual


Look for too much or too little in reject
Uses recovery (R) to determine concentration factor (CF)
CF = 100 / (100 - R))

Reject ppm = feed ppm x CF

More going in than coming out

EDI feed hardness = 0.8 ppm


EDI recovery = 90%
CF = 100 / (100 - 90)) = 10
Expected reject ppm = 0.8 x 10 = 8 ppm
Actual reject ppm = 0 ppm
Module is probably CaCO3 scaled!

EDI Troubleshooting: Resistance Increase

Check wiring connections for corrosion, tightness


Has feed temperature changed?
Check hardness and silica in feed and concentrate - mass balance calculation

EDI Troubleshooting: Low Flow

Check control valves


Is the RO making enough water?
Check EDI feed water for chlorine
Dont test just the RO feed only!
Check product (dilute) d/p in NaCl and again in NaOH
Large difference indicates resin oxidation

Flowcharts

Use Troubleshooting flowcharts included in instruction manuals


Cleaning is often used as a diagnostic tool
If all else fails, autopsy may be necessary

Oxidants

Free chlorine at EDI inlet should be < 0.02 ppm as Cl2


RO feed Cl2 may not = RO product Cl2
Oxidation of resin causes loss of flow and increase in d/p of resin-filled compartments
Other oxidants: Hydrogen Peroxide, Ozone
Improper cleaning or sanitization can cause oxidation
EDI module oxidation almost always associated with chemical dechlorination
(sodium sulfite or sodium bisulfite)
Resin oxidation in top 1 - 2 inches of cell
Rapid at high oxidant concentration (one day)
Gradual at low concentration (months)
Oxidation causes permanent damage - resin & membrane replacement required

Initial symptom of EDI module oxidation is decline in product water quality


Next symptom is increase in module electrical resistance
Final symptom is loss of product flow (increase in d/p caused by swelling of resin,
membranes due to de-crosslinking and therefore increased water content)

Scale

Causes drop in product water quality


Causes increase in electrical resistance
In severe cases causes drop in concentrate flow (increase in d/p)
Main culprits: CaCO3 and (rarely) SiO2
Analyze feed and concentrate for hardness or silica (mass balance calculation)

Figure 1
Heavy Scaling of Cathode
Organic Fouling

Causes drop in product water quality (often the only symptom)


Sometimes causes increase in electrical resistance
Does not affect pressure drop
Can occur post-RO

Figure 2
Result of Organic Fouling of IX Resins

Biofilm

Usually does not affect performance or stack resistance, only flow


Often associated with bisulfite feed
Use acid/caustic/percarbonate cleaning

Excessive Feed Pressure

Can cause increased external leaking


High product (dilute) >> concentrate can cause poor performance
High concentrate > product (dilute) can cause cross-leak
Design system to avoid pressure changes, water hammer

High Temperature

DC power on without water flow can cause burning of membrane and melting of
plastic piping

Figure 3
Burned Membrane Caused by DC Power w/No Flow
Particulates

Can cause plugging of concentrate & dilute


Normally doesnt affect product quality
Countercurrent cleaning sometimes gives better results

Cleaning &
Sanitizatio
n
Purpose of cleaning

Restore performance
o Quality
o Flow/pressure drop
Determine cause of performance drop
o Measure flow & /p during cleaning

Analyze cleaning solutions (especially acid)

When to Clean?

Reduced product quality (out of customer specification)


Reduced product flow
Reduced concentrate flow
50% increase in /p without change in feed temperature or flow
25% increase in electrical resistance (ohms) without change in feed temperature

Chemicals Used on EDI

Hydrochloric Acid
o Cleaning only
o Removes calcium scale, metal oxides
Sodium Chloride/Sodium Hydroxide
o Cleaning only
o Removes organic foulants from resin, membranes
Sodium Hydroxide
o Cleaning & sanitization
o Removes biofilm and silica
Sodium Percarbonate
o Cleaning & sanitization
o Removes organics, particles, biofilm and kills bacteria.
Peracetic Acid (Minncare)
o Sanitization only
o Kills bacteria

For Any Cleaning/Sanitization

RO or EDI product water should be used to make up solutions


Use nominal flow rates in all streams
Adjust valves as required
Turn off feed water supply to EDI
DC power supply off during cleaning
Make necessary connection to EDI
from cleaning tank
Clean EDI separate from RO

Typical Cleaning System


General Considerations

A brine flush is required prior to sodium percarbonate and peracetic acid


Remove ions that could otherwise catalyze oxidation of the resin or membranes.
After cleaning, flush with water for 10 min without DC, then apply normal current
If analyzing cleaning solutions, take samples before and after cleaning, label bottles

Multiple Cleanings

Do acid cleaning or brine flush before high pH cleaning such as caustic or percarbonate
It is not necessary to brine between high pH cleanings such as percarbonate and brine/caustic.
It is necessary to brine between low pH and high pH cleanings to avoid weakening the second solution

HCl Cleaning Procedure

2% HCl - cleaning
30-60 minutes recirculation
Overnight soak for severe scaling
Yellow color may indicate iron fouling
Add acid as required to maintain pH < 0.5

Brine flush to drain


Water flush (10-30 minutes) without power

Brine/Caustic Cleaning Procedure

Brine flush to drain to remove hardness


Recirculate 5% brine / 1% caustic solution for 30-60 minutes
Water flush for 10-30 minutes
Look for yellow-brown color or foaming
Multiple batches until discoloration of solution is minimal

Sanitization: Definitions

Sanitization or Disinfection
o Kills many (but not all) bacteria
o Destroys, neutralizes, and inhibits growth of microorganisms
Sterilization
o Free from bacteria 100%
o Absence of life (bacteria, viruses, spores, etc.)
o Requires steam or very high temperature

Caustic Cleaning/Sanitization Procedure

Brine flush to drain to remove hardness


Recirculate 2% caustic solution for 30-60 minutes
Water flush for 10-30 minutes
Increase temperature to 45C for removal of silica scale (Only if module compatible with elevated
temperature)

Sodium Percarbonate Cleaning/Sanitization Procedure

Flush once-through with 5% NaCl for 5 min


Water flush (to drain 1 min)
Recirculate 1% Percarbonate 30-60 min
Must mix hydrogen peroxide and sodium carbonate
Another salt flush to drain
Final water flush 10-30 min

Multi-Agent Cleaning Procedure>

Not for routine cleaning. Use only if heavily biofouled.


Recirculate 2%HCl
Flush once-through with 5% NaCl for 5 min
Water flush to drain
Recirculate 2% NaOH

Water flush (to drain 3 min)


Recirculate 1% percarbonate
Another salt flush to drain
Final water flush 10-30 min

Peracetic Acid Sanitization Procedure

5% salt flush once-through for 5 min


Water flush
Recirculate 400 ppm peracetic acid (dilute Minncare 100:1) for 30 minutes
Up to 1-1/2 Hour soak
Another salt flush
Final water flush

Hot Water Sanitization

Only if recommended my EDI manufacturer!


Maximum sanitization temperature 80C
Low inlet pressure (10 psig)
Follow manufacturer's directions

Hot Water Sanitization "watch-outs"

Hot water sanitization of upstream Carbon or RO could strip TOC and send to EDI
Could foul EDI resin
Elution of hardness from Carbon
Could scale EDI
Low rejection of RO during simultaneous RO/EDI sanitization (low pressure)
Use DI water for HWS when possible
Generally should flush & rinse carbon and RO before sanitizing EDI

Which Cleaning for High /p?

Perform brine flush - if this temporarily reduces pressure drop (shrinks resin) but problem returns with
operation, module may be irreversibly damaged by chlorine (Chlorine causes de-crosslinking of ion
exchange resin and subsequent increase in water content of resin and increase in swelling. It is the
swelling of the resin in a confined space that causes the increase in pressure drop.)
Percarbonate clean for particle fouling or biofouling

Which Cleaning for High Module Resistance?

Usually caused by hardness scale


Can also be caused by oxidation
Look for hardness mass balance; acid clean

In rare cases silica scaling, 2% caustic @ 45C (ONLY if recommended by manufacturer!)

Action for Quality Decline with no Other Symptoms?

Check for module cross leaks


Clean with brine/caustic for organic fouling

Shutdown/Preservation

Drain standing water


Seal inlet & outlet connections
Suitable for one year
O-ring prevents evaporation out edge of membrane
Propylene glycol incompatible with modules at concentrations above 75%
Typical
Systems

Typical Two-Pass RO / EDI System

EDI
Designs

P&ID of EDI System requiring


concentrate recycle and brine injection

P&ID of EDI System not requiring


concentrate recycle or brine injection
(Ionpure Technologies)

P&ID of EDI System not requiring

concentrate recycle or brine injection


(Electropure)
Typical
Systems

Typical Two-Pass RO / EDI System

Glossary
ABCDEFGHIJKLMNOPQRSTUVWXYZ
A
Absorption - The holding of a substance within a solid by cohesive or capillary forces.
Accumulator - A pulsation dampener installed on the suction and/or discharge lines of
pumps, generally plunger type, to minimize pressure surges and provide uniformity of flow.
Accuracy - The closeness of agreement between an observed value and an accepted reference
value. Where an accepted reference value is not available, a measure of the degree of
conformity of a value generated by a specific procedure to the assumed or accepted true
value, and includes both precision and bias.
Acetylation - Substitution of an acetyl radical for an active hydrogen. Specifically, formation
of cellulose acetate from cellulose.
Acidity - The quantitative capacity of aqueous media to react with hydroxyl ions.
Activated Carbon - Granulated or powdered activated carbon used to remove tastes, odor,
chlorine, chloramines, and some organics from water. A family of carbonaceous substances
manufactured by processes that develop adsorptive properties.
Adsorption - The holding of a substance onto the surface of a solid by chemical surface
forces, without forming new chemical bonds.
Aerobic Bacteria - Bacteria that require oxygen for growth, see Bacteria, aerobes.
Aggregate - Granular material such as sand, gravel, crushed stone.
Air Scour - Distributing air over the entire filter area at the bottom of a filter media flowing
upward to improve the effectiveness of backwashing and/or to permit the use of lower
backwash water flow rate.
Algae - Any of a group of chiefly aquatic mono cellular plants with chlorophyll often masked

by a brown or red pigment.


Alkalinity - The quantitative capacity of aqueous media to react with hydrogen ions. "M"
alkalinity is that which will react with acid as the pH of the sample is reduced to the methyl
orange endpoint about 4.5. "P" alkalinity is that which reacts with acid as the pH of the
sample is reduced to the phenolphthalein end point of 8.3. "M" is the total alkalinity which is
the sum of hydroxide plus carbonate plus bicarbonate contents; "P" includes all the hydroxyl
and half the carbonate content.
Alum - Aluminum sulfate, Al2(SO4) 3XH2O (X=14-18), a coagulant.
Ambient Temperature - The temperature of the surroundings, typically 20o-25oC.
Amorphous - Non crystalline, devoid of regular cohesive structure.
Anaerobic Bacteria - Bacteria that do not use oxygen, oxygen is toxic to them, see Bacteria
(anaerobes).
Amphoteric - Capable of acting as an acid or a base.
Angstrom ( ) - A unit of length equaling 10-10 meters, 10-8 centimeters and 3.937 x 10-9
inches. The symbol is or A.U..
Anion - An ion (electrically charged atom of group of atoms) that carries one or more
negative charges, e.g., Cl-, OH-, SO4 =
Anion Exchange Material - A material capable of the reversible exchange of negatively
charged ions.
Anisotropic Membrane - A nonuniform structure in cross-section; typically the support
substructure has pores much larger than the barrier layer. (See Asymmetric Membranes).
Anode - A positively (+) charged electrode that attracts anions
Anionic Polyelectrolyte - Usually acrylamide and acrylamide and acrylic copolymers,
negatively charged, used for coagulation/flocculation, see Polyelectrolytes.
Anthracite - A granular hard coal used as a filtration media, commonly used as the coarser
layer in dual and multimedia filters.
Antifoulant - See Antiscalant.
Antiscalant - A compound added to a water which inhibits the precipitation of sparingly
soluble inorganic salts.
Anti-Telescoping Device - A plastic or metal device attached to the ends of a spiral wound
cartridge to prevent movement of the cartridge leaves in the feed flow direction, due to high
feed flows.
AOC - Assimilable Organic Carbon.
Applied Voltage - The DC voltage across the anode and cathode of each module. The voltage
required depends primarily on the number of cells in the module.
Aquifer - A water-bearing geological formation that provides a ground water reservoir.
Aramid - A fully aromatic polyamide.

Array - An arrangement of devices connected to common feed, product and reject headers
i.e. a 2:1 array.
Asymmetric Membrane - Membrane which has a change in pore structure with depth. (See
Anisotropic Membranes).
ATD - See Anti Telescoping Device.
Atomic Weight - The relative mass of an atom based on a scale in which a specific carbon
atom (carbon 12) is assigned a mass value of 12.
ATP - Adenosine triphosphate.
Autopsy - The dissection of a membrane module or element to investigate causes of
unsatisfactory performance.
Availability - The on-stream time or rated operating capacity of a water treatment system.
A-Value - Membrane water permeability coefficient. The coefficient is defined as the amount
of water produced per unit area of membrane when net driving pressure (NDP) is unity, a unit
of measurement is m3/hr/m2/kPa.
AWWA - American Water Works Association.
AWWARF - American Water Works Association Research Foundation.
(Top)
B
Backwash - Reverse the flow of water w/wo air either across or through a medium or
membrane designed to remove the collected foreign material from the bed or membranes.
Bacteria - Any of a class of microscopic single-celled organisms reproducing by fission or by
spores. Characterized by round, rod-like spiral or filamentous bodies, often aggregated into
colonies or mobile by means of flagella. Widely dispersed in soil, water, organic matter, and
the bodies of plants and animals. Either autotrophic (self-sustaining, self-generative),
saprophytic (derives nutrition from non-living organic material already present in the
environment), or parasitic (deriving nutrition from another living organism). Often symbiotic
(advantageous) in man, but sometimes pathogenic.
Bactericide - Agent capable of killing bacteria.
Bacteriostat - Substance that prevents bacterial growth and metabolism but does not
necessarily kill them.
Bank - A grouping of devices. (See Array, Block, Train).
Bar - A unit of pressure. One bar is equal to 14.5 psi, 1.02 kg/cm2, or one Pascal. By
definition one bar is equal to one million dynes per square centimeter.
BAT - Best Available Technology.
Baume Scale - oBe - A measure of the density of a solution relative to water.
Bed Depth - The depth of the filter medium or ion exchange resin in a vessel.

Bed Expansion - The depth increase of filter medium or ion exchange resin that occurs
during backwashing.
Binders - In reference to cartridge filters, chemicals used to hold, or 'bind', short fibers
together in a filter.
Biocide - A substance that kills all living organisms.
Biological Deposits - The debris left by organisms as a result of their life processes.
Biomass - Any material which is or was a living organism or excreted from a microorganism.
Biostat- A substance that inhibits biological growth.
Blinding - In surface filtration, a build-up of particulates on the filter, restricting fluid flow
through the filter at normal pressures.
Block - A grouping of devices in a single unit having common control. (See array, Bank
Train).
BOD - Biochemical Oxygen Demand - The amount of dissolved oxygen utilized by natural
agencies in water in stabilizing organic matter at specified test conditions.
Body Feed - The continuous addition of filter medium (e.g.; diatomaceous earth) to sustain
the efficacy of the filter.
BOO - Build, Own Operate.
BOOT - Build, Own, Operate and Transfer.
Boundary Layer - A thin layer at the membrane surface where water velocities deviate
significantly less than those in the bulk flow.
Brackish Water - Water with an approximate concentration of total dissolved solids ranging
from 1000 to 10,000 mg/L. (See High Brackish Water and Sea Water).
Breakpoint Chlorination - The point at which the water chlorine demand is satisfied and
any further chlorine is the chlorine residual, the "free" chlorine species.
Break Tank - A storage device used for hydraulic isolation and surge protection.
Brine - The concentrate (reject) stream from a crossflow membrane device performing
desalination. Portion of the feed stream which does not pass through the membrane.
Brine (Concentrate) Seal - A rubber lip seal on the outside of a spiral wound cartridge which
prevents feed by-pass between the cartridge and the inside pressure vessel wall.
Brine Seal Carrier - See ATD.
Brine System Staging - A process in which the concentrate, under pressure, of a group of
membrane devices is fed directly to another set of membrane devices to improve the
efficiency of the water separation.
Bubble Point Pressure - The pressure necessary to displace a liquid held by surface tension
forces from the largest equivalent capillaries in a membrane filter.
Bubble Point Test - A non destructive membrane filter test used to assess filter integrity and

proper installation.
Bundle - A general term for a collection of parallel filaments or fibers.
B-Value - Salt diffusion coefficient. The coefficient is defined as the amount of salt
transferred per unit area of membrane when the difference in salt concentration across the
membrane is unity. A unit of measurement is m/hr.
BWRO - Brackish Water Reverse Osmosis.
(Top)
C
CAC - Combined Available Chlorine.
Calcium Carbonate Equivalents (mg/L as CaCO3) - A method for expressing mg/L as ion
in terms of calcium carbonate. Concentration in calcium carbonate equivalents is calculated
by multiplying concentration in mg/L of the ion by the equivalent weight of calcium
carbonate (50) and dividing by the equivalent weight of the ion.
Carbonate Hardness - The hardness in a water caused by carbonates and bicarbonates of
calcium and magnesium. The amount of hardness equivalent to the alkalinity formed and
deposited when water is boiled. In boilers, carbonate hardness is readily removed by
blowdown.
Calcium Hypochlorite - Ca (HClO)2, a disinfection agent.
Cartridge - See Spiral-Wound Cartridge.
Catalyst - A substance whose presence initiates or changes the rate of a chemical reaction,
but does not itself enter into the reaction.
Cathode - A negatively (-) charged electrode that attracts cations.
Cation - An ion that carries one ore more positive charges, e.g., Na+ , NH4 + , Ca+2.
Cation Exchange Material - A material capable of the reversible exchange of positively
charged ions.
Cationic Polyelectrolyte - A polymer containing positively charged groups used for
coagulation/flocculation, usually dimethyl - aminoethyl methacrylate or dimethyl-aminoethyl
acrylate. (See Polyelectrolyte.)
CDI - Continuous Deionization. Commercial electrodeionization devices developed and sold
by U.S. Filter.
EDI - Continuous Electrodeionization. Electrodeionization where the ion transport properties
of the active media are the primary sizing parameters
Cellulose - An amorphous carbohydrate (C6H10O5) that is the principal constituent of wood
and plants.
Cellulose Acetate (CA) - In the broad sense, any of several esters of cellulose and acetic
acid.

Celsius (oC) - The designation of the degree on the International Practical Temperature Scale.
Formerly called centigrade, oC = K minus 273.15 K= Kelvin.
Centigrade - Since 1948, now called Celsius, a temperature scale.
Ceramic Membrane - Generally a glass, silica, alumina, or carbon based membrane.
Generally used in micro and ultrafiltration. They tend to withstand high temperatures and
wide pH ranges and be more chemically inert than polymeric membranes.
CFU - Colony Forming Unit; unit used in the measure of Total Bacteria Count (TBC).
Channeling - Unequal flow distribution in the desalination bundle or filter bed.
Chelating Agents - A sequestering or complexing agent that, in aqueous solution, renders a
metallic ion inactive through the formation of an inner ring structure with the ion.
Chemical Feed Pump - A pump used to meter chemicals, such as chlorine or polyphosphate,
into a feed water supply.
Chloramine - A combination of chlorine and ammonia in water which has bactericidal
qualities for a longer time than does free chlorine.
Chlorine - Chemical used for its qualities as a bleaching or oxidizing agent and disinfectant
in water purification.
Chlorine Demand - The amount of chlorine used up by reacting with oxidizable substances
in water before chlorine residual can be measured.
Chlorine Residual - The amount of available chlorine present in water at any specified time.
Chlorine, Free Available - The chlorine (Cl2) hypochlorite ions (OCl-), hypochlorous acid
(HOCl) or the combination thereof present in water.
Chlorine, Total Available - The sum of Free Available Chlorine plus chloramines present in
water.
CIP - Cleaning-In-Place.
Citric Acid - C3H4(OH)(COOH)3, membrane cleaning chemical.
Clarifier - A tank in which precipitate settles and supernatant overflows, a liquid - solids
separation unit using gravity to remove solids by sedimentation.
Clark Degree - Number of grains of substance per one British imperial gallon of water
expressed as CaCO3. Concentration in Clark or English degree is calculated by dividing
concentration in calcium carbonate equivalents by 14.3. One grain weighs 1/7,000 pound and
one imperial gallon of water weighs 10 pounds at 25oC.
Coagulant - Chemical added in water and wastewater applications to cause destabilization of
suspended particles and subsequent formation of flocs that adsorb, entrap, or otherwise bring
together suspended matter that is so fine, it is defined as colloidal. Compounds of iron and
aluminum are generally used to form flocs to allow removal of turbidity, bacteria, color, and
other finely divided matter from water and waste water.
Coalescing - The separation of mixtures of immiscible fluids (such as oil and water) based on
different specific gravities and surface tensions. Can occur whenever two or more droplets

collide and remain in contact and then become larger by passing through a coalescer. The
enlarged drops then separate out of solution more rapidly.
COD - Chemical Oxygen Demand - The amount of oxygen required under specified test
conditions for the oxidation of water borne organic and inorganic matter.
Colloid - A substance of very fine particle size, typically between 0.1 and 0.001 microns in
diameter suspended in liquid or dispersed in gas. A system of at least two phases, including a
continuous liquid plus solid, liquid or gaseous particles so small that they remain in
dispersion for a practicable time.
Colony Forming Unit (CFU) - Unit used in the measure of Total Bacterial Count (TBC).
Compaction - In crossflow filtration, the result of applied pressure and temperature
compressing a polymeric membrane which may result in a decline in flux.
Compartment - A unit in the Ionpure module which includes a dilute spacer and a
concentrate spacer.
Composite Membrane - A membrane having 2 or more layers with different physical or
chemical properties. Membrane manufactured by forming a thin desalinating barrier layer on
a porous carrier membrane.
Concentrate - The stream exiting a crossflow membrane device which has increased
concentration of solutes and particles over the feed stream. Portion of the feed stream which
does not pass through the membrane. The stream in which dissolved solids and/or particulates
are concentrated in a membrane separation process.
Concentrate Stream - The flow of water through the parallel concentrating compartments,
where ions are collected.
Concentrate Spacer - Portion of the module compartment where salts are concentrated.
Concentrate Recycle - A technique for improving recovery in which a fraction of the
concentrate is recycled through the membrane system.
Concentration Factor - CF = ((Brine Concentration CB)/(Feedwater Concentration CF)) (1/(1 - Conversion))
Concentration Polarization - The increase of the solute concentration over the bulk feed
solution which occurs in a thin boundary layer at the feed side of the membrane surface,
resulting from the removal of the solvent.
Conductivity - The electrical measurement of water's ability to conduct an electrical current,
which is dependent on the concentration of ions in the water and its temperature. Units are
microsiemens/cm or mS/cm or micromhos/cm and are normalized to 25 deg C.
Contaminant - Any foreign substance present which will adversely affect performance or
quality.
Continuous Deionization - A deionization process that does not require regular interruptions
in service to discharge ionic materials collected from the water being processed.
Control Block - A group of devices having a common piping and control system.

Conversion (Y) - Product water flow rate divided by feed water flow rate. Also called
recovery; given as fraction or decimal. (See Recovery).
Corrosion Products - Products that result from chemical or electrochemical reaction
between a metal and its environment.
CPU - Chloroplatinate Unit (color indicator).
Crossflow Membrane Filtration - A separation of the components of a fluid by
semipermeable membranes through the application of pressure and flow parallel to the
membrane surface. Includes the processes of reverse osmosis, ultrafiltration, nanofiltration,
and microfiltration.
Current Draw - The DC current through each module. The current depends on the ion load
in the feed, on the module recovery rate, and on the amount of water splitting. Roughly
independent of the number of cells.
Current Efficiency - Theoretical current required to transport the feed ions divided by the
actual current. Expressed as %.
(Top)
D
Dalton - An arbitrary unit of molecular weight, 1/12 the mass of the nuclide of carbon 12.
Unit of measure for the smallest size of the molecule retained by an ultrafilter.
DBP - Disinfection By-Products (a rule as part of the SDWA).
DC Current - Current that does not change direction. Proportional in an EDI system to the
number of ions moved, including split water ions.
DC Potential - Voltage that does not change polarity. Electrodeionization is only possible
with this form of driving force.
Dead-End Filtration - A process in which water is forced through a medium which captures
the retained particles on and within it, where the process involves one influent and one
effluent stream.
Deaerator - A device to remove air from water.
Decarbonator - A device to remove carbon dioxide from water.
Degasification - The process of removing dissolved gasses from water.
Deionization (DI) - The removal of ions from a solution by ion exchange.
Demineralization - The process of removing minerals from water.
Depth Filter - A filter of spool-wound polypropylene; compressed, matted fibers; compacted
sand or diatomaceous earth; or sintered metal or glass that retains particles by random
adsorption or entrapment. The efficiency of depth filters is rated in nominal terms. Depth
filters were the main type of filters before the introduction of membrane filtration.
Desalination - See Demineralization.

Detergent - A cleansing agent; any of numerous synthetic water soluble or liquid-organic


preparations that are chemically different from soaps but resemble them in the ability to
emulsify oils and hold dirt in suspension.
Dialysis - A separation process dependent on different diffusion rates of solutes across a
permeable membrane without an applied hydraulic driving force.
Diatom - Single cell marine animal having a coating consisting principally of silica.
Diatomaceous Earth (DE) Filtration - Filtration using an amphorous, lightweight siliceous
earth medium occurring naturally as the fossil remains of diatoms.
Dilute Steam - Water that flows through the dilute spacers and is deionized.
Dilute Spacer - Plastic sheet fused to one cation and one anion membrane, and containing
ion exchange resin. Desalination occurs within the dilute spacer.
Disinfection - The process of killing organisms in a water supply or distribution system by
means of heat, chemicals, or UV light.
Dissolved Solids - The residual material remaining after filtering the suspended material
from a solution and evaporating the solution to a dry state at a specified temperature. That
matter, exclusive of gases, which is dissolved in water to give a single homogeneous liquid
phase.
Distillation - The process of condensing steam from boiling water on a cool surface.
Donnan equilibrium - If a non-permeable substance is separated from permeable substances
by a membrane, ions will pass through in different amounts and establish an electrostatic
difference (membrane potential); also known as the Gibbs-Donnan equilibrium.
DOC - Dissolved Organic Carbon.
Double Pass RO System - RO system in which the permeate from the first RO plant is
further desalinated by a subsequent RO system.
DOX - Dissolved Organic Halogen.
Draw-back - The reverse flow of product water from the product side across the membrane
to the feed/brine side as a result of osmosis (product to feed flow); occurs on shut down of
RO process.
Dual Media Filter - Filter with a bed consisting of two separate granular filter media, the
coarse, low density, at the top, the fine, high density, at the bottom.
Dynamic Membrane - A surface layer coating deposited on a semipermeable membrane
which aids as a separation barrier.
(Top)
E
EDTA - Ethylenediaminetetraacetic acid, a common chelating agent for removal of di and tri
valent ions.

Effluent - A stream exiting a treatment system.


Electrical Conductivity - The property of a fluid or solid that permits the passage of an
electrical current as a result of an impressed emf, measured in micro Siemens/cm or micro
mhos. The reciprocal of the resistance in ohms measured between opposite faces of a
centimeter cube of an aqueous solution at a specified temperature.
Electrode - A metal plate (anode or cathode) that conducts an electrical field and catalyzes
electrochemical reactions. Electrodes are equipped with an external post for attachment to a
power supply.
Electrode Stream - Water stream that flows by the anode and cathode, and accumulates
chemicals and gases electrochemically generated by the electrodes.
Electrodeionization - A process that removes ionizable species from liquids using
electrically active media and using an electrical potential to influence ion transport.
Electrodialysis (ED) - A process in which ions are transferred through membranes from a
less concentrated to a more concentrated solution using direct current electric power as the
driving force.
Electrodialysis Reversal (EDR) - Same as ED with the addition of a polarity reversal step
added to improve performance.
Element - The component containing the membrane, generally replaceable, such as a spiral
wound cartridge.
English Degree - Number of grains of substance per one British imperial gallon of water.
EPA - Environmental Protection Agency (USA) - an organization that has set the potable
water standards.
Equivalent per Million (EPM) - A unit chemical equivalent weight of solute per million unit
weights of solution. Concentration in equivalents per million is calculated by dividing
concentration in ppm by the equivalent weight of the substance or ion. Equivalent weight is
the atomic weight of the substance divided by the valence of the substance.
Equivalent Weight - The weight of an ion determined by dividing the sum of the atomic
weights of its component atoms by its valence.
ERD - Energy Recovery Device.
ERT - Energy Recovery Turbine.
ESWTR - Enhanced Surface Water Treatment Rule.
Evaporation - Process where a liquid (water) passes from a liquid to a gaseous state.
(Top)
F
FAC - Free Available Chlorine.
Fahrenheit (oF) - Designation of a degree on the Fahrenheit temperature scale that is related

to the International Practical Temperature Scale.


FDA - Food and Drug Administration (USA).
Feed Water - The water that is plumbed into the EDI module. It supplies the product
compartments and concentrating compartments. The source of this water is Reverse Osmosis
(RO) product or permeate.
Feed Channel Spacer - A plastic netting between membrane leaves which provides the flow
channel for the fluid passing over the surface of the membrane and increases the turbulence
of the feed-brine stream.
Feed Distributor - The plastic mesh cylinder at the core of the fiber bundle which distributes
the feed evenly.
Feed Water - That water entering a device or process.
Ferric Chloride - Crystalline form of FeCl3 . 6H2O, a coagulant.
Ferric Sulfate - Fe2(SO4)3.9H2O, a coagulant.
Ferrous Sulfate - FeSO4.7H2O, a coagulant.
Fiber Bundle - The heart of the permeator consisting of the hollow fiber polymer membrane,
epoxy tube sheet, nub and feed distributor.
Filter Cake - The accumulated particles on a filter surface, usually from a slurry mixture.
Filtrate - The portion of the feed stream which has passed through a filter.
Fixed Matter - Residues from the ignition of particulate or dissolved matter, or both.
Flat Sheet Membrane - A sheet type membrane may be coated onto a fabric substrate.
Floc - A loose, open-structured mass produced by the aggregation of minute particles.
Flocculant - Chemical(s) which, when added to water, form bridges between suspended
particles causing them to agglomerate into larger groupings (flocs) which then settle or float
by specific gravity differences.
Flocculation - The process of agglomerating fine particles into larger groupings called flocs.
Flow Balancing - The use of an imposed pressure drop (flow balancing tube), to minimize
conversion differences of modules operating in parallel.
Flow Balancing Tube - See Flow Balancing.
Flux - The membrane throughput, usually expressed in volume of permeate per unit time per
unit area, such as gallons per day per ft2 or liters per hour per m2.
Fouling - The reduction of flux due to a build-up of solids on the surface or within the pores
of the membrane, resulting in changed element performance.
Fouling Index (FI) - See SDI.
FRC - Free Residual Chlorine.
Freeboard - The space above a filter bed in a filtration vessel to allow for expansion of the
bed during back washing.
Free (available) Chlorine - The chlorine existing as hypochlorous acid (HOCl) or its

dissociated ions (H+ OCl-). The less dissociated the hypochlorous acid, the more effective the
sanitization. The degree of dissociation can be controlled by adjusting the pH. At a pH of 6,
96% of the chlorine will be in the HOCl form. At a pH of 9, only 3% will be the HOCl form,
97% of the chlorine will be in the H+OCl- form.
French Degree - Calcium carbonate equivalents expressed in parts per hundred thousand.
Concentration in French degree is calculated by dividing concentration in calcium carbonate
equivalents by 10.
FRP - Fiberglass Reinforced Plastic.
Fungus - Primitive plants distinguished from algae by the absence of chlorophyll.
(Top)
G
GAC - Granular Activated Carbon.
Galvanic Corrosion - Accelerated corrosion of a metal because of an electrical contact with
a more noble metal or non metallic conductor in a electrolyte.
GD - Gallons per day. (See GPD).
GPM (gpm) - Gallons per minute, a measurement of the flow of water.
GFD (GPDSF) - Unit of permeate rate or flux; gallons per day per square foot of effective
membrane area.
GPD - Unit of flow rate; gallons per day. (See GD).
German Degree - Calcium oxide equivalents expressed in parts per hundred thousand.
Concentration in German degree is calculated by dividing concentration in calcium carbonate
equivalents by 17.86.
Grain - Unit of weight, 0.648 grams, 0.000143 pounds. there are 7,000 grains in 1 pound.
Grains Per U.S. Gallon (GPG) - Number of grains of substance per one U.S. gallon of
water. Concentration in GPG is calculated by dividing concentration in ppm of the ion by
17.1. One grain weighs 1/7,000 pounds and one U.S. gallon weights 8.3 pounds.
GRAS - Materials "generally regarded as safe", as listed by the FDA.
Gravity Filter - A filter through which water flows through it by gravity.
Greensand - A mineral (glauconite), used as a filtration medium. (See Manganese
Greensand).
Groundwater - Water confined in permeable sand layers between rock or clay; that part of
the subsurface water that is in the saturated zone.
(Top)
H

HAA - A group of 6 Halo acetic acids regulated in drinking water (mono, di and trichloroacetic acid, mono and di bromoacetic acid and chlorobromoacetic acid).
Halogen - Any element of the family of the elements fluorine, chlorine, bromine and iodine
(definition for purpose of this standard).
Hardness - The solution in water of both calcium and magnesium as cations independent of
the nature of anions present. Hardness is generally expressed in terms of calcium carbonate
(CaCO3). "Temporary hardness" (carbonate hardness") is equal to or less than the alkalinity.
"Permanent hardness" ("non-carbonate hardness") is an excess of hardness over alkalinity.
Header - See Manifold.
Head Loss - The reduction in liquid pressure usually associated with the passage of a
solution through a filter media bed.
Heavy Metals - Elements having a high density or specific gravity of approximately 5.0 or
higher. A generic term used to describe contaminants such as cadmium, lead, mercury, etc.
Most are toxic to humans in low concentration.
High Brackish Water - Water with an approximate concentration of total dissolved solids
ranging from 10,000 to 30,000 mg/L. (See Brackish Water and Sea Water).
High-purity Water - Highly treated water with attention to microbiological, particle,
organics and mineral reduction or elimination.
Hollow Fiber (HF) Membrane - Self-supporting membrane fibers which have a hollow bore
like a cylinder. In reverse osmosis, the membrane is usually on the outside with bore
conveying the permeate. In ultra and micro filtrations the membrane is generally on the
inside.
HPC - Heterotrophic Plate Count. Use to be called SPC.
Humic Acid - A variety water-soluble organic compounds, formed by the decayed vegetable
matter, which is leached into a water source by runoff or percolation. Present in most surface
and some ground waters. Higher concentrations cause a brownish tint. Difficult to remove
except by adsorption, ultrafiltration, nanofiltration or reverse osmosis.
Humidity, Absolute - The mass of water vapor per unit volume of the atmosphere usually
measured as grams per m3.
Humidity, Relative - The ratio of the actual pressure of existing water vapor to the maximum
possible (saturation) pressure of water vapor in the atmosphere at the same temperature,
expressed as a percentage.
Hydrated Lime - Dry calcium hydroxide.
Hydrophilic - Having an affinity for water.
Hydrophobic - Lacking an affinity to water.
Hydroxyl Alkalinity - See Alkalinity.
Hyperfiltration - Separation of dissolved ions from a feed stream as in nanofiltration and
reverse osmosis.

(Top)
I
Imperial Gallon (IG) - 1.2 times U.S. gallon
In-Line Coagulation - A filtration process performed by continually adding a coagulant to
the raw feedwater and then passing the water through a filter(s) to remove the microfloc
which has been formed.
Interconnector - A device to connect adjacent membrane elements in series and to seal the
product channel from the feed-brine channel.
Ion - A charged portion of matter of atomic or molecular dimensions.
Ion-Exchange - A reversible process by which ions are interchanged between a solid and a
liquid with no substantial structural changes in the solid; ions removed from a liquid by
chemical bonding to the media.
Ion-Exchange Capacity (volume basis) - The number of milliequivalents of exchangeable
ions per milliliter of backwashed and settled bed of ion-exchange material in its standard
form.
Ion-Exchange Capacity (weight basis) - The number of milliequivalents of exchangeable
ions per dry gram of ion-exchange material in its standard form.
Ion-Exchange Material - A water insoluble material that has the ability to reversibly
exchange ions in its structure, or attached to its surface as functional groups, with ions in a
surrounding medium.
Ion Exchange Membrane - A membrane containing ion-exchange groups that are selectively
permeable to either anions or cations but is impermeable to water.
Ion-Exchange Particle - An ion-exchange material in the form of spheroids or granules.
Ion Exchange Resin - Resin beads containing ion-exchange groups that selectively adsorb
either anions or cations.
Ionic Strength - Measure of the overall electrolytic potential of a solution, the strength of a
solution based on both the concentrations and valencies of the ions present.
Ionization - The disassociation of molecules into charged particles (ions).
(Top)
J-K
Jackson Turbidity Unit, JTU - Unit of measure used with the Jackson Candle Turbidimeter.
Jar Test - A laboratory procedure for the evaluation of a treatment to reduce dissolved,
suspended colloidal and non settleable matter from water (ASTM D2035).
(Top)

L
Langelier Saturation Index, LSI - An index calculated from total dissolved solids, calcium
concentration, total alkalinity, pH, and solution temperature that shows the tendency of a
water solution to precipitate or dissolve calcium carbonate.
LD-50 - Concentration required for 50 percent mortality (lethal dose).
Leaf - The sandwich layer of flat-sheet membrane/product channel spacer/flat-sheet
membrane, glued together on the sides and across the outer end in a spiral wound element.
Lime - Ca(OH)2, calcium hydroxide, a common water treatment chemical.
Lime Soda Softening - Use of lime and Na2CO3 for softening water.
Limestone - Either calcite limestone (CaCO3) or dolomitic limestone (CaCO3 -MgCO3).
Loose RO - See Nanofiltration.
LSI - Langelier Saturation Index, A measure of the tendency of water to dissolve or deposit
calcium carbonate CaCO3 (scaling). The LSI is calculated from the total alkalinity, hardness,
total dissolved solids (TDS), pH and temperature of the water. (See also S&DSI).
(Top)
M
Manganese Greensand - A manganese dioxide coated greensand used as a filter medium for
removal of manganese and iron. (See Greensand).
Manifold - An enlarged pipe with connections available to the individual feed, brine and
product ports of a desalination device.
Mass Transfer Coefficient (MTC) - Mass (or volume) transfer through a membrane based
on driving force.
MCL - Maximum Contaminant Level.
Megohm - The unit of measurement typically used to quantify the ionic purity of product
water produced by a deionization system. It is a measurement of electrical resistance.
Ultrapure water with no impurities may reach 18.24 MW or Megohm.cm at 25 deg C. (See
Microsiemens, Ohm).
Membrane - Engineered thin semipermeable film which serves as a barrier permitting the
passage of materials only up to a certain size, shape, or electro-chemical character.
Membranes are used as the separation agent in reverse osmosis, electrodialysis, ultrafiltration,
nanofiltration, and microfiltration, as disc filters in laboratories, and as pleated filter
cartridges, particularly for microfiltration.
Membrane Area - The active area available for micro, nano and ultra- filtration and reverse
osmosis.
Membrane Compaction - See Compaction.

Membrane Configuration - The design and shape of a given membrane element (cartridge)
such as tubular, spiral wound or hollow fiber.
Membrane Element - A bundle of spiral membrane envelopes or hollow fiber membranes
bound together as a discrete entity.
Membrane Filter - Geometrically regular porous matrix; removes particles above pore size
rating by physical size exclusion.
Membrane Salt Passage - SPm is the concentration of a compound in the permeate related
to its average concentration on the feed/concentrate side.
Membrane Fouling - A buildup of adsorbed material on the surface of the membrane that
causes an increase in membrane electrical resistance.
Membrane Softening - Use of crossflow membrane to substantially reduce hardness ions in
water, see Nanofiltration.
MGD (MGPD) - Millions of gallons per day.
Mho - A measure of conductance. The ratio of the current flowing through a conductor,
measured in amperes, to the potential difference between the end of the conductor, measured
in volts, is the conductance in mhos. A mho is a unit of conductance equal to the reciprocal of
the ohm, expressed as amperes/volt.
Microaerophilic Bacteria - Aerobic bacteria that require 2-10% oxygen in order to grow.
See Bacteria (microaerophiles).
Microfiltration (MF) - Filtration designed to remove particles in the approximate range of
0.05 to 2 micrometers.
Microbe - Bacteria and other organisms that require the aid of a microscope to be seen.
Micromho - One millionth of a Mho.
Micron (micrometer) - A metric unit of measurement equivalent to 10-6 meters, 10-4
centimeters. Symbol is mm.
Microorganism - See Microbe.
Microsiemens - Unit of measurement of water purity by electrical conductivity; one
micromho; reciprocal of resistance. (See Megohm, Ohm).
Milliequivalent Per Liter (meq/L) - A weight-volume measurement obtained by dividing
the concentration expressed in milligrams per liter by the equivalent weight of the substance
or ion. If specific gravity is unity meq/L is the same as epm.
Milligram Per Liter (mg/L) - A weight-volume measurement which expresses the
concentration of a solute in milligrams per liter of solution. When specific gravity is unity
mg/L = ppm. When specific gravity is not unity, mg/l divided by specific gravity of solution
equals ppm.
Mixed-bed - A physical mixture of anion-exchange and cation - exchange materials.
Module (RO) - A membrane element combined with the element's housing. Pressure vessel

containing membrane element(s).


Module (EDI) - The basic desalting unit of the Ionpure EDI system.
Module Resistance - The electrical resistance across the module. Applied voltage divided by
current draw, typically expressed in ohms.
Molality (-mi) - Moles (gram molecular weight) of solute per 1000g of solvent.
Molarity (mi) - Moles (gram molecular weight) of solute per liter of total solution.
Molecular Weight Cut Off (MWCO) - The rating of a membrane for the size of uncharged
solutes it will reject. Also referred to as Nominal Molecular Weight Cut Off (NMWCO).
Monovalent Ion - An anion or cation with one valence (e.g. K+m Cl+, OH-)
Multimedia Filter - Filter with a bed consisting of three or more separate filter media. The
coarsest, lowest density at the top and the finest, highest density at the bottom.
m-Value - The negative slope of a curve plotting log flow vs. log time. A measurement of the
degrees of membrane compaction as a result of temperature, pressure and time.
(Top)
N
NaHMP - Sodium hexametaphosphate, an antiscalant.
Nanofiltration (NF) - A crossflow process with pore sizes designed to remove selected salts
and most organics above about 300 molecular weight range, sometimes referred to as Loose
RO.
NOM - Natural Organic Matter.
Nephelometer - A device used to measure mainly the turbidity of water with results
expressed in Nephelometric Turbidity Units (NTU) measures light at an angle of 90o.
Non-carbonate Hardness - Hardness caused by chlorides, sulfates, and nitrates of calcium
and magnesium.
Nonionic Polyelectrolyte - Neutral charged polymers usually polyacrylamides, used for
coagulation/flocculation. (See Polyelectrolytes).
Normalization - Converting actual data to a set of reference conditions in order to
"standardize" operation to common base.
NTU - See Nephelometer
(Top)
O
OEM - Original Equipment Manufacturer.
O&M - Operation and Maintenance.
Ohm - Unit of electric resistance equal to the resistance of a circuit in which a potential

difference of one volt produces a current of one ampere.


Defined by the equation:
R = E/I where:
R = resistance (ohms)
E = electrical potential (volts)
I = current (amps)
Operating Pressure - The gauge hydraulic pressure at which feedwater enters a device.
Organic Scavenger - Macroporous anion resin in the chloride form that removes some
organic compounds from water.
Osmosis - The spontaneous flow of water from a less concentrated solution to a more
concentrated solution through a semipermeable membrane until chemical potential
equilibrium is achieved.
Osmotic Pressure - A measurement of the potential energy difference between solutions on
either side of a semipermeable membrane. A factor in designing the operating pressure of
reverse osmosis equipment. The applied pressure must first overcome the osmotic pressure
inherent in the chemical solution in order to produce any flux.
Oxidation-Reduction Potential (ORP)- The electromotive force developed by a noble metal
electrode immersed in the water, referred to the standard hydrogen electrode.
Oxygen Demand - The amount of oxygen required for the oxidation of waterborne organic
and inorganic matter under the specified test conditions.
(Top)
P
PAC - Powdered Activated Carbon or poly-aluminum chloride.
Parts per Billion (ppb) - A measure of proportion by weight, equivalent to a unit weight of
solute per billion unit weights of solution (approximate mg/l or mg/m3 in dilute solutions).
Parts per million (ppm or mg/liter) - The unit used to identify the total dissolved solids
(TDS) in water, and is typically used to quantify the purity of feed water entering the EDI
module. At low conductivity, 1 ppm is approximately 2 microsiemens/cm.
Parts per Trillion (ppt) - A measure of proportion by weight, equivalent to a unit weight per
trillion unit weights of solution (approximate hg/l or mg/m3 in dilute solutions).
Pass - A treatment step or one of multiple treatment steps producing in a membrane system a
product stream.
Permanent Hardness - The total milliequivalents of hardness ions minus the milliequivalents of bicarbonate alkalinity in a water. (See Hardness, Alkalinity).
Permeable - Allowing material to pass through.
Permeate - That portion of the feed stream which passes through a membrane.

Permeate Collector Fabric - See Product (Permeate) Channel Spacer.


Permeator - A reverse osmosis module of the hollow fiber configuration consisting of
membrane (s) and pressure vessel.
Pervaporation - A separation process involving vaporization of one liquid from a mixture of
two or more liquids, with the aid of a membrane which functions as a barrier to the liquid
phase.
pH - The measure of the acidity or basicity of a solution, based upon the concentration of
hydrogen or hydronium ion. The equation for pH is: pH = -log [H+]
pH values are expressed on a scale from 0 (very acidic) to 14 (very basic). pH
is a logartithmic measurement, so a solution with a pH of 5 is ten times as
acidic as a solution with a pH of 6.
Phase - A state of matter, either solid, liquid, or gaseous.
Plant Capacity - Manufacture of product per unit time, expressed as m3/day, m3/h, GPD,
MGD.
Plugging Factor - See Fouling Factor and SDI.
Polarization - The splitting of water into H+ and OH- ions with an electrical current. This
will occur when there are relatively few ions present in the dilute compartments and the
applied voltage is excessive, causing water to disassociate in order to conduct current.
Fluctuation of pH in the module is typically associated with polarization. Polarization of
water regenerates the ion-exchange resins.
See Concentration Polarization.
Polyelectrolyte - Synthetic (or natural) molecules, containing multiple ionic groups, used as
coagulants and flocculants; available as anionic, cationic and nonionic.
Polymers - A substance consisting of molecules characterized by the repetition of one or
more types of monomeric units.
Polyvalent Ion -An anion or cation with more than one valence (e.g. Ca+2, Mg+2, Al+3).
Pore - An opening or void in a membrane or filter matrix.
Porous - Substances containing pores for fluids to pass due to an open physical structure.
Porosity - That portion of a membrane filter volume which is open to fluid flow, also known
as void volume.
Post Treatment - A process of applying chemical(s) to a membrane after formation to
improve its performance.
Post Treatment - The addition of chemicals to the product or concentrate stream to make it
suitable for the desired end use application.
Post Treatment - Utilization of equipment such as degasifiers to make the product and/or
concentrate stream suitable for the desired end use application.
Power Requirement - Electrical power needed to supply the necessary current and voltage.

Typically expressed as kWh/1000 gallons.


Power Supply - A device which converts an AC input to a DC output, which acts as the
driving force for removal of ions in the Ionpure module.
Precipitate - An insoluble product of a chemical reaction of soluble compounds in water.
Precoat - The initial coating of the septum in a diatomaceous earth filter to provide initial
straining medium.
Pressure Filtration - Filtration performed in an enclosed pressurized filter vessel.
Pressure Vessel - The vessel containing one or more individual membrane elements and
designed to withstand safely the hydraulic pressure driving the separation mechanism.
Pretreatment - Processes such as chlorination, filtration, coagulation, clarification,
acidification which may be used on feedwater ahead at membrane devices to improve quality,
minimize scaling and corrosion potential, control biological activity.
Product - The flow of water through the purifying or dilute compartments. This is where
deionization occurs.
Product Staging - A process in which the permeate from one membrane plant is used as the
feed to another membrane plant in order to further improve product quality.
Product Channel Spacer (Permeate Carrier) - The fabric or other material through which
permeate water flows after it passes through the flat sheet membrane.
Productivity - Flow rate of product water.
Product Tube - The tube at the center of the spiral wound cartridge which collects permeate
water.
Product Water - Purified water produced by a process. (See Permeate).
Projection - A calculation, usually performed by a software package, which predicts the
performance of parts or all of a water plant.
Pyrogens - Any substance capable of producing a fever in mammals. Often a bacterial
endotoxin such as a lipopolysaccaride generated by gram negative bacteria at destruction.
Chemically and physically stable, pyrogens are not necessarily destroyed by conditions that
kill bacteria.
(Top)
Q
Quicklime - CaO, calcium oxide.
(Top)
R
Ranney Collector - An underground water collection system sometimes called Ranney

Wells.
Raw Water - Water which has not been treated. Untreated water from wells, surface sources,
the sea or public water supplies.
Recovery - Y (conversion) - The ratio of product quantity (permeate stream flow rate) over
the feed quantity (feed stream flow rate), given as fraction or in percent.
Regeneration - In ion exchange systems, the process of using either an acid, alkali, or salt
solution to remove the accumulated cations or anions. The cation exchange resins take on
hydrogen or sodium ions and the anion exchange resins take on hydroxide ions to restore
themselves to the original hydrogen or hydroxide form when using strong acid and strong
alkali solutions for the process.
Reject - Brine, (concentrate) stream from a desalination device. Portion of the feed stream
which does not pass through the membrane.
Resin Particle - Specially manufactured polymer beads used in the ion exchange process to
remove dissolved salts from water.
Resistivity - The electrical measurement of water's ability to resist the flow of electrical
current. Resistivity increases as the concentration of ions decreases. It is highly affected by
temperature. This measurement is associated with the level of deionization achieved with
EDI. Ultrapure water with no impurities may reach 18.24 MW or Megohm.cm at 25 deg C.
Resistivity is the inverse of conductivity, and is measured in megohm-cm.
Retentate - See Concentrate.
Reverse Osmosis (RO) - The separation process where one component of a solution is
removed from another component by flowing the feed stream under pressure across a
semipermeable membrane. RO removes ions based on electro chemical forces, colloids, and
organics down to 150 molecular weight. May also be called hyperfiltration.
RO Train - One of two or more complete RO installations, including membranes and high
pressure pump operating in parallel.
Ryznar Stability Index (RSI) - An index indicating if a water has a tendency to corrode or
precipitate CaCO3; equals 2 pH (CaCO3 saturation) - pH (actual), RSI < 6.0 scale formation,
>7.0 corrosive.
(Top)
S
Salinity - The concentration of inorganic salts in water.
Salt Flux - Amount of dissolved salt passing through the membrane, moles per day per
square unit of membrane area.
Salt Passage - The ratio of product (CP) and feed (CF) salt concentrations expressed as
percent. SP = (CP/CF) x 100

Salt Rejection - , SR= (100 - salt passage) expressed as percent,


Sanitization - Reduction in the number of bacterial contaminants to safe levels, see
Disinfection.
Saturation - The point at which a solution contains enough of a dissolved solid, liquid, or gas
so that no more will dissolve into the solution at a given temperature and pressure.
SBS - Sodium bisulfite, NaHSO3.
Scale Inhibitor - A chemical which inhibits the growth of micro-crystals (inhibits
precipitation of sparingly soluble salts). See Antiscalant.
Scaling - The build-up of precipitated salts on a surface, such as membranes, pipes, tanks, or
boiler condensate tubes.
SDI - Silt Density Index. An index calculated from the rate of plugging of 0.45um membrane
filter. it is an indication of the amount of particulate matter in water, sometimes called
Fouling Index.
S&DSI - Stiff and Davis Saturation Index, measure of CaCO3 solubility in seawater or highly
saline water. See also LSI.
SDWA - Safe Drinking Water Act of the United States, specifying required purity levels of
municipal potable water.
Sea Water - Water with a approximate concentration of total dissolved solids ranging from
30,000 to 60,000 mg/L. (See Brackish Water and High Brackish Water).
Sedimentation - The precipitation or settling of insoluble materials from a suspension, either
by gravity or artificially e.g. centrifuge, pressure.
Semipermeable Membrane - A membrane which preferentially allows the passage of
specific compounds while rejecting others.
SHMP - Sodium hexametaphosphate. (NaHMP)
Siemens - A measure of electrical conductance in water, equivalent to a mho. (See Mho,
Ohm).
Slime - Biological deposits of gelatinous or filamentous matter.
Sludge - A water-formed sedimentary deposit.
Sludge Blanket - Suspended bed of solids in a solids contact or sludge blanket clarifier.
SBS - Sodium bisulfite, NaHSO3.
SMBS - Sodium metabisulfite, Na2S2O5.
STP - Sodium Triphosphate - Na5P3O10, a cleaning agent.
STPP - Sodium Tri Poly Phosphate - See STP.
Softening - See Membrane Softening.
Softener - Water treatment equipment that uses a sodium based ion-exchange resin
principally to remove cations as calcium and magnesium.
Solids Contact Clarifier - Water treating device used in lime softening, waste water

treatment and coagulation processes.


Solubility Product - [M+]a [X-]b/[MX] where the brackets indicate the concentrations of the
components of the ionization equilibrium MaXb _ aM+ +bX. For sparingly soluble salts
[MX] is essentially unity.
Solutes - Matter dissolved in a solvent.
Solvent - Here defined as water.
SPC - Standard (Heterotrophic) Plate Count - Measurement method for enumerating bacteria.
Specific Gravity - The ratio of the mass (density) of a sample material to the mass (density)
of an equal volume of water at the same specified temperature.
Specific Flux - Flux divided by net pressure driving force.
Spiral Wound Cartridge - A crossflow membrane element design consisting of a product
tube, flat membrane leaves, feed channel spacers, anti-telescoping devices, and brine
(concentrate) seal.
Spiral Wound Membrane - A flat sheet membrane with one or more feed channel spacers
and barrier layers, all of which are rolled into a spiral configuration.
Stage - A sequential arrangement of pressure vessels, usually reject staged such as 2:1 array,
sometimes permeate staged as in double pass RO.
Staging - See Brine Staging and Product Staging.
Standard Test Conditions - The parameters under which a membrane manufacturer tests
devices for flow and salt rejection.
Sterilization - Destruction or removal of all viable organisms.
Stiff & Davis Stability Index, S&DSI - An index calculated from total dissolved solids,
calcium concentration, total alkalinity, pH and solution temperature that shows the tendency
of a water solution to precipitate or dissolve calcium carbonate. S&DSI is used primarily for
seawater RO applications.
Supersaturation - A state in which the inorganic salt (s) are in solution at a level higher than
the respective solubility product.
Suspended Solids (SS) - Solid organic and inorganic particles that are held in suspension in a
liquid.
SWRO - Seawater Reverse Osmosis.
Symmetric Membrane - Membrane and bulk polymer have equivalent characteristic
(isotropic).
(Top)
T
Tangential Flow - Action whereby the untreated water (feed) flow is directed across the
surface of the membrane. This "sweeping" action helps keep material retained by the

membrane from settling, and eventually restricting, permeate flow (fouling).


TBC - Total Bacteria Count, the total number of viable microorganisms present in the
sample, excluding anaerobic organisms.
TCC - Total Colony Count.
TDS - Total Dissolved Solids, usually expressed as mg/L or ppm (parts per million).
Telescoping - The movement of the outer layers of a spiral wound cartridge in the direction
of the feed flow caused by excessive pressure drop through the feed channel spacer.
Temperature Correction Factor (TCF) - Defines the effect of temperature on permeate
flow relative to a base temperature (25oC), is mainly a function of fluid characteristics but
also membrane polymer.
Temporary Hardness - Usually the bicarbonate salts of calcium and magnesium.
Thickener - A vessel designed to concentrate sludges; similar to a clarifier.
Thin Film Composite (TFC) - See Composite Membrane.
Threshold Treatment - The process of stopping precipitation at the start of occurrence,
usually does not stop the formation of nuclei but does inhibit growth. (See Antiscalant).
THM - Trihalomethanes; a group of low molecular weight molecules which can result from
chlorination of organics typically found in surface water.
THMP - Trihalomethane Precursors; organic molecules found in water which have the
potential of reacting with chlorine to form THMs.
Thrust Collar - A plastic cylinder placed between the last spiral wound cartridge and vessel
end plate to support the last cartridge in a pressure vessel against telescoping.
TOC - Total organic carbon. A quantitative measure of the level of organic compounds in a
water sample. The total inorganic carbon (CO2) is subtracted from the total carbon measured
in a sample. Expressed as ppm or mg/liter.
TOCI - Total Organic Chlorine.
TOX - Total Organic Halides.
TOXFP - Total Organic Halide Formation Potential.
Train - A grouping of devices (See Array, Bank, Block).
Transmembrane Pressure - The net driving force across the membrane. The hydraulic
pressure differential from the feed side to permeate side less the osmotic pressure differential
on each side.
TRC - Total Residual Chlorine.
Trisodium Phosphate (TSP) - Na3PO4 . 12H2O, a cleaning agent.
TSS - Total Suspended Solids. Concentration of undissolved solids in a liquid, usually
expressed in mg/l or ppm.
TTHM - Total Trihalomethane.
Turbidity - A suspension of fine particles that scatters or absorbs light rays.

Turbidity, Jackson Candle (JTU) - An empirical measure of turbidity in special apparatus,


based on the measurement of the depth of a column of water sample that is just sufficient to
extinguish the image of a burning standard candle observed vertically through the sample.
Turbidity, Nephelometric (NTU) - An empirical measure of turbidity based on a
measurement of the light-scattering characteristics (Tyndall effect) of the particulate matter in
the sample.
Tyndall Effect - The path of light through a heterogeneous medium made visible by the solid
particles.
(Top)
U
Ultrafiltration (UF) - A process employing semipermeable membrane under a hydraulic
pressure gradient for the separation of components in a solution. The pores of the membrane
are of a size which allow passage of the solvent(s) but will retain non-ionic solutes based
primarily on physical size, not chemical potential.
Ultra Pure Water - Water generally used in semiconductor industry having specifications,
chemical, physical and biological, for extremely low contaminant levels.
USEPA - U.S. Environmental Protection Agency.
Ultraviolet (UV) Radiation - Wave lengths between 200-300 nm. These wave lengths have a
strong germicidal effect, the maximum effect is at 253.7 nm.
(Top)
V
Viable - Ability to live or grow e.g. bacteria, plants.
VOC - Viable Organism Count - A measure of biological activity (living or growing) in
water.
VOC - Volatile Organic Compound - An organic compound with a vapor pressure higher than
water.
(Top)
W
WQA - Water Quality Association.
Water Softener - A vessel having a cation resin in the sodium form that removes cations
such as calcium and magnesium from water and releases another ion such as sodium. The
resin is usually regenerated. (See Softener).
(Top)

X-Y-Z
Y - Conversion, recovery.
Zeolite - Any of various natural or synthetic hydrated aluminum silicates used as ion
exchange substrates in water softening.
Zero Discharge - A condition whereby a facility discharges no process effluent.
Zeta Potential - Colloidal stability measured in millivolts. High negative value (-10 to
-30mv) results in particulate stability.
Asked Questions
This section will address some questions commonly asked. If you do not find the answer to
your question, please contact
techcenter@cediuniversity.com
.
Q: What is Electrodeionization?
A: Electrodeionization is the chemical-free process in which ions are removed from water via
electric potential. Ions are forced from the dilute compartments into adjacent reject
compartments resulting in high purity water. This electric potential also constantly
regenerates the resin in the module rather than conventional ion exchange units which use
acid and caustic to regenerate. EDI is currently designed to polish RO permeate (product).
Q: What are the advantages of using EDI rather than conventional mixed bed
deionization?
A: EDI is a continuous process and the only thing you need is electricity to regenerate the
resin. Conventional ion exchange is a batch process which requires large vessels filled with
resin, acid/caustic injection systems, a neutralization waste system and all associated controls.
Such chemicals can be costly, hazardous and typically requires more space than the
equivalent EDI system.
Q: Can EDI Systems handle small amounts of free chlorine the same way TFC RO
membranes can?
A: Actually, neither can tolerate free chlorine. We have seen damage to both RO membranes
and EDI Modules at concentrations as low as 0.05 ppm. However, it is possible for EDI
Module damage to be seen before the RO membranes show signs of chlorine attack.
Q: Why are the limits for hardness concentration in the EDI system feed water so low?
A: Scale can form inside the EDI Module concentrate compartments even at relatively low
hardness concentrations because of the high recovery of the EDI System and pH shifts which
can occur in the module.
Q: Why are there limits on the iron and manganese concentrations going into EDI
system? Aren't these materials removed by ion exchange?
A: Iron and manganese can be removed by ion exchange but can cause problems in a EDI
Module because they are held tightly by the resins and may oxidize and precipitate in the

resin before they can be transferred out to the waste stream.


Q: Why must the feed water temperature be at least 41oF (5oC)? Why is the minimum
temperature specification even higher when silica removal is important?
A: The electrical resistance of the EDI system increases with decreasing temperature. The
power supplies for EDI systems are sized to compensate for this down to a temperature of
41oF (5oC) below that temperature the performance may decline due to DC voltage
limitations.
Q: Can EDI Systems be installed outdoors?
A: EDI systems must be sheltered from direct sunlight. Also, if ambient temperatures
routinely exceed 95oF (35oC) or go below 41oF (5oC) it is not recommended to install the EDI
system outdoors. In any outdoor installation, special control enclosures are required to protect
the EDI system from humidity, excessive temperatures and other natural elements (rain, sleet,
snow).
Q: Is it OK to install a EDI system in the distribution loop instead of directly after the
RO system?
A: Yes, EDI systems can be and have been installed in a distribution loop. However, the
reject stream from the EDI system results in 5% of the loop water being continuously drawn
out of the loop. This water can be returned ahead of the system for re-use.
Q: Do EDI Systems have a difficult time removing CO2? Why is the CO2 concentration
in the EDI feed water so important?
A: CO2 is removed by EDI systems just as dissolved salts such as sodium chloride are. We
call particular attention to the feed water CO2 concentration for two main reasons. The first is
that CO2 is weakly ionized in water so its concentration is not measured by conductivity
measurements. The second reason is that CO2 is often a significant portion of the load on the
EDI System because it is not removed by RO systems.
Q: Can I raise the pH of the EDI system feed water with caustic to eliminate the CO2?
A: pH adjustment has to be done in the feed to the RO system to improve the EDI
performance. Raising the pH of RO permeate will not improve the EDI product water quality.
Raising the pH before the RO allows the RO to remove the CO2 as bicarbonate, preventing
the CO2 from reaching the EDI.
Q: How much back pressure can a EDI module handle?
A: The back pressure on the EDI module is limited only by the requirement that the inlet
pressure be 100 psi or below. So, if the EDI System is operating at a flow rate which results
in a 20 psi pressure drop the maximum back pressure would be 80 psi.
Q: How long does the EDI System take to come back up to quality after being in standby?
A: Typically a EDI System will return to quality within a few seconds of starting up from
stand-by. It takes much longer however, to come back to quality after cleaning or sanitization,

several hours or overnight.


Q: What is the recommended frequency of sanitization for EDI Systems?
A: The frequency of sanitization varies greatly with the customer's bacteria specifications and
the system design. Some systems are never sanitized but typical frequencies for systems
producing Purified Water range from weekly to quarterly.
Q: What is the typical life of the EDI resin and membrane cell packs?
A: Three to five years is considered a normal life for the cell packs. The useful life obtained
at any given site will be a function of the raw water quality, the pretreatment design and how
well the system is monitored and maintained.

You might also like