You are on page 1of 54

Chlor- Alkali production by electrochemical process

ABSTRACT

The chlor-alkali industry is the industry produces chlorine gas (Cl2), hydrogen gas (H2), and
alkali; sodium hydroxide (NaOH) and to less extent potassium hydroxide (KOH) by
electrolysis of a salt solution using direct current (DC). The salt solution is sodium chloride
or potassium chloride. The main technologies applied for chlor-alkali production are mercury
cell, diaphragm cell and membrane cell techniques. The amount of electricity required for the
electrolysis process depends on electrolytic cell parameters such as current density, voltage,
and anode and cathode material.

i
ii
INTRODUCTION

Electrolysis reactions are the basic foundations of today's modern industry. There are various
elements, chemical compounds that are only produced by electrolysis and are an
electrochemical process in which a direct current is passed between two electrodes through
an ionized solution (electrolyte). It deposits positive ions (cations) on the negative electrode
(cathode) and negative ions (anions) on the positive electrode (anode). The processes are used
in many industries. Here are three examples: chlor-alkali industries, sodium and Aluminum
industries [1]. The chlor-alkali electrolysis process industry is used in the manufacture of
chlorine, sodium hydroxide (caustic soda) solution (an alkali) and hydrogen hence the term
chlor-alkali. The term chlor-alkali industry also includes the production of chlorine with
potassium or lithium hydroxide. The primary product is chlorine; chlorine is one of the more
abundant chemicals produced by industry and has a wide variety of industrial uses [2]. The
technologies applied for chloro-alkali production are mercury cell, diaphragm cell and
membrane cell. In all 3 methods, the chlorine (Cl2) is produced at the positive electrode
(anode) and the caustic soda (NaOH) and hydrogen (H2) are produced, directly or indirectly,
at the negative electrode (cathode).The 3 processes differ in the method by which the anode
products are kept separate from the cathode products. Chloralkali raw material is sodium
chloride salt solution or brine as feed and to a lesser extent potassium or lithium chlorides
salt.Chlor-Alkali products are precursors for many leading chemical industries all over the
world [7].

1
CHAPTER

CHLOR-ALKALI RAW MATERIAL

The basic raw material is the natural occurring salt lakes of sodium chloride (brine), lake
katwe and rock salts obtained by mechanical mining as sodium chloride and are all raw
materials for chlor-alkali industry. Solar evaporation of seawater or brine is another source of
raw material salts. Sodium chloride (NaCl) is the major raw material in chlor-alkali industry
with the exception when potassium or lithium hydroxide is the desired products. Mercury cell
and membrane cell plants mostly used vacuum salt and rock salt solution; diaphragm cell
plants used solution-mined brine.

Brine Preparation
When solid salt is the raw material, a dissolving operation becomes necessary, and this may
be carried out in open or closed vessels. The water and/or depleted brine can be sprayed onto
the salt or introduced at the base of the saturator for progressive saturation when running
through it. In the latter case, the saturated brine overflows the equipment at the top. Modern
saturators are closed vessels to reduce emissions of salt spray or mist, as well as of mercury
in the case of the mercury cell technique. Sodium chloride (NaCl) concentrations in the
saturated brine reach values of 310–315 g/l.

Chlor-Alkali Raw Material Chemical Treatment


The natural brine solution and the prepared solution by mixing rock salt of sodium chloride
with water contain traces of divalent ions such as magnesium, calcium, sulfate, and barium
ions that must be removed before the electrochemical process in the electrolytic cell to prevent
fouling. Figure 1illustrates a simple brine treatment block diagram, table 1 shows typical
sodium chloride chemical composition used in chlor-alkaly electrolysis.

2
B5
B4
WATER B6
B8
3
5
6
RAWSALT

Figure 1 Simple brine treatment block diagram

B4=Salt dissolution

B5= primary Brine treatment precipitation

B6= secondary brine treatment filtration and exchange

B8= Pure brine for electrolysis

Table 1. Typical compositions of sodium chloride used in chlor-alkali electrolysis [7]

component Salt source


Rock salt Washed solar Vacuum salt
salt
NaCl 93-99% 99% 99.95%
SO42- 0.2-1% 0.2% 0.04%
Ca2+ 0.05-0.4% 0.04% 0.0012%
Mg2+ 0.01-0.1% 0.01% 0.0001%

Primary purification

The initial stage of purification uses sodium carbonate and sodium hydroxide to precipitate
calcium and magnesium ions as calcium carbonate (CaCO3) and magnesium hydroxide
(Mg(OH)2). Metals (iron, titanium, molybdenum, nickel, chromium, vanadium, and tungsten)
may also precipitate as hydroxide during this operation. The usual way to reduce the
3
concentrations of metals is to specify maximum concentration values in the purchase
specifications for the salt. Sulphate ion is precipitate by adding calcium chloride (CaCl2) or
barium salts as barium carbonate (BaCO3) or barium chloride(BaCl2). The precipitation of
barium sulphate can take place simultaneously with the precipitation of calcium carbonate
and magnesium hydroxide, whereas the precipitation of calcium sulphate requires a separate
vessel. When vacuum salt is used as raw material, only a part of the brine stream might be
treated in the primary purification unit, while the total stream is usually treated when using
other salt types. Some plants using vacuum salt omit primary brine purification completely.
The sulphate content can also be reduced without the use of expensive barium salts by purging
a part of the brine, by cooling the brine stream and crystallizing Na2SO4·10H2O, by
precipitating the double salt Na2SO4·CaSO4, by ion exchange, or by nanofiltration combined
with purging of the brine. In the diaphragm cell technique, the removal of sulphate is not
always necessary because sulphate can be removed from the cell liquor as pure Na2SO4 during
the concentration process. In the case of the membrane cell technique, the use of barium salts
is generally avoided to protect the membrane against potential precipitations .Chemical chlor-
alkali reaction treatment is shown in the chemical equations below.

Calcium is precipitated as calcium carbonate by reaction with sodium carbonate.

Ca+2 + Na2CO3 ------------> CaCO3(s) + 2Na+


Magnesium is precipitated as magnesium hydroxide by reaction with caustic soda.

Mg+2 + 2NaOH ------------> Mg(OH)2(s) + 2Na+

Sulphates are precipitated as barium sulphate by reaction with barium carbonate.


SO4-2 + BaCO3 ------------> BaSO4(s) + CO3-2

Strontium, is precipitated mainly as strontium carbonate.

Sr+2 + CO3-2 ------------> SrCO3(s)


The solids are then removed using a combination of clarifying and filtration steps. This step
will also remove heavy metals if they are present, pH is usually maintained in the 9–11 range
at this stage to optimize the removal of the contaminants, and control is achieved by regulating
the hydroxide feed rate. The purified brine should ideally contain [ 5].

4
•Calcium (Ca2+): < 2 mg/l
•Magnesium (Mg2+): < 1 mg/l
•Sulphate (SO42-): < 5 g/l
Following the filtration steps, the brine is purified by ion exchange chelation, which reduces
the calcium and magnesium to ug/l or parts per billion (ppb) levels. Current technology calls
for brine with 20 ppb levels of calcium and magnesium to enable the membranes to have a
useful performance.

Before the brine enters electrolysis cells, it is usually acidified with hydrochloric acid to
maintain a pH less than 6 (pH < 6), which increases the lifetime of the anode coating and
reduces the formation of oxygen, hypochlorite ion and chlorate ion.

Secondary purification: membrane cell technique

To maintain the high performance of the ion-exchange membrane in the membrane cell
technique, the feed brine must be purified to a greater degree than in the mercury or diaphragm
cell techniques. The precipitation step alone is not enough to reduce the levels of calcium and
magnesium, and additional softening is thus required. Figure 2 shows the flow diagram of a
possible layout for the brine system used in the membrane cell technique. The secondary brine
purification generally consists of polishing filtration and brine softening in an ion-exchange
unit. The polishing filtration generally consists of candle-type, plate frame or pressure leaf
filters (either with or without a cellulose-based pre-coat) in order to sufficiently reduce
suspended particles and protect the ion-exchange resin from damage. The ion-exchange
chelating resin treatment is designed to decrease the sum of magnesium and calcium
concentrations to less than 20 μg/l; part per billion levels (ppb).

5
Figure 2 Flow diagram of a possible layout for the brine system used in the membrane cell technique
Table 2 indicates typical specifications required for metals, sulphate and other impurities.
These specifications can vary if the users want to operate at a low current density (< 4 kA/m2)
or at a high current density. The specifications are more stringent for high current densities.
Specifications also depend on the interaction of impurities. While the presence of one impurity
may not be harmful, its synergistic combination with others may be (e.g. the combination of
aluminum, calcium and silica). The resin is periodically regenerated with high-purity
hydrochloric acid and sodium hydroxide solutions. Generally, one resin exchange column is
in operation while another resin exchange column is regenerated.

Control of nitrogen compounds in the brine

6
The presence of some nitrogen compounds in the brine gives rise to the formation of nitrogen
trichloride (NCl3), which is an explosive substance. Techniques are applied to reduce the
concentration of nitrogen compounds in the brine.

Table 2. Typical impurities with sources and effects on the membrane cell technique and brine
specifications

Impurity Source Typical upper Effects Mechanism


limit of brine
specifications

Ca2+ + Salt 20 pbb Ca CE Ca: Precipitation with various anions


Mg2+ Mg V near the cathode side of the
membrane, precipitation with silica
and iodine in the membrane Mg:
Fine precipitation with OH- near the
anode side of the membrane,
precipitation with silica in the
membrane
St2+ Salt 0.1–4 ppm CE Precipitation with hydroxide on the
cathode side of the membrane,
precipitation with silica and iodine in
the membrane
Ba2+ Salt 0.05–0.5 CE Very fine precipitation with iodine in
ppm the mem- brane, precipitation with
silica in the membrane
Al3+ Salt 0.1 ppm CE Precipitation with silica in the
membrane, precipitation of
calcium/strontium aluminosilicates
near the cathode side of the
membrane
Fe3+ Salt, pipe work, 0.05–0.1 ppm V Deposition on the cathode,
tank precipitation with hydroxide on the
material, anti- anode side of the membrane or in the
caking agent membrane (depending on pH of the
brine)
Hg2+ Parallel 0.2 ppm V Deposition on the cathode
operation of heavy metals

7
mercury cell
plant

Ni2+ Salt, pipe work, 0.2 ppm heavy V Deposition on the


tank metals cathode,
material, absorption in the membrane
cathode
ClO3- Process 10 g/l 0 Chlorination of the ion-exchange
side (as resin
reactions NaClO3)
I2 Salt 0.1–0.2 ppm CE, V Very fine precipitation with calcium,
(e.g.H2IO6) strontium or barium in the membrane,
precipitation with sodium on the
cathode side of the membrane
F- Salt 0.5 ppm V Destruction of the anode coating
SO42- Salt < 4–8 g/l (as CE Precipitation with sodium near the
dichlorination Na2SO4) cathode side of the membrane, anode
with NaHSO3 coating with barium
SiO2(e.g. Salt 10 ppm CE Silica itself is harmless, but in the
SiO32- ) presence of magnesium, calcium,
strontium,
barium or aluminum, silicates can be
formed (see above)
Suspended Salt 0.5–1 ppm V Precipitation on the anode side of the
solids membrane
Total Salt 1–10 ppm V Increased foaming, over plating
organic
carbon
NB: CE = current efficiency decreases; O = other effects; V = cell voltage increases. Source: [ 7] •Brine
dichlorination and resaturation

Brine dechlorination and resaturation


Mercury and membrane cell plants usually operate with brine recirculation and resaturation.
Some plants operating on a once-through basis, one of them use mercury cell technique, the
membrane cell technique or use both the mercury and the membrane cell technique.
Diaphragm cell plants always use a once through brine circuit, but some employ brine
saturation using the salt recovered from the caustic evaporators.

In recirculation circuits, the depleted brine leaving the electrolyzer is first dechlorinated:

8
• Only partially for the mercury cell technique (leaving active chlorine in the brine keeps
the mercury in its oxidized form as HgCl3- and HgCl42- and avoids the presence of
metallic mercury in the brine purification sludge.

• Totally for the membrane cell technique (necessary here because the active chlorine
can damage the ion-exchange resins of the secondary brine purification unit). For this
purpose, the brine containing 0.4–1 g/l of dissolved chlorine is generally acidified to
pH 2–2.5 and sent to an air-blown packed column or sprayed into a vacuum system of
50–60 kPa to extract the majority of the dissolved chlorine to a residual concentration
of 10–30 mg/l. The chlorine-containing vapors are subsequently fed back to the raw
chlorine collecting unit or directed to the chlorine absorption unit. The water that
evaporates from the dechlorinated brine is condensed in a cooler. The condensate can
be sent back to the brine system or chemically dechlorinated.

For the membrane cell technique, complete dichlorination is achieved by passing the brine
through an activated carbon bed, by catalytic reduction, or by using chemical reducing agents
such as sulphate. Residual levels were reported to be < 0.5 mg/l or below the detection limit.
No such dichlorination treatment is required for the diaphragm system, since any chlorine
passing through the diaphragm reacts with caustic soda in the catholyte compartment to form
hypochlorite or chlorate.

If the saturation is carried out with impure salt (followed by a primary purification step of the
total brine flow), the pH of the dechlorinated brine is then brought to an alkaline value with
caustic soda to reduce the solubilization of impurities from the salt. If the saturation is carried
out with pure salt (with subsequent primary purification on a small part of the flow), there is
no alkalization step prior to the resaturation (only in the purification phase). In the case of
mercury or membrane cell plants operating with solution-mined brine, brine resaturation is
achieved by evaporation. During this step, sodium sulphate precipitates and can be recovered,
purified and used for other purposes. In the case of diaphragm cells, the catholyte liquor (10–
12 wt-% NaOH, 15–17 wt-% NaCl) is directly used or transferred to the caustic evaporators,
where solid salt and 50 wt-% caustic are recovered together. Fresh brine may be saturated
with recycled solid salt from the caustic evaporators before entering the diaphragm
electrolyzer.

9
Chlorate destruction: membrane cell technique
In order to reduce the build-up of chlorate in the brine circuit, which could have negative
effects on the ion-exchange resins, the caustic quality, and on emissions to the environment,
some membrane cell plants operate a chlorate destruction unit prior to the dichlorination.
Techniques include the reduction of chlorate to chlorine with hydrochloric acid at
temperatures higher than 85°C and the catalytic reduction of chlorate to chloride with
hydrogen.

10
CHAPTER
CHLOR-ALKALI ELECTROLYSIS PROCESS
TECHNIQUES

There are three basic techniques for the electrolytic production of chlorine, caustic soda and
hydrogen. These three techniques are the mercury cell, the diaphragm cell and the membrane
cell. The three techniques differ from each other in terms of electrode reactions, the amount
of electricity and in the way the produced chlorine and caustic soda/hydrogen are kept
separate. The products of the electrolysis are formed in a fixed stoichiometric ratio, which is
1070–1128 kg of caustic soda; NaOH (100 wt- %) and approximately 28 kg of hydrogen gas
(H2) per one ton of chlorine gas (Cl2) produced. This product combination is often referred to
as the electrochemical unit (ECU). Figure 3 is a simple block diagram of the three chlor-alkali
process techniques.

11
Figure 3 Typical flow diagrams of the three cells techniques
Mercury Cell Process Technique

General description

The mercury cell technique includes an electrolysis cell and a horizontal or vertical
decomposer as shown in figure 4. A 25 wt-% sodium chloride brine having pH 2-5 enters the
electrolysis cell at 60–70 °C. The temperature can be achieved by preheating the saturated
brine, and is increased in the cell by the heat of resistance to approximately 75–85 °C. At this
temperature, the conductivity of the brine solution and the fluidity of the mercury are higher
compared to operation at ambient temperature. The cell operation voltage is 3.15-4.80 V and
current density 2.20-14.50 kA/m2, total electrical power 3 MW/h per one ton produced
chlorine gas (Cl2) [7].

In the electrolysis cell, the brine flows through an elongated trough that is slightly inclined.
In the bottom of this trough the cathode which is a shallow film of mercury metal (Hg) flows
along the brine cell together with the brine. Closely spaced above the cathode, a titanium
anode assembly is suspended as shown in figure 4 [1].

Figure 4. Mercury cell

Electric direct current flowing through the cell decomposes sodium chloride in the brine
passing through the narrow space between the electrodes, liberating chlorine gas (Cl 2) at the
12
anode and metallic sodium (Na) at the cathode. The chlorine gas is accumulated above the
anode assembly and is discharged to the purification process. As it is liberated at the surface
of the mercury cathode, the sodium (Na) immediately forms an amalgam with mercury
(Na/Hg) (a mixture of metal sodium and mercury metal). The concentration of the amalgam
is maintained at 0.2–0.4 wt-percent sodium so that the amalgam flows freely. Sodium
concentrations higher than 0.5 wt-percent can cause increased hydrogen evolution in the cells
[1]. The liquid amalgam flows from the electrolytic cell to a separate reactor, called the
decomposer or denuder where it reacts with demineralized water in the presence of a graphite
catalyst to form sodium hydroxide and hydrogen gas. The decomposer operates at a
temperature of approximately 90-130 ºC, which is caused by the chemical reactions in the
decomposer and the input of warm amalgam from the electrolyzer. The sodium-free mercury
is fed back into the cell and is reused. The depleted brine anolyte leaving the cell is saturated
with chlorine and must be partially dechlorinated before being returned to the dissolvers.

The sodium hydroxide is produced from the decomposer at a concentration of


approximately50 wt-percent [1]. For its operation, the mercury cell depends on the higher
overpotential of hydrogen on mercury to achieve the preferential release of sodium rather than
hydrogen. However, impurities such as vanadium (V), molybdenum (Mo), and chromium (Cr)
at the 0.01–0.1 ppm level and other elements (Al, Ba, Ca, Co, Fe, Mg, Ni, W) at the ppm level
that can contaminate the mercury surface may lack this overpotential protection and can cause
localized release of hydrogen into the chlorine. There is a risk that the hydrogen concentration
in the chlorine can increase to the point at which the cell and downstream chlorine handling
equipment contain explosive mixtures [1].

Mercury cells are usually operated to maintain a 21–22 wt-% concentration of salt in the spent
brine discharged from the cell. This corresponds to the decomposition of 15–16 % of the salt
during a single pass. Further salt decomposition to a lower concentration in the brine would
decrease brine conductivity, with the attendant loss of electrical efficiency. However, in plants
with once-through brine systems, approximately 40 % of the salt is electrolyzed in the cells
[1].

The mercury cathode electrolyzer and decomposer


The cell is made of an elongated, slightly inclined trough and a gas-tight cover. The trough is
made of steel, and its sides are lined with a protective, non-conductive coating to prevent

13
contact with the anolyte, to confine brine-cathode contact to the mercury surface, and to avoid
the corrosive action of the electrolyte. Modern electrolyzer is10-25 m long and 1 -2.5 m wide.
As a result, the cell area today can be greater than 30 m2. The size of the cells can be varied
over a broad range to give the desired chlorine production rate. At the design stage, computer
programs can be used to optimize the cell size, number of cells, and optimum current density
as a function of the electricity cost and capital cost [1]. The steel base is made as smooth as
possible to ensure mercury flow in an unbroken film. In the event of a break in the mercury
surface, caustic soda will be formed on the bare (steel) cathode, with simultaneous release of
hydrogen, which will mix with the chlorine. Because hydrogen and chlorine can forma highly
explosive mixture, great care is necessary to prevent hydrogen formation in the cell.

Characteristics of the cathode: The cathode is made by a shallow layer of mercury which flows
from one extremity of the cell to the other because of the slight inclination from the horizontal
of the cell. Characteristics of the anode: Electrolytic cell anodes made of titanium coated with
ruthenium oxide (RuO2) and titanium oxide (TiO2) were used of reduces energy consumption
by about 10% and their life expectancy is higher. In recent years there have been competitive
developments in detailed anode geometry, all with the aim of improving gas release in order
to reduce ohmic losses and increase the homogeneity of the brine to improve anode coating
life. Compartments for collecting the chlorine gas and weirs for separating the mercury and
brine streams, washing the mercury and permitting the removal of thick mercury “butter” that
is formed by impurities. The whole electrolyzer is insulated from the floor to prevent stray
ground currents. Usually, several electrolyzers are placed in series by means of electrically
connecting the cathode of electrolyzer to the anodes of the next electrolyzer. Individual cells
can be by-passed for maintenance and replacement. The decomposer may be regarded as a
short-circuited electrical cell in which the graphite catalyst is the cathode and sodium
amalgam the anode.

Mercury cell chemical reactions:


• Anode (positive electrode): Titanium
• Anode reaction (oxidation): 2Cl-(aq) → Cl2(g) + 2e-
• Cathode (negative electrode): Mercury flowing along bottom of cell
• Cathode reaction (reduction): 2Na+(aq) + 2e- → 2Na(s)
• The overall reaction is

14
2Na+(aq) +2Cl-(aq) →2Na(s) +Cl2(g)

Na(s)+ Hg(l) → Na/Hg(l)

2Na/Hg + 2H2O(l) → 2Na+ + 2OH- + H2(g) + 2Hg(l)

The problem:

Industrial installations consist of some 200 mercury cells in series, each one measuring 15 m
x 2 m x 0.3 m, the volume of each cell 9 m3 and the total volume is 1800 m3. The process,
which takes place at a very high voltage, uses an enormous quantity of electricity: 3 MW/h
per ton Cl2.This method only produces a fraction of the chlorine and sodium hydroxide used
by industry as it has certain disadvantages: mercury is expensive and toxic, and whilst it is
recirculated, some always escapes with the produced sodium hydroxide and the spent brine
with which it reacts to form mercury (II) chloride (HgCl2), if the effluent is discharge into
lakes, rivers and seas lead to the accumulation of high levels of mercury in fish, which
absorbed the mercury compound but could not re-excrete it, Now a days the spent brine is
treated before discharge, the mercury being precipitated as mercury(II) sulphide (HgS). In
recent years a large share of chlorine and sodium hydroxide production has been produced in
two other types of cells, which do not use mercury: the diaphragm cell and the membrane cell.

The Diaphragm Cell Technique

General description

In the diaphragm cell, a 25 wt-% sodium chloride brine having pH2.5-3.5 enters anode
compartment as shown in figure 5. The brine is decomposed to approximately 50 wt-% of its
original concentration in a passage through the cell. Cell voltage 2.90-3.60 V and current
density 0.80-2.70 kA/m2. Heating caused by passage of current through the diaphragm cell
raises the operating temperature of the electrolyte to 80-99 ºC. On the negatively charged
cathode surface caustic soda (NaOH) and hydrogen gas (H2) are formed directly and chlorine
gas (Cl2) is formed at the anode. To separate the chlorine from the sodium hydroxide, the two
half-cells were traditionally separated by a porous asbestos diaphragm, which needed to be
15
replaced every two months. This was environmentally detrimental owing to the need of
disposing of large quantities of asbestos. Such frequent replacement is fortunately no longer
necessary, the asbestos having now been replaced in part by metal oxide with polymers
resulting in diaphragms with a much longer life. Diaphragm cells generally produce cell liquor
that contains 10–12 wt-% NaOH and 15–17 wt-% NaCl. Generally, this solution is evaporated
to 50 wt-% sodium hydroxide. The caustic liquor produced may contain lower concentrations
of about 7 wt-% sodium chloride if it is used directly without further concentration [7]. During
evaporation, most of the sodium chloride precipitates, except a residual of approximately 1.0
wt-%. The salt generated is very pure and is typically used to make more brine. This high
quality sodium chloride is sometimes used as a raw material for mercury or membrane cells.
Low concentrations of oxygen (0.5– 2.0 vol-%) in chlorine are formed by the electrolytic
decomposition of water. Furthermore, chlorate is formed in the cell liquor by anodic oxidation
and the disproportionate of hypochlorous acid. (0.04–0.05 wt-% before concentration and ~
0.1 wt-% after concentration). However, a high amount of steam may be necessary for the
caustic soda concentration, and the quality of the caustic soda and chlorine produced are low.

Figure 5 Diaphragm cell


The cell

Various designs of diaphragm cells have been developed and used in commercial operations,
an example is the monopolar diaphragm cell the typical anode areas per cell range from 20
to100 m2and the cell voltage 2.90–3.60 V; current density 0.8–2.70 kA/m2[5].Cathodes used

16
in diaphragm cells consist of carbon steel with an active coating which lowers the hydrogen
overpotential, thus providing significant energy savings. The coatings consist of two or more
components. At least one of the components is leached out in caustic to leave a highly porous
nickel surface [5]. The coatings have to be robust, as a powerful water jet is used to remove
the diaphragm at the end of its lifetime from the cathode mesh, which can adversely affect the
coatings. Anodes used in diaphragm cells consist of titanium coated with a mixture of
ruthenium dioxide, titanium dioxide and tin dioxide. The lifetime of the coatings is at least 12
years [7]. The most commercially accepted design is that of an expandable anode, which
allows compression of the anode structure during cell assembly and expansion when the
cathode is in position. The spacers initially placed over the cathode create a controlled gap of
a few millimeters between the anode and cathode. Minimization of the gap leads to a reduced
power consumption. The lifetime of the diaphragm can be several years. Their service life has
also increased due to a change in composition. The earliest diaphragms were made of sheets
of asbestos paper. Due to the potential exposure of employees to asbestos and emissions to
the environment, efforts have been made to replace the asbestos with other diaphragm
materials. The basis of the material used is the same in all asbestos-free diaphragms, i.e. a
fluorocarbon polymer, mainly polytetrafluoroethylene (PTFE). The differences lie in the
fillers used and the way the hydrophobic PTFE fibers are treated and deposited in order to
form a permeable and hydrophilic diaphragm [5]. The diaphragm cell is now technologically
the most advanced of all three cells and has a high electrochemical performance.

Reactions in the diaphragm cell are as follows:

• Anode (positive electrode): carbon (graphite) or titanium coated with Ru-Ti oxide.

• Anode reaction (oxidation): 2Cl-(aq) → Cl2(g) + 2e-


• Cathode (negative electrode): steel mesh

• Cathode reaction (reduction): 2H2O(l) + 2e- → H2(g) + 2OH-(aq)

• Na+ migrates across diaphragm to cathode compartment combining with OH- to form
NaOH.

• 2OH-(aq) + 2Na+ → 2NaOH(aq)


17
• Overall cell reaction (showing Na+ spectator ions):

2H2O(l) + 2Cl-(aq) + 2Na+(aq) → 2Na+(aq) + 2OH-(aq) + H2(g) + Cl2(g)

The Membrane Cell Process Technique

General description

In this technique, the anode and cathode are separated by an ion-conducting membrane figure
6. Cell voltage is 2.35-4 V and current density 1.0-6.50A kA/m2. A 28 wt-% brine having
pH 2-4 flows through the anode compartment, where chloride ions are oxidized to chlorine
gas. The sodium ions, together with approximately 3.5 to 4.5 moles of water per mole of
sodium, migrate through the membrane to the cathode compartment, which contains a caustic
soda solution [5]. The water is electrolyzed at the cathode, releasing hydrogen gas and
hydroxide ions. The sodium and hydroxide ions combine to produce caustic soda, which is
typically kept at 32 ± 1 wt-% in the cell by diluting a part of the product stream with
demineralized water to a concentration of approximately 30 wt-percent and subsequent
recycling to the catholyte inlet. Caustic soda is continuously removed from the circuit.
Depleted brine is discharged from the anode compartment and resaturated with salt. The
membrane largely prevents the migration of chloride ions from the anode compartment to the
cathode compartment; therefore, the caustic soda solution produced contains little sodium
chloride (i.e. approximately 50 mg/l). Back-migration of hydroxide is also largely prevented
by the membrane but nevertheless takes place to a certain extent and increases the formation
of oxygen, hypochlorite and chlorate in the anode compartment, thereby resulting in a loss of
current efficiency of 3–7 % with respect to caustic soda production [5].

18
Figure 6. Process flow of the membrane cell process

Some electrolyzer produces more diluted 23–24 wt-% caustic soda. In this case, the caustic
entering the cell has a concentration of approximately 20–21 wt-% and the heat of the
electrolysis can be used to concentrate the 23–24 wt-% caustic solution to 32–34 wt-%. The
overall energy efficiency is comparable to the aforementioned process with the 32 wt-%
caustic solution but more equipment is required for the caustic evaporation. On the other hand,
simpler and cheaper construction materials can be used in the caustic circuit around the
membrane cells. Generally, the caustic produced in a concentration of 30–33 wt-percent is
concentrated to the usual commercial standard concentration of 50 wt-percent by evaporation
using steam. Another possibility is to use the caustic produced in the membrane cells as feed
to the decomposers of mercury cells.

The concentration of sodium chlorate in the produced caustic soda typically ranges from less
than 10 to 50 mg/kg. The level depends on the membrane characteristics, the operational
current density and the chlorate levels in the brine. The chlorine produced in membrane cells
contains low concentrations of oxygen (0.5–2.0 vol-%). The formation of oxygen and chlorate
ion can be depressed by selecting an anode coating with suitable characteristics and/or by
decreasing the pH in the anode compartment [5]. Brine depletion in membrane cells is two or

19
three times greater than in mercury cells, which allows the brine system to be smaller,
resulting in significantly lower recycling rates and less equipment needed compared to
mercury cell plants of the same capacity [5]. The membrane cell technique has the advantage
of producing a very pure caustic soda solution and of using less energy than the other
techniques. In addition, the membrane cell technique uses neither mercury, which is very
toxic, nor asbestos, which is classified as toxic (carcinogenic) [7]. The membrane cell process
brine specifications are more stringent than that of the mercury and diaphragm processes and
calls for impurities to be at the parts-per- billion (ppb) levels. If this level of purity is not
reached the membrane will be damaged. Disadvantages of the membrane cell technique are
that the caustic soda produced may need to be evaporated to increase its concentration and,
for some applications, the chlorine gas produced needs to be processed to remove oxygen,
usually by liquefaction and evaporation.
Furthermore, the brine entering a membrane cell must be of a very high purity, which requires
additional purification steps prior to electrolysis.

2.3.2The Cell
Various designs of membrane cells have been developed and used in commercial operations
figure 6 is a membrane cell diagram. The cathode material used in membrane cells is nickel.
Like the cathodes of diaphragm cells, they are often coated with a catalyst that is more stable
than the substrate and which increases surface area and reduces the over potential. Coating
materials include Ni-S, Ni-Al, and Ni-NiO mixtures, as well as mixtures of nickel and
platinum group metals. The cathode coatings for membrane cells have to be more chemically
resistant than those of diaphragm cells, because of the higher caustic concentration. The
anodes used consist of titanium coated with a mixture of ruthenium dioxide, titanium dioxide
and iridium dioxide.

The membranes used in the chlor-alkali industry are commonly made of per fluorinated
polymers. The membranes may have one to three layers, but generally consist of two layers
see the figure 8. One of these layers consists of a per fluorinated polymer with substituted
carboxylic groups and is adjacent to the cathodic side. The other layer consists of a per
fluorinated polymer with substituted sulphonic groups and is adjacent to the anodic side. The
carboxylate layer exhibits a high selectivity for the transport of sodium and potassium ions
and largely prevents the transport of hydroxide, chloride, hypochlorite, and chlorate ions,

20
while the sulphonate layer ensures good mechanical strength and a high electrical
conductivity. To give the membrane additional mechanical strength, it is generally reinforced
with polytetrafluoroethylene (PTFE) fibers. The membranes must remain stable while being
exposed to chlorine on one side and a strong caustic solution on the other. Commercially
available membranes are optimized for use in a specific strength of caustic. Depending on the
particular design, membrane sizes range from 0.2 to 5 m2.The general economic lifetime of
chlor-alkali membranes is approximately three to five years [5]. In the design of a membrane
cell, minimization of the voltage drop across the electrolyte is accomplished by bringing the
electrodes close together. However, when the gap is very small, the voltage increases because
of the entrapment of gas bubbles between the electrodes and the hydrophobic membrane. This
effect is avoided by coating both sides of the membrane with a thin layer of a porous inorganic
material to enhance the membrane's ability to release the gaseous products from its surface.
These improved membranes have allowed for the development of modern cells with zero-gap
or finite-gap cathode structures.

Figure 8. Schematic view of a membrane

Monopolar and bipolar electrolyzer


Electrolysis containing a multitude of membrane or diaphragm cells are classified as either
monopolar or bipolar. The designation does not refer to the electrochemical reactions that take
place, which of course require two poles or electrodes for all cells, but to the electrolyzer
construction or assembly. In a bipolar arrangement, the elements are connected in series with

21
a resultant low current and high voltage (Kirchhoff's circuit laws). The cathode of a cell is
connected directly to the anode of the adjacent cell as shown in figure 9. In the monopolar
arrangement, all anodes and cathodes are connected in parallel, forming an electrolyzer with
a high current and low voltage. The current has to be connected to every single anodic and
cathodic element, while in a bipolar electrolyzer the power supply is connected only to the
end part of the electrolyzer. Due to the long current path, ohmic losses in monopolar
electrolyzer are much higher than in equivalent bipolar electrolyzer, leading to increased
energy consumption.

Figure 9. Simplified scheme of monopolar and bipolar electrolyzer

Multiple electrolyzers are employed in a single direct current circuit as illustrated in figure 9.
Usually bipolar electrolyzers are connected in parallel with a low current and high voltage.
Monopolar electrolyzers are often connected in series, resulting in a high current circuit and
low voltage. Table 3 shows the differences between typical configurations of monopolar and
bipolar membrane cell plants with the same production capacity. Monopolar membrane cell
plants are characterized by a larger number of electrolyzer while the number of cells per
electrolyzer is lower than in a bipolar membrane cell plant.

Table 3. The differences between typical configurations of monopolar and bipolar

22
The maximum current density of an electrolyzer is determined by the resistance (ion
conductivity) of the membrane and the hydraulic conditions of the cells (elimination of the
gas formed). The first monopolar electrolyzer worked at maximum current densities of 4
kA/m2 but bipolar electrolyzer can now be operated at current densities of 6–7 kA/m2 and
pilot cells are currently tested at up to 10 kA/m2. The trend to develop higher current density
electrolyzer aims to reduce investment costs while developments in membranes, electrolysis
technology such as anode-cathode gap reduction and catalyst developments strive to reduce
energy consumption.

Monopolar membrane electrolyzer were only commercialized to maintain existing plants and
for new plants with small capacities. That change is due to the following advantages of bipolar
electrolyzer which lead to reduced investment and operating costs, as well as to improved
safety.

• Easier manufacturing;

•Smaller copper busbars due to the lower current; the only copper current distributors needed
are the main busbars connected to the end parts of the electrolyzer;

•Possibility to operate at higher current density without dramatic effects on energy


consumption due to very efficient internal recirculation;

•Membrane area more effectively used (from 85–87 % to 90–92 %);

23
•Better energy performance due to smaller voltage drop;

•No need for spare bipolar electrolyzer (only some spare individual cells are necessary,
compared to spare electrolyzer required for the monopolar technique);

• Shorter duration of shutdown and start-up phases to replace membranes due to the
easy and simple filter press design;

• Higher flexibility of operation (each electrolyzer could be operated independently of


the others due to a parallel connection); no need for expensive mobile short-circuit
switches for the isolation of troubled electrolyzer (Figure 10).

• New possibility to operate bipolar electrolyzer under slight pressure on the chlorine
side (no need for a blower).
• Easier detection of faulty cells by monitoring of individual cell voltages.

Figure 10. Electrolyzer architecture

2.3.4 Reactions in the membrane cell are as follows:

• Anode (positive electrode) : titanium


• Anode reaction (oxidation): 2Cl-(aq) → Cl2(g) + 2e-

Cathode (negative electrode): nickel

24
• Cathode reaction (reduction): 2H2O(l) + 2e- → H2(g) + 2OH-(aq)
• Na+ migrates across the membrane to cathode compartment combining with OH- to
form NaOH.

• Overall cell reaction (showing Na+ spectator ions):

2H2O(l) + 2Cl-(aq) + 2Na+(aq) → 2Na+(aq) + 2OH-(aq) + H2(g) + Cl2(g)

• Product is concentrated sodium hydroxide

The overall reaction for the diaphragm and membrane cells is:

The products of the electrolysis are formed in a fixed ratio, which is 1070–1128 kg of NaOH
(100 wt- %) and approximately 28 kg of H2 per ton of Cl2 produced. This product combination
is often referred to as the electrochemical unit (ECU).

Some side reactions occur during electrolysis, leading to a loss of efficiency [7]
At the anode, oxidation of water to oxygen and of hypochlorous acid to chlorate takes place:

2H2O → O2 + 4H+ + 4 e- or 4OH- → O2 + 2H2O + 4e- 12HClO


+ 6 H2O → 4ClO3- + 8 Cl- + 24 H+ + 3 O2 + 12 e-

Hypochlorous acid is formed by disproportionate of chlorine in water:


Cl2 + H2O HClO + H+ + Cl-
Chlorate is also produced by chemical reactions in the anolyte:
2HClO + ClO- → ClO3- + 2Cl- + 2H+
These four major side reactions are repressed by lowering the pH value.

The main typical characteristics of the different electrolysis technique are shown in table 4.

Table 4. Main typical characteristics of the different electrolysis techniques [7]

Criterion Mercury Diaphragm Membrane


Anode RuO2 + TiO2 coating RuO2 + TiO2 + SnO2 RuO2 + IrO2 + TiO2 coating
on Ti substrate coating on Ti substrate on Ti substrate

25
Cathode Mercury Steel (or steel coated Nickel coated with high area
with activated nickel) nickel-based or noble
metalbased coatings
Separator None Asbestos, Ion-exchange membrane
polymermodified
asbestos, or non-
asbestos diaphragm
Cell voltage 3.15–4.80 V 2.35–4.00 V 2.35–4.00 V
Current 2.2–14.5 kA/m2 0.8–2.7 kA/m2 1.0–6.5 kA/m2
density
Temperature Inlet: 50–75 °C Outlet: NI NI
80–90 °C
pH 2–5 2.5–3.5 2–4
Cathode Sodium amalgam (Na 10–12 wt-% NaOH 30–33 wt-% NaOH and
product HgX) and H2
Decomposer 50 wt-% NaOH and No decomposer needed No decomposer needed
product H2

Evaporator No evaporation needed 50 wt-% NaOH 50 wt-% NaOH


product
Quality of NaCl: ~ 50 mg/kg NaCl: ~ 10 000 mg/kg NaCl: ~ 50 mg/kg
caustic soda NaClO3: ~ 5 mg/kg (15 000–17 000 mg/kg NaClO3: ≤ 10–50 mg/kg
(50 wt-% Hg: ~ 0.1 mg/kg before concentration)
NaOH) NaClO3: ~
1 000
mg/kg
(400 –500 mg/kg
before concentration)
Chlorine O2: 0.1–0.3 vol-% O2: 0.5–2.0 vol-% O2: 0.5–2.0 vol-%
quality H2: 0.1–0.5 vol-% H2: 0.1–0.5 vol-% H2: 0.03–0.3 vol-%
N2: 0.2–0.5 vol-% N2: 1.0–3.0 vol-%
Advantages 50 wt-% high-purity Low quality Low total energy consumption,
caustic directly from requirements of brine, low
cell, high-purity low electrical energy investment and operating costs,
chlorine and hydrogen, no use of mercury or asbestos,
consumption high-purity caustic, further
simple brine
purification improvements expected

26
Disadvantages Use of mercury, High steam High-purity brine required,
expensive cell consumption for caustic low chlorine quality, high
operation, costly concentration cost of membranes
environmental in
protection, large floor expensive multi-effect
space evaporators, low-purity
caustic, low chlorine
quality, some cells are
operated with asbestos
diaphragms
NB: NI = No information provided

27
CHAPTER

CHLORINE, HYDROGEN AND CAUSTIC SODA


PROCESSING

Chlorine Processing, Storage and Handling

General description
Generally, for the production of dry gaseous or liquid chlorine most downstream process units
are required to get anhydrous chlorine with low hydrogen and oxygen content. The typical
chlorine gas is Cl2 97 - 99.5 vol%, O2 0.5 - 2.0 vol% and H2 0.03 - 0.3 vol%. To get the
typical chlorine specifications, it goes through a series of processes:

1. Cooling and filtration of hot wet chlorine gas.

2. Drying of “pre-dried” chlorine gas.

3. Scrubbing and cooling of dried chlorine gas, decomposition of NCl3.


4. Compression of dried chlorine gas.
5. Liquefying of compressed chlorine gas.
6. Vaporization of liquid chlorine.
7. Storage, transfer and filling.
In some applications, chlorine gas can be used as dry gas without the need for liquefaction.
Very occasionally it can be used directly from the electrolyzers. A general flow of chlorine
from the electrolyzers to storage is presented in figure 11. The chlorine process usually takes
hot, wet cell gas and converts it to a cold, dry gas. Chlorine gas leaving the electrolyzers has
a temperature of approximately 80–90 ºC and is saturated with water vapor. It also contains
brine mist, impurities such as N2, NCl3, H2, O2, CO2 and traces of chlorinated hydrocarbons.
Electrolyzers are operated at essentially atmospheric pressure with only a few mbar
differences between the pressure of the anolyte and the catholyte.

28
Figure 11. Chlorine processing

Material
The strong oxidizing nature of chlorine requires a careful choice of construction materials at
all stages of processing, depending on the operating conditions (temperature, pressure, state
of matter, moisture content). Most metals are resistant to dry chlorine at temperatures below
100 °C. Above a specific temperature for each metal, depending also on the particle size of
the metal, spontaneous ignition takes place (150–250 °C for iron). Carbon steel is the material
most commonly used for dry chlorine gas (water content below 20 ppmw). Wet chlorine gas
rapidly attacks most common metallic materials with the exception of tantalum and titanium,
the latter being the preferred choice in chlor-alkali plants. However, if the system does not
remain sufficiently wet, titanium ignites spontaneously (ignition temperature ~ 20 °C). Other
construction materials such as alloys, graphite, glass, porcelain and polymers may be used,

depending on the conditions. Oils or greases generally react with chlorine upon contact unless
they are fully halogenated [7].

Cooling
29
Chlorine cooling is generally designed for continuous cooling, filtration and, if required,
boosting of wet chlorine gas generated in the electrolysis cells. The purpose is to lower the
water content to the lowest possible value by means of cooling and chilling, to separate and
remove brine mist (aerosol) which is carried over from the electrolysis cells and optionally to
increase the gas pressure as may require for the downstream plant sections.

·Modern Concept
The modern concept of chlorine cooling see figure 12, is designed for:
·Indirect cooling in shell and tube heat exchangers.
·High efficiency filtration of the gas by use of glass fiber elements to separate the entrained
brine mist. With these measures the dew point of the gas could be brought to the absolute
minimum and the brine mist separation to the maximum.

The modern system is described below:


Chlorine Cooling
Cooling of the wet chlorine gas is normally done in two stages.
· First stage cooling from 95 to 40 °C by means of cooling water.
· Second stage cooling from 40 to 15 °C by means of chilled water.
·Indirect cooling through a titanium surface (usually in a single-pass vertical shell-and-tube
heat exchanger). The resultant condensate is either fed back into the brine system of the
mercury or membrane cell technique or is dechlorinated by evaporation in the case of the
diaphragm cell technique. Indirect cooling lowers the water content to approx. 4500 ppm. Gas
temperature lower than 13 °C is not recommended due to the formation of chlorinehydrate
(white crystals; Cl2·nH2O; n = 7–8), which may plug of pipelines or filter elements. The final
temperature of the cooled gas is also subject to the temperature increase in the blower and the
remaining water content at this temperature to avoid the ignition of a Ti-Cl2 fire. Below 13°C
the water vapour pressure in the gas becomes too low to be safe. Under conditions of chlorine-
hydrate formation ignition is still possible. All parts that are coming into contact with the
chlorine gas are made of titanium; these are the tube sheet as well as the tubes, which are
seamless or welded. The shell is made of normal carbon steel suitable for cooling water. The
heat exchangers are installed either vertically or horizontally with a slope of approx. 3° in
direction of the flow. The temperature split at 40 °C results in 90% heat removal in the first
stage and 10% in the second stage and similar heat transfer area for both coolers. Condensate

30
from both coolers is chlorine saturated and is recycled to depleted brine. There is no need to
provide standby heat exchangers. Also cleaning of the tubes is normally not necessary. The
advantages of this concept are:

·No possibility of NCl3 formation


·Low space requirement
·Very low maintenance requirement
·Lowest possible gas temperature and water vapour content in the gas, which reduces acid
consumption for chlorine drying.

·Low operating cost.


·Less chlorine to be condensed or absorbed ·Less
chlorine-saturated water for disposal.

·Indirect cooling can be carried out in once-through, open-recirculating, or closed-loop


systems.

·Another method is direct cooling with water (or brine or other fluids). The chlorine gas is
cooled by passing it directly into the bottom of a tower. Water is sprayed from the top and
flows counter currently to the chlorine. The cooling water is generally free of traces of
ammonium salts, to avoid the formation of nitrogen trichloride. This method has the
advantage of better mass transfer characteristics and higher thermal efficiencies. Direct
cooling is usually carried out in closed-loop systems.

Figure 12. Modern Concept of Chlorine Cooling

Filtration and boosting


31
Filtration
Cooled and water vapour saturated chlorine gas contains remaining brine mist which must be
separated from the gas to avoid formation of sodium-sulfate during chlorine drying with
sulfuric acid. The gas is fed horizontally through a chlorine resistant glass fiber bed. Particles
are contacted and collected on individual fibers of the bed and then coalesce to form liquid
films and droplets, which are pushed through the bed by the gas flow. The collected liquid
then is drained off the downstream face of the bed by gravity. The particles are collected in
three different ways: by internal agglomeration (mainly for larger particles > 3 micron) by
direct collision for particle sizes between 1and 3 microns and by Brownian diffusion for
particles smaller than 1 micron. The filter-elements are designed for an overall separation
efficiency of 98% and a pressure drop of max. 150 mm WC (water column). The brine mist
quantity in the feed is normally less than 350 ppm NaCl. To achieve the high separation
efficiency the gas flow rate should be kept at design conditions.

The filter elements are installed vertically and fixed in a filter plate. Gas is entering the filter
vessel below the lowest point of the elements, flowing up and then horizontally through the
elements. Inside the element the filtered gas leaves the vessel at the top. The constant
saturation of the gas with water vapour is important. Therefore, the filter is installed upstream
of the blower (if any) and close to the second stage chlorine gas cooler. To achieve saturation
of overheated gas, water addition via a spray nozzle located at gas inlet nozzle of the filter is
provided. Water addition is normally not necessary. If required it is added discontinuously.
The water collected from the bottom of the filter vessel is drawnoff and added to the
condensate from the chlorine coolers. The filter elements are normally self-cleaning and do
not require maintenance. A replacement is required after 5-6 years of operation.

The advantages of this concept are:


· Very high separation efficiency for brine mist.
· Very low maintenance and operating cost.
· High reliability.

Boosting
If required, the cooled and filtered chlorine gas is boosted by means of centrifugal blowers to
a pressure of up to 2000 mm WC. Boosting the entire or part of the gas stream is only required

32
if wet chlorine gas is branched off for the production of hydrochloric acid or ironchloride.
Alternatively the gas stream could be branched-off after the chlorine gas compression. The
blower is made of carbon steel with hard rubber lining and titanium impeller. Normally is the
shaft sealed with a double acting mechanical seal, flushed and cooled with water. The
mechanical seal is normally very sensitive to water failures, vibration etc. and special
protections like water flow- and pressure control is required. The improvement consists in
using a labyrinth sealing with gas flushing. To avoid over heating of the boosted gas and
therefore ignition of a Ti-Cl2 fire at the blower impeller, recirculating part of the gas stream
back to the inlet of the first chlorine gas cooler must control the capacity [4].

Drying
Chlorine from the cooling system is more or less saturated with water vapour. The water
content is typically 1–3 vol-%. This must be reduced in order to avoid downstream corrosion
and to minimize the formation of hydrates.

The drying of chlorine is carried out almost exclusively with concentrated sulphuric acid (96–
98 wt-%) in countercurrent contact towers in two to six stages, which reduce the moisture
content to less than 20 mg/m3 as shown in figure 13. The remaining moisture content depends
on the temperature and concentration of the sulphuric acid in the last drying stage. For low
temperature liquefaction, a lower moisture content is required, which can be achieved by
adding more equilibrium stages to the drying towers or by using molecular sieves to levels of
3–9 mg/m3 [7]. The number of stages is usually increased to lower the final strength of the
spent sulphuric acid. For example, three stages are needed to reach a spent acid concentration
of 50–65 wt-percent while six stages are needed for a final concentration of 30–40 wt-percent.
The columns contain plastic packing resistant to chlorine and sulphuric acid to improve fluids
distribution, increase efficiency and lower pressure drops, and thus reduce energy
consumption. The heat liberated during dilution of the circulating acid is removed by titanium
heat exchangers, and the spent acid is dechlorinated chemically or by stripping. The
concentration of the spent acid depends on the number of drying stages and the further
potential use or method of disposal. In some cases, the acid is reconcentrated to 96 wt-% by
heating it under vacuum and then it is subsequently recirculated. Sometimes the acid is sold
or used for other purposes. Rarely, it becomes waste.

33
Figure 13. Chlorine drying process

Cleaning of dry chlorine


When leaving the top of the drying tower, dry chlorine passes through high efficiency
demisters or a packed bed to prevent the entrainment of sulphuric acid droplets.

Further potential cleaning steps after chlorine drying include:


•adsorption on carbon beds to remove organic impurities;
• scrubbing with concentrated hydrochloric acid to remove nitrogen trichloride;
•scrubbing with liquid chlorine to remove nitrogen trichloride, organic impurities, carbon
dioxide and bromine;

•irradiation with UV destroys nitrogen trichloride and hydrogen.

Compression
After drying and potential further cleaning, chlorine gas may be compressed by a variety of
compressors, depending on the throughput and the desired pressure [7], Rotary compressors,
such as:

•Sulphuric acid liquid ring compressors for throughputs of 150 t/d per compressor and for
pressures of 4 bar or, in two-stage compressors, 12 bar;
34
•Screw compressors for low throughputs and for pressures of up to 16 bar.
Reciprocating compressors, such as:
•Dry ring compressors for throughputs of 200 t/d per compressor and for pressures of up to
16 bar.

Centrifugal compressors, such as:


•Turbo compressors in mono- or multi-stage operation for throughputs of up to ~ 1800 t/d per
unit and for pressures of up to 16 bar;

•Sundyne blowers for throughputs of 80–250 t/d per compressor and for pressures of up to 3
bar.

Because of heat build-up from compression, multi-stage units with coolers between stages are
usually necessary. Compressor seals are generally fitted with a pressurized purge to inhibit
the leakage of chlorine to the atmosphere. Dry chlorine at high temperatures can react
spontaneously and uncontrollably with iron. Chlorine temperatures are therefore usually kept
below 120 °C.

Liquefaction
Liquefaction can be accomplished at different pressure and temperature levels: at ambient
temperature and high pressure (for example 18 ºC and 7–12 bar), at low temperature and low
pressure (for example -35 ºC and 1 bar) or any other intermediate combination of temperature
and pressure. Important factors for selecting appropriate liquefaction conditions include the
composition of the chlorine gas, the desired purity of the liquid chlorine and the desired yield.
Increasing the liquefaction pressure increases the energy consumption of compression,
although the necessary energy for cooling decreases, resulting in an overall reduction in
energy consumption. The liquefaction yield is typically limited to 90–95 % in a single-stage
installation, as hydrogen is concentrated in the residual gas and its concentration needs to be
kept below the lower explosive limit. Higher yields of up to 99.8 % can be achieved by multi-
stage liquefaction. Typically, small volume liquefiers which are protected against explosions
are used after primary liquefaction, and inert gas is added to keep the mixture below the lower
explosive limit [7]. Another possibility is to remove hydrogen from the system by reaction
with chlorine gas in a column, yielding hydrogen chloride which can be recovered in a
hydrochloric acid unit. The remaining chlorine gas can then be safely further condensed. This

35
solution can be chosen if hydrochloric acid is a saleable product or if it can be used as a
feedstock for downstream production, such as for ferric chloride.

The choice of the refrigerant in a certain stage of the liquefaction depends on the pressure of
the chlorine. When the pressure is sufficiently high, water can be used as an indirect
refrigerant. When the pressure is relatively low, other refrigerants such as
hydrochlorofluorocarbons (HCFCs) or hydrofluorocarbons (HFCs), typically
chlorodifluoromethane (HCFC-22) and 1,1,1,2-tetrafluoroethane (HFC-134a) (indirect
cooling), ammonia (indirect cooling) or liquid chlorine (direct cooling) are used.

The use of HCFCs such as HCFC-22 is generally prohibited but reclaimed or recycled HCFCs
may be used for the maintenance or servicing of existing refrigeration equipment until 31
December 2014.

The temperature of the chlorine gas in a certain stage depends mainly on the pressure after
compression. A pressure greater than 8 bar generally enables water cooling but implies an
increased hazard.

The residual chlorine in the tail gas can be used to produce hypochlorite, iron (III) chloride or
hydrochloric acid. The residual chlorine which cannot be valorized is then led to the chlorine
absorption unit. In some cases, it is recovered by absorption-desorption process with carbon
tetrachloride which is strictly prohibited. The latter has the disadvantage of using a toxic
substance with a high ozone depletion and global warming potential.

Handling and storage


Liquefied chlorine is stored at ambient or low temperature. The pressure corresponds to the
vapour pressure of the liquefied chlorine at the temperature in the storage tank. Pressure
storage at ambient temperatures (~ 7 bar at 20 °C) has advantages of simplicity of operation,
ease of visual external inspections, as well as lower energy and investment costs. Low-
pressure storage operating around the boiling point of liquid chlorine (-34 °C) requires more
complex infrastructure, particular safety measures and higher energy costs. Within a plant and
over distances of several kilometers, chlorine can be transported by pipelines, either as a gas
or a liquid. The liquid chlorine from the bulk tank can be used as a feedstock for on-site
processes or loaded into containers, or road or rail tankers.

Vaporization
36
Liquid chlorine is usually vaporized prior to use. The easiest option is to use ambient heat by
which approximately 5 kg of chlorine per hour and square meter of container surface can be
vaporized. For higher flowrates, it is necessary to use a chlorine vaporizer.

Hydrogen Processing, Storage and Handling


Hydrogen leaving the cells is highly concentrated (> 99.9 vol-%) and normally cooled to
remove water vapour, sodium hydroxide and salt. The solution of condensed saltwater and
sodium hydroxide is recycled to produce caustic, as brine make-up or is treated with other
waste water streams.

In the case of the membrane or diaphragm cell technique, the cooling is usually carried out by
one or more large heat exchangers. In the mercury cell technique, primary cooling is carried
out at the electrolyzer, allowing mercury vapour to condense into the main mercury circuit.
Further cooling and mercury removal takes place at a later stage using a variety of techniques.
Some uses of hydrogen require additional removal of traces of oxygen, which may be
achieved by react the oxygen with some of the hydrogen over a platinum catalyst. Hydrogen
may be distributed to users using booster fans or fed to the main compression plant, which
usually comprises a number of compressors and a gas holder (surge chamber). The hydrogen
gas holder is incorporated into the system to minimize fluctuations in gas pressure from the
primary stage. The hydrogen product gas stream is always kept pressurized to avoid the
ingress of air. All electrical equipment taken into the hydrogen compression plant area must
be

'intrinsically safe' (i.e. the equipment will not produce a spark) or explosion proof (i.e. a local
small explosion is contained within the equipment). A relief valve is normally provided within
the system to relieve high pressure to atmosphere. Hydrogen is normally monitored for
oxygen content, and the compression will shut down automatically in critical situations. The
hydrogen sold to distributors is usually compressed at pressures higher than 100 bar and is
injected into a pipeline network. Otherwise, the hydrogen is transported in dedicated tank
lorries or in steel bottles at pressures of up to 300 bar. For these high pressures, the gas is
further dried and traces of oxygen are usually removed.

37
The main utilizations of the co-produced hydrogen are combustion to produce steam (and
some electricity) and chemical reactions such as the production of ammonia, hydrogen
peroxide, hydrochloric acid and methanol.

Caustic Soda Processing


The caustic soda solution from the three techniques is treated in slightly different ways due to
the difference in composition and concentration. In the case of the mercury cell technique, 50
wt-percent caustic soda is obtained directly from the decomposers. The caustic soda is
normally pumped through a cooler, then through a mercury removal system and then to the
intermediate and final storage sections. In some cases, the caustic is heated before filtration.
The most common method for the removal of mercury from caustic soda is a plate (or leaf)
filter with a carbon pre-coat, under normal operating conditions, mercury cell caustic soda (as
100 % NaOH) contains 20–100 ppmw of sodium chloride and 40–60 μg Hg/kg NaOH.

In the case of the diaphragm and membrane cell techniques, the caustic soda is usually
concentrated to 50 wt-percent by evaporation before storage. Steam is used as the source of
evaporative energy.

In the case of the diaphragm cell technique, this is achieved by triple- or quadruple-effect
evaporators. Increasing the number of effects reduces energy consumption and operating costs
but increases investment costs. The presence of salt in the diaphragm cell liquor requires that
the evaporator be equipped with scraper blades or other devices to draw off the precipitated
salt, which is then usually reused for brine preparation. Sodium sulphate present in the cell
liquor (0.12–0.65 wt-%) also crystallizes in the later stages of evaporation and may be isolated
to avoid contamination of the main portion of the recovered salt. The residual level of sodium
chloride in sodium hydroxide from diaphragm cells is approximately 1 wt-% and sodium
chlorate approximately 0.1 wt-percent. For this reason, it is unsuitable for certain end
applications such as the manufacture of rayon. The concentrations of salt and sodium chlorate
in the caustic soda from diaphragm cells can be reduced by extraction with anhydrous liquid
ammonia to increase marketability, but at an increased cost.
In the case of the membrane cell technique, concentration of the caustic soda is normally
achieved in two or three stages using either plate or shell-and-tube evaporators. The number
of stages depends on factors such as plant size and the cost of steam. The caustic soda from
membrane cells is of high quality, although the caustic soda produced (usually around32 wt-
38
% NaOH) needs to be concentrated to 50 wt-% NaOH to be traded as a commodity. The NaCl
content of the membrane-cell caustic soda lies between 20 and 100 ppmw(in 100 % NaOH)
but is on average slightly higher than mercury cell caustic. In some plants, the caustic soda is
further concentrated to a 73 wt-% solution and to solid caustic pills or flakes with a water
content of < 0.5–1.5 wt-%, using multi-effect evaporators. Some chlor-alkali production
facilities combine the caustic production process from mercury and membrane cells to
minimize energy costs. Because of its highly reactive and corrosive properties, caustic soda
may corrode containers and handling equipment. Construction materials must be suited to the
caustic soda handled and stored.

Caustic solutions require steam or electrical heating where temperatures can fall below the
freezing point. Depending on the concentration, the freezing point can be higher than 0 °C;
for example it is 5 °C for 32 wt-percent NaOH and 12 °C for 50 wt-percent NaOH. Frozen
pipelines present both safety and environmental risks when attempts are made to unblock
them.

Storage tanks may be lined in order to minimize iron contamination of the product or to
prevent stress corrosion cracking of the tank. Tanks are usually included in procedures to
prevent over flow or spillage of caustic soda. Such procedures include containment and
mitigation. Dissolved hydrogen gas can be released into the vapour space above the liquid in
storage tanks. Tanks are normally vented from the highest point. Testing for an explosive
mixture of hydrogen and air normally precedes any maintenance activity in the area.

39
CHAPTER
CHLOR-ALKALI MATERIAL AND ENERGY BALANCE

Material quantities, as they pass through processing operations, can be described by material
balances. Such balances are statements on the conservation of mass. Similarly, energy
quantities can be described by energy balances, which are statements on the conservation of
energy.

If the unit operation, whatever its nature is seen as a whole it may be represented
diagrammatically as a box, as shown in Figure. 14. The mass and energy going into the
process must balance with the mass and energy coming out the process.

Figure 14: Mass and Energy Balance

4.1- Chlor-Alkali Material Balances


Table 5. Typical data for chlor-alkali cells

40
Mercury Diaphragm Membrane
Cell Cell Cell
Cell voltage/V -4.4 -3.45 -2.9
Current density KA/m2 8-13 0.9-2.6 3-5
Current efficiency for Cl2/% 97 96 98.5
Energy consumption KWh/MT Cl2 3360 2720 2650
Electricity Utilities 200 250 140

Steam Requirement for Heating Brine


Q = mCPΔT
Q = quantity of energy kJ
m = mass flow rate of Aq NaCl in kg/hr CP= specific
heat capacity of Aq NaCl at 60O C

ΔT = Temperature Difference λ = latent


heat of vaporization in kJ/kg

CHAPTER
USES OF PRODUCTS
Important Uses of Chlorine
• Chlorine is largely used in the synthesis of chlorinated organic compounds. Vinyl
chloride monomer (VCM) for the synthesis of polyvinyl chloride (PVC) still remains
the driver of chloro-alkali production in the world.

• Manufacture hydrochloric acid by the direct reaction with hydrogen gas


H2O + Cl2 ==> HCl + HClO
• Synthesize a wide range of organic compounds such as Methylene Chloride,
(CH2Cl2), Chloroform, (CHCl3), Carbon Tetrachloride, CCl4.

• Extract Gold from its ores

• Chlorine's principal applications are in the production of a wide range of industrial


and consumer products, it is used in making plastics, solvents for dry cleaning and
41
metal degreasing, textiles, agrochemicals and pharmaceuticals, insecticides,
dyestuffs, etc.

• Chlorine is an important chemical for water purification (such as water treatment


plants), in disinfectants.

• Chlorine is usually used (in the form of hypochlorous acid) to kill bacteria and other
microbes in drinking water supplies and public swimming pools.

• Bleaching powder industry.


Important Uses of Hydrogen
• A perfect fuel.
– The byproducts of hydrogen combustion are electricity, water and heat.
– Energy conversion devices using hydrogen are highly efficient and
produce very little or no harmful emissions; safe and harmless fuel.

• It is used as a shielding gas in welding methods such as atomic hydrogen welding.


• H2 is used as the rotor coolant in electrical generators at power stations, because it has
the highest thermal conductivity of any gas

• In more recent applications, hydrogen is used pure or mixed with nitrogen


(sometimes called forming gas) as a tracer gas for minute leak detection. Applications
can be found in the automotive, chemical, power generation, aerospace, and
telecommunications industries

• Hydrogen is used for catalytic hydrogenation of edible vegetable oils.


Important Uses of Sodium Hydroxide
• General applications
Sodium hydroxide is the principal strong base used in the chemical industry. In bulk it is most
often handled as an aqueous solution, since solutions are cheaper and easier to handle. It is
used to drive chemical reactions and also for the neutralization of acidic materials. It can be
used also as a neutralizing agent in petroleum refining. It is sometimes used as a cleaner.

• Soap production

42
Sodium hydroxide was traditionally used in soap making (cold process soap, saponification).
Persians and Arabs began producing soap in this way in the 7th century, and the same basic
process is used today.

• Biodiesel
For the manufacture of biodiesel, sodium hydroxide is used as a catalyst for the
transesterification of methanol and triglycerides. This only works with anhydrous sodium
hydroxide, because combined with water the fat would turn into soap, which would be tainted
with methanol. It is used more often than potassium hydroxide because it is cheaper, and a
smaller quantity is needed

43
CHAPTER

ENVIRONMENTAND SAFETY MEASURES

Environmental Relevance of the Chlor-Alkali industry

Inputs and pollutant outputs from the chlor-alkali industry are quite specific to the cell
technology used, the purity of the incoming salt and the specifications of the products.
Because of the huge amount of electricity needed in the process, energy can be considered as
a raw material. The chlor-alkali process is one of the largest consumers of electrical energy.

Historical mercury and PCDD/Fs contamination of land and waterways from mercury and
diaphragm chlor-alkali plants is a big environmental problem at some sites.

For many years, the mercury cell has been a significant source of environmental pollution
because some mercury is lost from the process to air, water, products and wastes. Inorganic
mercury can be metabolized to form highly toxic methyl mercury by anaerobic bacteria, and
this organic mercury is bio-accumulated1 in the food chain. It is recognized that the main part
of the mercury losses is in the different wastes from the process. Considerable emissions of
mercury can also occur with run-off water. The soil at many sites is contaminated with
mercury due to deposition of diffuse emissions and/or historical disposal of mercury
contaminated wastes. The mercury leaks from the soil and ends up in the runoff water.
Another big entry is the ‘Difference to Balance’. The annual mercury balance for a site is
never zero. This is because mercury accumulates in plant equipment and structures during the
life of the plant.
The chlor-alkali industry was the largest domestic user of mercury in the years 1989 to 1990
in the world.

The European Atmospheric Emission Inventory of Heavy Metals and Persistent Organic
Pollutants in the 15 EU countries, the highest emitters of mercury to air in 1990 include: coal
burning electric utilities (highest figure of 90.5 ton), municipal and hazardous waste
incinerators and cement industry (37.7 ton). Chlor-alkali emissions to air are reported to be

44
28.4 tones. A USEPA report, dated 1998, identifies the same sources of emissions: municipal
waste incinerators, medical waste incinerators, hazardous waste incinerators and industrial
boilers.

In addition, certain manufacturing processes, most notably chlor-alkali and cement plants, are
listed although their emissions are substantially lower than those of incineration sources.

According to Euro Chlor, the total mercury emission to air, water and products from
chloralkali plants in Western Europe was 9.5 ton in 1998, ranging between 0.2-3.0 g Hg/ton
of chlorine capacity at the individual plants. Decision 90/3 of 14 June 1990 of the Commission
for the Protection of the Marine Environment of the North-East Atlantic (PARCOM2)
recommends that existing mercury cell chlor-alkali plants should be phased out as soon as
practicable. The objective is that they should be phased out completely by 2010. As regards
the diaphragm technology, due to the potential exposure of employees to asbestos

and releases to the environment, the use of good practices is needed and some efforts are being
made to replace the asbestos with other diaphragm material. With the inputs/outputs of the
chlor-alkali sector, it is also relevant to point out the special importance of safety aspects
related to production, handling and storage of chlorine.

Safety Measures
In hazard and risk assessment studies, the design of chlor-alkali installations and equipment
and the operating and maintenance routines are examined in detail to reduce the risks for
people and the environment at source as much as possible. The most important substance to
consider is chlorine. Preventative measures are the most important, although corrective and
emergency measures are also of importance. The design principles of the plant, scheduled
maintenance and inspection, procedures and instrumentation (control systems) for operating
the unit are covered, as well as hardware. An overview is given, below, of measures that can
be applied to reduce the risks in operating a chlor-alkali plant, including storage and loading
of products. The measures presented are an indication of possible measures that can be taken.
The local situation will determine the final package of measures that is required.

General measures

45
General measures related to safety are mainly linked to reliable and efficient safety
management systems based on: Training of personnel including:

- Basic knowledge of chlorine properties


- Correct operating practice
- Emergency procedures
- Frequent refresher training
- Making sure that contractors’ personnel on the site are familiar with the safety regulations
and procedures for the site.

Identification and evaluation of major hazards


- written information for personnel about safety measures in normal and abnormal conditions

Instructions for safe operation including:


- Permanent monitoring of the installations under the responsibility of a designated person
specially trained for chlorine hazards

- Good compliance with safety parameters defined in the safety report, including periodic
Inspection and control of materials specified according to safety hazards
- Maintenance programs for the installations: for example, storage, maintenance of pipework,
pumps, compressors, monitoring of moisture concentrations, impurities in liquid chlorine

Planning for emergencies and recording of accidents and near-misses


- Preparation, test and review of emergency plans
The safety management systems should be completed with appropriate technical measures,
such as:

- High quality preventive and protective systems, in particular in the loading area
- Enhanced leak detection and leak isolation
- Good protection of employees and temporary workers on the site with appropriate and well-
maintained equipment.

In the chlor-alkali plant


In the chlor-alkali plant, safety measures refer in particular to the prevention of liquid chlorine
Spillage, using means such as:

- overfilling protection systems

46
- correct choice of construction material and regular inspection of vessels
- bonding for vessels containing hazardous material and also:

- Preventing impurities to avoid any explosive mixture. Measuring and control of hydrogen
concentration in the chlorine gas from the cell room and after each liquefaction step and
measuring and prevention of possible NCl3 accumulation.

- Prevention of failure of the electrical supply. Emergency generators for supply of power to
vital equipment when/if grid power fails and also the prevention of failure of the instrument
air supply.

- Prevention of chlorine releases by a collection of chlorine releases during maintenance


operation to the absorption unit, a correct warning for process deviations and irregularities.
In a more general way, a good layout of the installation and provision for instant shut-down
of some compartments help to prevent domino-effects. Manual emergency push buttons
which can be used by any of the site personnel on discovery of a chlorine leak should be
present around the plant.

In the loading area


In any liquid chlorine loading (or unloading) system, the weakest link is the connection
between the static plant and the mobile tank. The use of pneumatic valves with automatic shut
down in case of failure of the link, at both ends of the link, is essential to limit the leak.
Only a risk assessment study, proper to each installation, can give the most likely modes of
failure during loading and the associated safety measures to avoid them. Examples of some
standard measures are:

- Improved chlorine detection and location and rapid isolation of sources feeding a leak
- Connection of loading area to a chlorine absorption system
- PTFE hoses should not be used (based on information reported to accident databases)
- Articulated arms or correctly specified flexible hoses and coils for chlorine transfers
In the storage area
Some standard safety measures to achieve optimum risk reduction in the storage area are:
-Chlorine detection and location and rapid isolation of sources feeding a leak
-Availability of at least one empty tank of sufficient capacity as an emergency spare

47
-Good pipe work design to minimize the length of pipeline containing liquid chlorine -
Limitation of the overall liquid chlorine inventory to what is really needed. A simplified
layout and a reduction in the number of valves, pipes and connections reduce the risk of
leakage

- For large storage capacities, low-temperature storage at -34 °C is recommended.

48
REFERENCE

[1] Clark J. 2007. Manufacturing chlorine using a diaphragm and a membrane cells..

[2] Pamphlets provided by Arthur E. Dungan of The Chlorine Institute, Inc., Washington,
D.C. January 1991.

[3] 1991 Directory of Chemical Producers: United States of America. Menlo Park, California:
Chemical Information Services, Stanford Research Institute, 1991.

[4] O'Brien T, Bommaraju TV, Hine F. 2005 a Handbook of chlor-alkali technology: Brine
treatment and cell operation. Plenum Publishing Corporation.

[5] Ullmann's, 'Chlorine', Ullmann's Encyclopedia of Industrial Chemistry, 7th edition,


electronic release, 2006.

[6] SRI Consulting, Chemical Economics Handbook - Chlorine/Sodium Hydroxide, SRI


Consulting, 2008.

[7] Thomas Brinkmann, Germán Giner Santonja, Frauke Schorcht, Serge Roudier and
Luis Delgado Sancho, 2014, Best Available Techniques (BAT), Reference Document for the
Production of Chlor-alkali. Industrial Emissions Directive 2010/75/EU. Integrated Pollution
Prevention and control. [www document], http://eippcb.jrc.ec.europa.eu
https://ec.europa.eu/jrc/en/institutes/ipts

[8] Industrial Project on Production of Sodium Hydroxide from sodium chloride by membrane
cell technology process [J.S.M. Mahedi 2016] Department of Chemical Engineering and
Polymer Science Shahjalal University of Science & Technology Sylhet3114, Bangladesh
[2/2018 www document]https://www.slideshare.net/MahediJsm/projectpaper-on-
chloralkali-process.

[9] Vladimir M. Sedivy, 2008, Economy of Salt in Chloralkali Manufacture, National Salt
Conference 2008, Gandhidham. Email: vladimir.m.sedivy@salt-partners.com.

49
[10] An Alternative to Purchasing Chlor-Alkali Commodity Chemicals, 2018 UniChlor
Technology. Powell Fabrication & Manufacturing, Inc. websites: www.powellfab.com,
www.valveclosures.com, email: info@powellfab.com [5/2018 www document.

50

You might also like