Sie sind auf Seite 1von 56

SHORT COURSE

WATER PURIFICATION AND TREATMENT

Everpure, Inc.
660 North Blackhawk Drive
Westmont, IL 60559-9005
Phone: 630-654-4000
Fax: 630-654-1115
E-mail: info@everpure.com

Prepared by
William H. Beauman
Everpure, Inc. 1998
Updated January, 2001

EVERPURE, QC4 and MICRO-PURE


are registered trademarks of Everpure, Inc.

TABLE OF CONTENTS
page
SECTION I: INTRODUCTION 5
A. General Information About Water5
1. The Water Cycle ...........................................................................5
2. The Basics of Water Chemistry.....................................................6
a) Dissolved, Particulate, or Colloidal .........................................6
i) Turbidity...........................................................................7
ii) Discussion of Sizes and dimensions...............................7
b) Organic or Inorganic ...............................................................8
c) Biological or Mineral ...............................................................9
d) Ionized or not..........................................................................9
i) Acids, Bases, and Salts ...................................................9
e) Aesthetic, Health-related or Nuisance ..................................10
f) Adsorption .............................................................................10
I) Carbon Adsorption of Chlorine ......................................11
ii) Activated Carbon ..........................................................12
iii) Freundlich Carbon Isotherms.......................................14
g) Chemical Equilibrium.............................................................15
i) The Water Equilibrium ...................................................16
ii) The Carbonic Acid Equilibrium......................................16
iii) About Buffers ...............................................................17
h) Oxidation and Reduction .......................................................18
i) Oxidation-Reduction Potential ........................................19
ii) Electromotive Series ......................................................21
B. Drinking Water Quality Standards ......................................................21
C. Centrally Treated Water Supplies22
D. Individual Water Supplies...................................................................22
SECTION II: DESCRIPTION OF WATER PROBLEMS................................23
A. Microbiological Problems..................................................................23
1. Viruses .........................................................................................23
2. Bacteria ........................................................................................23
i) Silver Filters ............................................................................24
ii) Bacteriological Testing..........................................................25
3. Algae ...........................................................................................26
4. Fungi (molds) ...............................................................................26
5. Protozoa and larger parasites ......................................................26
B. Iron Water, including Manganese27
C. Sulfur Water28
D. Acid Water28
E. Alkaline Water and Excessive Alkalinity29
F. Hard Water and Scale............................................................30
i) The Langelier Index.......................................................................31

ii) Polyphosphates and Threshold Treatment32


G. Brackish Water and Excessive Total Dissolved Solids ....................34
i) Osmotic Pressure............34
H. Turbidity 35
I. Taste & Odor ..................................................................................36
i) Chlorine T&O................................................................................36
ii) Earthy-musty-moldy-mildewy-fishy T&O......................................37
iii) Chlorophenols38
J. Color
...........................................................................................39
K. Toxic Organic Contamination40
SECTION III: TYPES OF WATER TREATMENT AND THE EQUIPMENT
USED
.................................................................. A.
Mechanical Filtration41
1. Pressure Drop .............................................................................42
2. Hydraulic Capacity.......................................................................42
3. Everpure Precoat Filters..............................................................43
4. Carbon Blocks .............................................................................45
B. Adsorption46
C. Ion Exchange46
1. Softening (Conditioning)...............................................................47
2. Dealkalization ...............................................................................48
3. Demineralization...........................................................................49
D. Oxidation Filtration..............................................................................50
1. Manganese Greensand...............................................................50
2. Granular Brass51
E. Chlorination, Coagulation, and pH adjustment .................................52
F. Membrane Systems ...........................................................................56
1. Ultrafiltration .................................................................................56
2. Nanofiltration ................................................................................56
3. Reverse Osmosis .........................................................................57
G. Ultra-Violet Irradiation........................................................................58

A note on typefaces:
Underlining shows emphasis.
Italics identifies foreign terms, including the Latin names of organisms.
Bold-face sets apart important terms or industry jargon to remember.
SECTION I

INTRODUCTION
Congratulations, and welcome! Your initiative in enlisting in this Short Course on
potable water treatment should be rewarded with sufficient expertise in the field
to understand most water problems and their solutions and to design effective
treatment systems. There will always be much more to learn than can be given
here, but this course should provide the minimum background for a water
treatment professional.
A. General Information About Water
Water is arguably the single most important substance in our lives. It has been
called the universal solvent, and it is the fundamental requirement of all life.
Virtually all of the water in existence today has been on Earth for billions of
years; any new water comes only from comets. All the rest is continually being
recycled. More than 80% of the Earths surface is covered by water, but less
than 1% of it is drinkable, non-saline water. Water comprises about 2/3 of our
bodies. Most of us could go without food for weeks, but we would die of thirst
after only a couple of days. People in the developed nations each use some 123
gallons (466 L) of water every day, and we each drink an average of about a liter
a day. The U.S. EPA bases regulatory matters on the assumed consumption of
two liters per day per person.
The Water Cycle
1. The water cycle begins with water vapormoisture in the air from evaporation
of water from the surfaces of oceans, lakes, etc.; transpiration from ground water
through plants into the air from their leaves; and water vapor belched up from
volcanoes. The vapor condenses under the influence of reduced temperature to
form clouds, and then condenses further onto the surface of particles to form
precipitation. Rain washes much of the air pollution out of the air, making rain a
significant cause of water pollution. Acidic gases from combustion sources
cause acid precipitation (acid rain), but even without any man-made acids in
the air, normal levels of carbon dioxide (about 0.04%) dissolve in the rain to
produce carbonic acid, H2CO3.
2. This slightly acidic rainwater lands on the ground and percolates through the
soil, where it begins to dissolve grains of limestone (CaCO3) and dolomite
(Ca/MgCO3). The resulting solution then contains calcium and magnesium
ions(Ca+2 and Mg+2), called hardness; plus carbonate (CO32) and bicarbonate
(HCO3) ions, together called alkalinity. These are the dominant ions in most
waters, and they form the basis of water chemistry. The water also begins to
dissolve many other minerals and organic matter in the soil, and excess acidity
can cause leaching of valuable nutrients, which then pollute streams. As the
rainwater soaks into the soil and moves downward through many layers of Earth,
it is filtered by particles of dirt, which also adsorb many dissolved contaminants,
becoming relatively pure ground water. If it penetrates as deep as 10 meters,

this natural purification may be sufficient to render the water potable without
any further treatment, but wells from shallower depths cannot be considered free
of influence from the surface (penetration of pathogens from surface contamination, especially protozoan cysts). The water collects above impervious layers
of clay or stone, often extending over vast regions, and moves very slowly
through pores, cracks, and other channels in the ground as an aquifer. Most are
very ancient, and it may have been 50,000-250,000 years or more since the
water there fell as rain. Thus, well water from deep aquifers is considered to be
a non-renewable resource, like petroleum. Pumping out water faster than it
can be renewed can lead to a significant lowering of the surface terrain, called
subsidence (pronounced sub-SI-dense). Areas with varied geologic histories
may have several aquifers stacked on top of one another, each receiving water
from its own regional source and having its own unique chemical composition.
The upper level of an aquifer is called the water table, and in most places the
water table of the uppermost aquifer coincides with the water level of the lakes
and rivers in the region. That water flows directly to the surface, but water from
the deeper aquifers may be pumped out by wells, spewed out of volcanoes, or
drawn into the roots of plants by osmotic pressure and transpired out of the
leaves to become water vapor again, completing the cycle.
The Basics of Water Chemistry
Some understanding of basic chemistry is needed to understand water
chemistry, but the only background needed to begin this course is the ability to
recognize the symbols of the elementsC for carbon, H for hydrogen, Ca for
calcium, etc. The rest is explained below.
The various impurities that can be found in water can be categorized in several
different ways: Are they dissolved, particulate, or colloidal? Are they organic or
inorganic? Are they biological or mineral? Are they ionized or not? Are they
aesthetic contaminants, health-related contaminants, or just process nuisances?
Each of these will be discussed, along with the concepts of oxidation/reduction,
adsorption, and chemical equilibrium, which will be explained and demonstrated
with the pH scale and the alkalinity continuum.
a) Dissolved, Particluate or Colloidal: Water molecules are very tiny, and they
are also polar in nature, which means the molecules are not symmetrical: one
end is bigger than the other, so theyre pointed, and the big end also has a
partial positive electrical charge, while the small end is partially negative. They
look a bit like a Mickey Mouse head in silhouette, with the ears being the two
hydrogen atoms. Their size, shape, and electrical polarity enable water
molecules to be invasive and squeeze into the smallest spaces between the
molecules of solids and pry them apart. This is dissolution, and the molecules,
atoms, or small clusters that are liberated into the water (called a solution now)
are said to be dissolved if they are too small ever to settle under the influence
of gravity. Anything solid that is suspended in water but is large enough to settle
eventually is a particle. Molecules and clusters of intermediate size are called

colloids. Examples of colloidal materials include biological polymers such as


protein molecules, tannins and lignins leached from vegetation, tiny aggregates
of rust and clay, fragments of microbes and larger organisms, even complete
viruses, asbestos fibers, etc., up to a size of a few tenths of a micron in diameter.
This is also the same size range as the wavelengths of visible light (violet light is
about 0.40 um or 400 nm; red light, about 0.77 um or 770 nm), and that is why
particles of this size cause haziness or cloudiness in water, called turbidity.
Turbidity is measured by shining light through a test tube of the water and
measuring the intensity of the light that comes out the other side. However,
unlike ordinary photometers used in chemistry labs (which have the light source,
sample, and detector all in a line), the detector for turbidity measurements is
positioned at a 90 angle from the incoming light beam, so that only the reflected
haze is measured. This is called nephelometry, and that is the origin of the
Nephelometric Turbidity Unit, NTU. One NTU is not detectable by eye, but a
reading of 15 NTU is noticeably cloudy, and murky river water may have a
turbidity of several hundred. Public health or environmental regulations everywhere require the turbidity to be low at the time of disinfection, because pathogens can be shielded by particles. In the U.S., the requirement is <1 NTU 95%
of the time, (<0.5 NTU for those systems using conventional or direct filtration), with an absolute maximum of 5 NTU. When this is exceeded, Boil Water
Orders may be required by law. That is mostly because of protozoan cysts in
surface waters, which are difficult or impossible to kill. The large urban waterworks using the most advanced water treatment methods are able to produce
water with less than 0.5 NTU, but it invariably picks up millions of particles from
the pipes and mains of the distribution system and may reach the point of use
with several units of turbidity.
Discussion of Sizes and Dimensions: How big is 0.40 um or 400 nm? The
fundamental length or distance is the meter (symbol: m), but the unit most used
in describing particles in water is the micro-meter (one-millionth of a meter)
(symbol: um or m), commonly called a micron. A meter is about a yard long; a
millimeter (one-thousandth of a meter) (symbol: mm) is about 1/32 inch, and a
micron is one-thousandth of that: 0.00003937 in. Also, many people think in
terms of thousandths of an inch, or mils for very tiny measurements: a mil or
0.001 in. is 25.4 microns. A particle 40 microns in diameter is just barely visible
to people with 20/20 visionsmaller than 40 um is microscopic. The table below
will help illustrate the range.

Comparison of Particle Sizes


1-inch ball

25.4 mm (millimeters)
= 25,400 um or m or microns (micro-meters)
pollen
10 - 100 microns
smallest item visible to naked eye 40 microns
1 mil or 0.001 inch
25.4 microns

fog droplet
dirt
silt and clay
pathogenic protozoan cysts
Cryptosporidium oocysts
Cyclospora cysts
Giardia cysts
Entamoeba cysts
red blood cells
most bacteria and algae
turbidity
colloids
wavelengths of visible light
cigarette smoke
viruses
protein molecules
individual atoms

2 - 50 microns
40+ microns
- 20 microns
3 - 20 microns
3 - 7 microns
8 - 10 microns
8 - 12 microns
12 - 20 microns
7 microns
- 5 microns
0.1 - 5 microns
0.1 - 5 microns
0.40 (blue) 0.77 (red) microns
= 400 (blue) 770 (red) nm (nanometers)
10 1000 nm
10 250 nm
2 50 nm
0.05 0.25 nm
= 0.5-2.5 Angstroms
= 50 250 pm (picometers)

b) Organic or Inorganic: the term, organic originally meant made by an


organism, but now it refers to chemical compounds containing the element
carbon, in some form other than the simple carbonate minerals such as limestone. Inorganic generally refers to minerals dissolved in the watersalts,
metals, hardness, corrosion products, etc. Over the years since pollution
became an issue, organic came to mean man-made in factoriessynthetic, and
therefore unnatural and dangerous. Organic foodstuffs are those grown without
the aid of synthetic pesticides and fertilizers. These trends in the language are
dangerous, because they ignore significant inorganic hazards that are also the
result of industrialized life, such as toxic corrosion products and asbestos. In
water chemistry, the organic contaminants are either named individually or
grouped with acronyms such as:
TOC = Total Organic Carbon
NPTOC = Non-Purgeable TOC (purgeable CO2 [carbonate] removed first)
TOX = Total Organic Halide (containing chlorine, bromine, and/or iodine)
THM = Tri-Halogenated Methane
VOC = Volatile Organic Chemical
AOC = Assimilable Organic Carbon (digestible by microbes)
NOM = Naturally-occurring Organic Matter
DBP = Disinfection By-Products
c) Biological or Mineral: This is nearly the same as organic or inorganic, but a
distinction needs to be made for films on surfacesbiofilms versus scale
deposits. Usually the two are mutually exclusive, because the high temperatures that often lead to scale formation usually kills micro-organisms. However,
commercial ice makers of the recirculating cascade or spray types are susceptible to both, simultaneously. Scale develops because the minerals in the water
become progressively more concentrated as the ice grows, but psychrophilic

(cold-loving) bacteria and molds also form complex slimes that require regular
cleanup.
d) Ionized or not: When dissolved in water, some chemical substances split up
as unequal fragments with electrical charges, positive and negative. Each
charged atom or group is called an ion. Opposite charges attract and like
charges repel, just like the North and South poles of a magnet. Positive ions are
called cations because they migrate toward the cathode in an electric field, and
negative ions are called anions because they go toward the anode. (These are
pronounced CAT-ions and AN-ions.) As a generalization, inorganic chemicals
are most usually ionic compounds that ionize when dissolved, and organic
chemicals are usually non-ionic, but there are many exceptions. Three chemical
types are always ionic, whether organic or inorganic, and those are acids,
bases, and salts:
Acids: chemicals that liberate a hydrogen ion, H+, when dissolved in water.
The H+ is one charged fragment of the original molecule, and the
remainder becomes negatively-charged as a result. Example: citric
acid:
C6H8O7 C6H7O7 + H+
citric acid
citrate hydrogen
molecule
ion
ion
Bases: chemicals that liberate a hydroxide ion, OH, when dissolved in water.
The OH is one charged fragment of the original molecule, and the
remainder becomes positively-charged as a result. Example: sodium
hydroxide:
NaOH Na+ + OH
sodium
sodium hydroxide
hydroxide
ion
ion
Salts: chemicals produced by mixing an acid solution and a base solution
and then crystallizing or evaporating to dryness. Two things combine:
the H+ and OH combine to make water, and the other two oppositelycharged fragments combine to make a salt. When salts are dissolved
again, they always dissociate into + and ions. Example: sodium
citrate:
Na+ +
sodium
ion

C6H7O7
citrate
ion

NaC6H7O7
sodium citrate
molecule

e) Aesthetic Contaminants, Health-Related Contaminants, Process Nuisances:


These are not as obvious as they might seem. Regulatory agencies the world

over recognize that the offensive taste of iron in water is harmless and that toxic
levels of lead or a pesticide may have no taste or odor at all, and regulations are
always separated into mandatory health-related requirements and other aesthetic
requirements which are only recommended. However, there are cross-overs, and
others that cause problems only for equipment rather than for people. The
recommended pH range of 6.5-8.5 becomes mandatory if there is a problem with
corrosion of lead, copper, and cadmium plumbing materials. Excess sulfate or
TDS (Total Dissolved Solids) causes short-term diarrhea in visitors unaccustomed to a water supply, but this is considered only an aesthetic problem. Likewise, excessive silver causes a deathly graying of the skin and the whites of the
eyes, and excessive fluoride ion causes ugly stains and even malformations of
bones and teeth, but these are not considered to be health problems. Too much
zinc can cause vomiting, but since there is no permanent damage, zinc is also
only an aesthetic contaminant. As for process nuisances, there is nothing toxic
about silica or hardness and alkalinity, and there are no regulations for them, but
in excess they can lead to very damaging and costly scale buildup. Similarly,
turbidity is very important at the time of disinfection (when it must be very low),
but its importance to our industry lies in its ability to scratch valves, add to scale,
and plug filters.
f) Adsorption: Not to be confused with absorption (which is what a sponge
does), adsorption is the attraction of tiny particles or dissolved molecules to a
solid surface and holding them there by weak intermolecular forces. It is similar
in concept to magnetism and the attraction due to static electricity, but much
weaker. In theory, every atom in the universe has some degree of affinity for
every other atom in the universe, just like gravity. But, just as gravity requires
enormous masses like planets and stars to show its effects, adsorption requires
extremely tiny distances to show its effects. In adsorption, the particle in
question is randomly bounced around the solution by collisions with water
molecules and other molecules in the water. (This is called Brownian motion.
It is estimated that an atom or molecule in water is involved in a million-billiontrillion or 1027 collisions with other atoms or molecules every second. This is part
of the definition of temperature.) Eventually, by chance, it will be bounced so
close to the surface of a wall or another larger particle that there are very few
water molecules separating it from the surface. When that happens, those few
molecules produce only a few collisions from that side, and the particle is
overwhelmed by collisions from the other sides and tends to become plastered
to the surface by a continual barrage of collisions from the solution. This is the
physical half of adsorption. The chemical half occurs if there is any chemical
affinity between the particle and the material of the surface. If there is, the
particle will become attached (adsorbed) and stay there; if not, it will bounce off
right away or just diffuse away, later.
The adsorptive forces (called van der Waals or London forces) are so weak that
adsorbed substances can become desorbed rather easilyby adding certain
acids, by heating the system, or by merely removing the contaminant from the

influent water. For example, activated carbon filters or ion exchange beds
nearing exhaustion are subject to desorption if the water quality suddenly
changes for the better. That shows that these treatment techniques are
equilibrium (balance) phenomena in which sorption and desorption both occur
and achieve an average condition, like a well-matched tug-of-war. Since
adsorption requires a surface, commercial adsorbent materials have very large
surface areas and are exemplified by activated carbon, activated alumina, and
fine powders such as baking soda. But many substances are so very insoluble
or otherwise so readily adsorbable that even small surface areas can make a big
difference. For example, most heavy metal ions (lead, mercury, copper,
cadmium, silver, chromium) adsorb so strongly to the walls of both glass and
plastic sample bottles that more than half of the total contamination can be
missed in an analysis if the sample bottles are not treated with nitric acid first, to
cause desorption. Similarly, many chlorinated hydrocarbons like the polychlorinated biphenyls (PCBs) adsorb so readily to both metal and plastic
plumbing and filter materials that even coarse prefilters remove them very well.
The adsorption and reduction of disinfectant chlorine by activated carbon is a
special case. Activated carbon is a mild reducing agent and chlorine is a strong
oxidizing agent, so after chlorine becomes adsorbed, it then actually reacts with
the carbon. The chlorine is reduced to chloride ion (as in table salt and sea
water), one atom of carbon is oxidized to carbon dioxide, and both are released
to the solution (desorbed). Meanwhile, most of the spots on the activated carbon
where all this took place become auto-regenerated back to their original, likenew condition, ready to adsorb again. For free available chlorine (FAC), this
takes only about fifteen minutes, which means that a small amount of carbon can
achieve an acceptable steady-state condition if the flow rate is slow or
intermittent. For combined chlorine (monochloramine), the reaction is much
slower, and more carbon or more contact time is needed to achieve equivalent
reductions. The chemical reactions between activated carbons active sites
(C*) and these forms of chlorine are shown below. Note that any surface oxides
on the carbon are recycled when reacted with monochloramine, while they are
oxidized to CO2 and lost when reacted with free chlorine.
Free Chlorine
Cl2 + H2O HOCl

Cl

C* + 2Cl2 + 2H2O C*O2

4H+

C* + HOCl C*O +

H+

Cl

C*O + HOCl C*O2 + H+

Cl

+ H+

Combined Chlorine: Monochloramine


C* + NH2Cl + H2O C*O + NH3 + H+

10

(forming aqueous chlorine)


4Cl

Cl

(the overall reaction)

C*O + 2NH2Cl C* + N2 + H2O + 2H+

2Cl

Finally, most dissolved/suspended particles and molecules in drinking water that


are highly adsorbable to something usually do become adsorbed to a larger
particle before reaching the point of use. Thus, adsorbable contaminants can
often be removed by mechanical fine-filtration because the contaminant in
question is already adsorbed to a larger particle. If you remove the particle, you
remove the adsorbed contaminants along with it. This commonly applies to
heavy metal ions, many pesticides, other chlorinated hydrocarbons, viruses, and
asbestos fibers.
About Activated Carbon: Granular activated carbon (GAC) and powdered
activated carbon (PAC) are the predominant adsorbents used in our industry.
They can be made from nearly anything organic: coal, petroleum, wood, coconut
shells, peach pits, ion exchange resin beads, fabrics, even waste plastics. The
starting material is first charredheated without air or oxygen, so it doesnt burn
up. Everything that can be vaporized or melted bubbles out as tar or pitch,
leaving many holes and channels. Then the charred material is heated further,
to above 1000C (hot enough to melt aluminum and lead), with the introduction
of live steam or other activating chemicals. The superheated water vapor is
extremely corrosive, etching more holes and extending channels to an amazing
degree. Metallic impurities are preferentially attacked and washed out, resulting
in a significant purification of the original material.
However, the heat of activation does more than extend holes and channels and
increase the surface area of carbon; it also changes the fundamental crystal
form from amorphous carbon black to the perfect crystalline array of graphite
plates. The carbon atoms in graphite are arranged in sheets or plates of
interlocking six-atom rings that look like slices through a honeycomb. Such a
perfect arrange-ment causes the London forces to focus and concentrate at the
surface, making activated carbon the best (strongest and most general)
adsorbent known.
After activation, the carbon may be treated further to produce specific chemical
qualities on the surface. For example, an acidic environment produces carbon
with maximum capacity for heavy metals but minimal capacity for chlorinated
organics, while an alkaline environment does the opposite. Most grades used in
our industry are made for organic adsorption. When activation is complete, the
carbon is a delicate, airy material that is so full of holes, it can barely hold
together. It is crushed to a powder, and then proprietary binders are added to
form granules of the desired size. The final product has a total internal and
external surface area of more than 1000 square meters per gram, or half a
football field inside a piece the size of a pea.
Activated carbon adsorption is useful because the material has strong chemical
affinities for several important classes of contaminants that are common in
water. These are:

11

1. Disinfectant chlorine: Free available chlorine (FAC) is readily


adsorbed, then chemically reduced, and finally desorbed as chloride ion
along with one molecule of carbon dioxide, with auto-regeneration of most
of the carbons active sites and nearly infinite capacity. Combined
chlorine (monochloramine) is less easily adsorbed, requiring more
carbon or reduced flow rate for equivalent performance.
2. Organic compounds containing chlorine and other halogens: Simple
halogenated hydrocarbons are highly adsorbable to activated carbon. This
includes a great many pesticides (DDT, Endrin, Lindane, Chlordane, etc.),
industrial solvents (trichloroethylene, trichloroethane, tetrachloroethylene,
carbon tetrachloride, etc.), and disinfection byproducts (THMs including
chloroform, chloral hydrate, etc.).
3. Organic compounds containing benzene rings: These include some of the
most toxic chemicals, such as benzene, toluene, dioxins, polychlorinated
biphenyls (PCBs), and phthalate esters (plasticizers for vinyls).
4. Heavy metals: Lead, cadmium, and mercury adsorb readily, both as
dissolved ions and colloidal oxide or carbonate particles, but the capacity
is limitedsimilar to the capacity for THMs.
5. Taste and Odor (T&O) compounds: The substances produced by
microbes that are responsible for the common musty-earthy-mildewy T&O
are extremely well adsorbed and with very great capacity.
There is great variation in the adsorbability of dissolved/suspended substances,
and also great variability in the adsorptive capacity of different adsorbents. A
bed of granular activated carbon (GAC) may be exhausted with respect to
chloroform and other volatile organic compounds (VOCs) after only a few
hundred bed volumes, yet continue to adsorb PCBs for many thousands more.
Different grades and types of activated carbon have different capacities for the
same contaminant as well as various contaminants, which means that one must
be very careful and specific in making comparisons.
Chemists have developed a standard procedure for comparison of adsorption
capacities, called an isotherm. A Freundlich carbon isotherm is determined by
preparing several identical bottles of powdered activated carbon suspended in
water. (In German, eu is pronounced oi, so Freundlich sounds like
Froindlich.) Varying amounts of a contaminant are added to the bottles, and all
are mixed until adsorption has reached equilibrium under those conditions of
temperature and pressure. Then the carbon is filtered out and the solutions are
analyzed to find how much contaminant remains unadsorbed in each one. The
Freundlich equation is used to calculate the amount of contaminant that was
adsorbed per milligram of carbon, and each data point is graphed on logarithmic
graph paper with the carbon capacity in mg/g on the Y-axis and the final
equilibrium concentration in mg/L on the X-axis. Finally, the average line
representing all of the data points is drawn. That average line on the graph
paper is called the isotherm for that contaminant and that carbon under those
conditions. But that line covers a range of capacities; the one value used for

12

comparison purposes has been designated by international agreement to be the


capacity in mg/g of carbon on the Y-axis that corresponds to the value of 1.0
mg/L on the X-axis. If the isotherm does not cross the 1.0 mg/L point, the line is
artificially extended (extrapolated) to that level for the purpose. The Freundlich
Equation can be represented as:
q e = K (C e )

1
n

where qe = concentration of contaminant on the carbon at equilibrium (Y-axis


value)
Ce = concentration of contaminant in solution at equilibrium (X-axis
value)
K = a constant
1/n = another constant
The Ce is determined by analysis; the qe is calculated using the equation and the
two Freundlich constants that were determined by the chemist who published the
data.
By convention in our industry, chloroform, the main THM, has been selected as
the least adsorbable contaminant that activated carbon can be claimed to adsorb
effectivelyany contaminant that has a Freundlich isotherm qe value less than
that for chloroform cannot be said to be removed by carbon adsorption. We
draw the line at chloroform; anything less adsorbable than chloroform is
deemed not worth the trouble. Examples: in one test series, the Freundlich
capacity for chloroform is 2.6 mg/g GAC (the isotherm passes through 1.0 mg/L
above the X-axis at the point where the Y-axis reads 2.6 mg/g GAC), while the
equivalent value for trichloroethylene is 30 mg/g GAC. That is a much higher
value, meaning that that particular GAC will adsorb trichloroethylene much more
easily than chloroform. However, in the same data set, the capacity value for
methylene chloride is only 1.3 mg/g, and thus we say that methylene chloride
cannot be adsorbed efficiently by that GAC. The carbons capacity for it is too
small to make an economically viable product. See the example below.

13

S a m p l e F r e u n d lic h C a r b o n is o t h e r m s
100

mg/g Carbon Capacity (qe)

10

0 .1

0 .0 1

0 .0 0 1
0 .0 0 0 1

0 .0 0 1

0 .0 1

0 .1

10

m g /L a t E q u ilib riu m (C e )
T ric h lo ro e th y le n e

C h lo ro f o rm

M e th y le n e C h lo rid e

It is important to remember that carbon capacity figures derived from Freundlich


isotherms are to be used only for comparisonse.g., carbon A is better than
carbon B, or contaminant X is easier to remove than contaminant Y. They
should not be used as concrete capacities to calculate how long a filter should
last. The reason is that the isotherm data are produced at equilibrium, which may
take several days of stirring in the lab to achieve. But a bed of GAC or a filter
cartridge operates on a dynamic, flowing basis, and equilibrium conditions may
not be achieved even after a weekend downtime.
g) Chemical Equilibrium is a special property of the reactions and solutions of
many substances. Whenever you see a chemical equation with arrows pointing
both forward and backward, it is an equilibrium reaction. In an ordinary reaction,
the starting materials simply react until one of them is all gone, and then it stops
dead. But an equilibrium reaction does not automatically go to completion.
Instead, it grinds to a stall (achieves equilibrium: the reverse reaction equals the
forward reaction) at some characteristic point, even though there may still be
plenty of starting materials left. Every reaction has its own regular balance-point:
one reaction may stall after only 0.0000001% of the starting materials have been
used up; another may go to the point of being 99.999999% complete. The endpoint characteristic to each equilibrium reaction is defined by a number called the
Equilibrium Constant, KEq , which never changes. Whenever the concentration
of any of the reactants or products is changed (by reacting or adding or removing
something), all of the other substances in the equation change in compensation,
and they change in such a way as to preserve the value of the KEq. For
example, consider the most important equilibrium of all, which is also the
simplest:
The Water Equilibrium describes the dissociation of water molecules into their
component parts, hydrogen ions and hydroxide ions (H+ and OH):
H2O H+ + OH

KEq = 1014
14

1014 is a very small number, which means the reaction doesnt go very far, and
only about one molecule in ten million is actually dissociated at any instant. In
pure water with nothing added, H+ and OH will be identical in concentration,
equal in this simple case to the square root of the KEq, or 107. This reaction and
relationship are so important that a special shorthand notation has been
developed, called the pH system: use the exponent only, drop the sign and use
p as a symbol, and you get pH = 7.0 for pure water. (The pOH = 7.0, also.) All
this is important because its water were talking about, but also because H+ is
very important to water chemistry. The hydrogen ion is the smallest and most
chemically active piece of ordinary matter known. Hydrogen is the smallest of
the elementsa single proton with a single electron whizzing around it. Remove
the electron, and you have a single, lone, naked proton in solution that is highly
reactive, invasive, and corrosive. (Its not really naked and alone; sometimes it is
shown attached to a water molecule as H3O+.) H+ defines acidity; H+ is acid. So,
the water equilibrium defines acidity: pH less than 7 is acidic; pH greater than 7
is alkaline or basic. If either H+ or OH is increased artificially, by adding or
removing acid or alkali, the other changes in the same proportion so as to maintain the value of the KEq, which is chiseled in stone. Thus, if acid is added to
pure water to make the pH = 3, the equilibrium shifts so that the pOH = 11 and
the KEq is preserved as 1014.
The Carbonic Acid-Calcium Carbonate Equilibrium is the other most important
equilibrium in water chemistry. When water falls as rain it absorbs carbon
dioxide gas from the air, forming carbonic acid, which instantly dissociates into
bicarbonate and carbonate ions:
CO2 + H2O H2CO3 H+ + HCO3 2H+ + CO32
carbon water carbonic hydrogen bicarbonate
carbonate
dioxide
acid
ion
ion
ion
When this slightly acidic water soaks into the ground and contacts limestone
(calcium carbonate) or dolomite (mixed calcium and magnesium carbonate), it
dissolves some of the rock, making hardness and alkalinity:
2H+
acid

Ca/MgCO3
lime/dolomite

Ca+2 + Mg+2
hardness

+ H2CO3
alkalinity

(which dissociates
as shown above)

Hardness is the sum of all ions that react with soap to inhibit lathering and
precipitate soap scum or bathtub ring. It happens that they are all metals with
more than one + chargemostly Ca+2 and Mg+2 in most water supplies, but
also including Fe+2, Cu+2, Zn+2, Mn+2, etc., if present.
Alkalinity is a confusing term because it came into use before the chemistry
was understood. As a word, alkalinity just means the opposite of acidity, or
something that consumes or neutralizes acidity. When it was realized that both

15

of the carbonic acids and their ions in the two equations just above are the same
thing, alkalinity became the term in water chemistry to represent the sum of
CO2 + H2CO3 + HCO3 + CO32 , and thus the total buffering power of the water
to resist changes in acidity. All the double arrows show that the reactions go
both forward and backward, which means that limestone dissolved in one place
may deposit as lime scale in another place later, if the equilibrium shifts. The
equilibrium can be shifted by adding or consuming any of the constituents
involved. Remember that the relationship between reactants and productsthe
KEqis chiseled in stone, and any change from the outside will be compensated
for, instantly, in a way that preserves the value of the KEq.
The values of the three KEq for the three carbonic acid equilibria (one for each
arrow) are intentionally absent from this discussion, to avoid getting you bogged
down in numbers. What matters is that you get a feel for the concept of
equilibrium. It makes sense that adding more reactants (on the left side of an
equation) would make a reaction go further (to the right). But with equilibrium,
you can also make a reaction go to the right by removing some of the final
products. Its as if the removal creates sort of a chemical vacuum which
demands to be filled. The reverse is also true: to make a reaction reverse (go to
the left), you can either load up the right side (add more final product) or lighten
the left side (remove reactants). It is as though the reactants and products are
physically connected by the arrow, and you can push or pull on the reaction from
either end. And in chemical systems with several reactions connected in series,
like the first one, reprinted here, the connection goes through all of them. Note
that one of the final products of the third reaction (far right) is H+, or acid:
CO2 + H2O H2CO3 H+ +

HCO3 2H+ + CO32

If you add more acid, the entire three-reaction continuum will reverse (go to the
left), causing some CO2 to bubble away and be lost forever, thus reducing the
total alkalinity of the system. This is called dealkalization.
About Buffering: Just above, it was said that the presence of alkalinity buffers
water against changes in pH. Buffers are chemical stabilizers that use equilibrium to establish and maintain a chemical balance. The common mineral
acids such as hydrochloric, sulfuric, and nitric acid do not work as buffers
because their dissociation to produce H+ ions is not governed by equilibrium:
they all become virtually 100% dissociated the instant they are dissolved.
Because of this, they are called strong acids. Only weak acids and their salts
like carbonic acid and sodium carbonate can be buffers. The KEq of carbonic
acid is about 107, which means that normally, only one out of millions of
carbonic acid molecules is split into H+ and HCO3 ions at any instant. Note that
the value of the KEq is close to the value of the pH (and also the pOH) when
acidity is neutral: pH = 7 = pOH. That means carbonic acid is good for
buffering water near pH 7. However, note that the bicarbonate ion, HCO3 , still
has one hydrogen to lose, and bicarbonate ion therefore has its own KEq, which

16

is about 1011. That means sodium bicarbonate would be a good buffer for water
at pH values near pH 11. For comparison, the KEq of acetic acid (vinegar) is
about 105, which means that acetic acid would be good for buffering water in the
vicinity of pH 5.
So, how do buffers work, exactly? They work by having half of their capability
inactive, in reserve, so to speak, from both the acidic and alkaline point of view.
When the pH is equal to the exponent on the KEq, 50% of the weak acid
molecules in solution (carbonic acid in this case) are whole and 50% are
dissociated into H+ and HCO3 . If extra H+ is artificially added to the solution, it
will instantly be incorporated into the carbonic acid-bicarbonate-carbonate
alkalinity continuum, with very little effect on the pH. Likewise, if some OH ion
is artificially added, it will instantly react with and be neutralized by a H+ ion, but
that H+ will instantly be replenished from the alkalinity, again with very little
impact on the pH. As long as any total alkalinity remains in the water, it will be
able to absorb either H+ or OH in the vicinity of pH 7 without changing the pH
very much. Incidentally, it is possible to buffer a solution against changes in ions
other than H+ --chloride ion, Cl, for example.
h) Oxidation and Reduction are the result of an atom or molecule losing or
gaining an electron. Before that makes any sense, you need to understand that
atoms are made of a nucleus containing positively-charged protons and a
cloud of an equal number of negatively-charged electrons orbiting
around it, as shown in the familiar solar system icon for atoms and
atomic energy. Some of the outermost electrons may be lost rather
easily (or not, depending on the element), thus destroying the balance
between protons and electrons and changing the character of the atom greatly.
When an electron is lost, the protons outnumber the electrons by one + charge,
and the atom as a whole then becomes oxidized with a +1 charge, as in the
sodium ion, Na+. Conversely, if the electron cloud is capable of accepting
another electron, then the electrons would outnumber the protons by one
charge and then the atom would become reduced and have a 1 charge, as in
the chloride ion, Cl. In their elemental forms of sodium metal and chlorine gas,
the atoms are balanced and electrically neutral and have a net charge (oxidation
state) of zero.
Consider the enormous difference between a piece of sodium metal and a cloud
of greenish chlorine gas on the one hand and a spoonful of salt crystals or sea
water on the other. Yet they are almost identicalboth are a collection of
sodium and chlorine atoms, and they both have the same number of electrons
over all, but in the salt crystals, the sodium atoms all have one electron too few,
and the chlorine atoms all have one electron too many. They balance out in the
NaCl crystal, but that is not the same. Na+ ion is simply different from Na0, and
Cl ion is nothing like Cl0. So, electrons are extremely important to chemistry,
and when an atom becomes oxidized or reduced, its chemistry changes
completely. The iron atoms in the centers of the hemoglobin molecules in our

17

red blood cells must be in the +2 oxidation stateferrous iron, Fe+2to work
properly in carrying oxygen to our tissues. If they get oxidized to ferric iron, Fe+3,
by nitrite, cyanide, or carbon monoxide, for example, the hemoglobin doesnt
work properly, and we die. One tiny electron on each iron atom makes all the
difference.
Oxidation and reduction always occur together, because they are like the two
sides of a coin. When an atom or molecule loses an electron (becomes
oxidized), that electron must go somewhere, and the atom or molecule that
accepts it becomes reduced. When a reducing agent does its job, it becomes
oxidized in the process. When an oxidizing agent does its job, it becomes
reduced in the process. Oxidation is loss of electrons, and reduction is
acceptance of those same electrons by a partner in the reaction. Every
chemical has its own unique tendency to gain or lose electrons, called its
oxidation-reduction, or redox potential, or ORP, which is expressed in volts
and is available from reference books. These voltages can be used to calculate
whether a particular reaction will go without introducing energy from the
outside, but that will not be discussed here. Some examples:
the simple battery used for teaching electricity, made by dipping a zinc bar and
a copper bar into a solution of copper sulfate: Zinc loses electrons more easily
than copper does, so the zinc becomes oxidized and the copper becomes
reduced. If you simply dip a zinc bar into a solution of copper sulfate, you get a
copper-plated zinc bar which is badly corroded, because every atom of copper
that plates out on the zinc bar is matched by a zinc ion liberated into the
solution. If you use both bars and connect them with a wire to form an electrical
circuit, the plating process will proceed until the zinc bar disintegrates and breaks
the circuit. The reaction below describes both situations:
Cu+2 + Zn0 Cu0 + Zn+2
the oxidation of hydrogen sulfide (rotten egg smell) by chlorine: if only the
minimum amount of chlorine is used, the result is elemental sulfur, which is a
solid that must be filtered afterward. However, it is possible to oxidize the sulfur
atom all the way to sulfate if more chlorine is used. The first reaction below
shows the sulfur being oxidized from the -2 state in H2S to the zero state in
elemental S, and the second one shows the sulfur being oxidized to the +6 state
as sulfuric acid.
H2S

Cl2

H2S

4 Cl2

2 H+ + 2 Cl
+ 4 H2O

8 H+

S
+

8 Cl

H2SO4

Most elements have at least two possible oxidation states. Note chlorine, below,
which has more possibilities than most elements:

18

Oxidation
Chemical State
Name
Comments
ClO4
7
perchlorate ion
used in rocket fuel

ClO3
5
chlorate ion
used in explosives
ClO2
4
chlorine dioxide gas used for oxidation and
disinfection

3
chlorite ion
used as a bleaching agent
ClO2
ClO
1
hypochlorite ion in laundry bleach
Cl2
0
chlorine gas
elemental chlorine
Cl
-1
chloride ion
as in table salt and sea water
This also demonstrates some useful rules of nomenclature:
-ate = suffix denoting a high oxidation state; same as -ic suffix for acids*
-ite = suffix denoting a low oxidation state; same as -ous suffix for acids*
per- (short for hyper-) = prefix denoting higher oxidation state than -ate
hypo- = prefix denoting lower oxidation state than -ite
-ide = suffix denoting the complete absence of oxygen
*names of some acids: nitric acid, nitrous acid, sulfuric acid, sulfurous acid,
perchloric acid, chloric acid, chlorous acid, hypochlorous acid
Potassium
The Electromotive Series is a listing of metallic elements in
descending order of their oxidation-reduction potential or ORP.
Sodium
One of the characteristics of metals is the free movement of the
Magnesium
outer valence electrons, which makes them good electrical
Aluminum
conductors. The ORP voltage is a measure of each elements
Zinc
readiness to lose electrons (be oxidized). They are arranged
Iron
at right, with the most reactive element at the top and the least
Tin
reactive at the bottom. Elemental potassium and sodium are
Lead
so reactive that they react with water as if it were strong acid,
Hydrogen
making NaOH or KOH and hydrogen gas. At the other end,
Copper
everyone knows that silver, platinum and gold are noble
Mercury
metals that are difficult or impossible to corrode. This list can
Silver
Platinum
also be used in the reverse: the ions of a metal can cause
the corrosion of any metal above it in the list. For example,
Gold
if a copper wire is immersed in a solution of silver ions, the
Ag+ Ag (plating) and the Cu Cu+ (corrosion). Thus, silver metal is relatively
non-corrodable, but silver ion is highly corrosive. Alchemists old name for silver
nitrate (AgNO3) was lunar causticthe Ag+ ion is so caustic and corrosive that
it will even react with the proteins in your flesh. Perhaps that is why Ag+ and
Cu++ have some weak antibacterial activity, even at low concentrations.

19

B. Drinking Water Quality Standards


Every sovereign nation adopts its own set of standards, or norms for drinking
water quality. The most influential and widely-recognized standards are those
developed by the United Nations-World Health Organization, the European
Union, Japan, and the United States EPA. They are all changed or updated
regularly, so any listing here would be out of date quickly. There are always two
types of regulations:
Primary requirements relate to health effects, and compliance is mandatory.
Secondary requirements relate to aesthetic effects and are only recommended.
In addition, the U.S. EPA also establishes maximum contaminant level (MCL)
goals for some contaminants which for some reason cannot presently be
attained. For example, all carcinogens are automatically given an MCLG of zero,
even though everyone understands that zero is impossible to achieve.
In the United States, all water systems that regularly supply drinking water to at
least 25 people or 15 service connections are Public Water Supplies which
must comply with the regulations. If there is only one building or subscriber,
such as a rural factory or school, it is a Non-Community Public Water Supply. If
there is only one site and the 25 people are not necessarily the same 25 people
every day, as at a rural restaurant or service station, it is a Non-Community,
Transient Public Water Supply.
C. Centrally Treated Water Supplies
As a company, we support the practice of large-scale, central treatment of
drinking water by professionals. The municipal waterworks generally do an
excellent job, and they make our job easier. But the smaller the customer base,
the fewer resources are available, and the smallest systems have the most
difficulty meeting the requirements. Also, the smaller the system, the more likely
it has a ground water source, and those generally require less treatment than
water from lakes and rivers. Surface water often contains color (tannins) and
high turbidity which must be coagulated with special chemicals before filtering.
These tannins are also the major source of unwanted byproducts of chlorine
disinfection such as THMs and many other chlorinated organic compounds.
Many surface waters are mostly snow-melt or rain water with relatively little total
mineral content, called Total Dissolved Solids (TDS). This allows the water to
be corrosive, or aggressive to plumbing materials so that asbestos fibers and
lead, zinc, cadmium, copper, and iron corrosion byproducts leach into the water
from asbestos-cement pipe, brass fittings, lead-based solder, and copper and
galvanized pipes. The microbiological quality of tap water is generally very good
as the water leaves the waterworksoften less than 10 HPC/mL (Heterotrophic
Plate Count), but after a few hours or miles of transit through the distribution

20

system, the chlorine dissipates and allows the bacteria that survived disinfection
to multiply. In addition, biofilms containing many species build up and flake off
into the water. It is important to remember that less than one-half of 1% of the
potable tap water produced by large waterworks is actually consumed by people.
The rest is used for flushing toilets, washing everything from laundry to streets,
watering lawns, fighting fires, etc.
D. Individual Water Supplies
The smallest of all water supply types, whether public or private, is the individual
system for a single user. This is usually a private well, because the minimum
treatment for a surface supply is often more trouble than is worthwhile: chemical
disinfection and fine-filtration to remove parasites are unavoidable, and the
addition of flocculants and the holding time needed to promote coagulation may
also be needed. We usually recommend that well waters be chlorinated (for
disinfection) and that will also change any iron or sulfide to filterable particles
which are easy to remove. Well water often requires softening or conditioning
(removal or reduction of hardness), and that can also remove small to moderate
amounts of iron and manganese by ion exchange.

SECTION II
DESCRIPTION OF WATER PROBLEMS
A. Microbiological Problems (Primary MCL: absence of Fecal Coliforms,
viruses, cysts)
Drinking water may contain all types of microbesviruses, algae, molds,
bacteria, and parasitesbut they are not all necessarily dangerous to health.
Some just cause bad taste and odor or clog filters. Algae and molds seldom
have any health significance, but all products and claims relating to the other
three microbial types are watched carefully by public health officials. At present,
Cyst Reduction is the only permitted microbiological reduction claim, unless
bacteria, viruses, and cysts are all killed/removed simultaneously. The requirement for that is 3-log reduction (99.9%) for Cryptosporidium oocysts, 4-log
reduction (99.99%) of two kinds of virus, and 6-log reduction (99.9999%) of
Klebsiella terrigena bacteria, which is an environmental coliform. Any product
that can document all three abilities can be labeled a microbiological purifier.
1. Viruses are the smallest microbesonly a few hundredths of a micronand
they are not truly alive, because they are not fully functional. They are parasites
of individual cells and cannot do anything by themselves. Most are very specific
and limited in the type of cell they can infect, and the only ones in water that are
of interest to humans are those that come from human sewage. Their numbers
in contaminated raw waters are usually about one per cent of the fecal coliform

21

count (see below, in bacteria). They are too small and too few to cause any
mechanical or maintenance problems for filters; the only problem they cause is
disease. They are only a minor problem in developed nations with good sewage
and water treatment practices, but rotaviruses in particular are the #1 cause of
human death and misery world-wide: infant diarrhea caused by rotaviruses in the
drinking water accounts for fully half of all human deaths on Earth each year.
2. Bacteria are the next-smallest microbes, single cells ranging from about 0.2
microns to 10 microns in size. They are the smallest, simplest, and most ancient
fully-functional life form. They have only one chromosome, and the genetic
material is not even organized in a nucleus. Bacteria are the most numerous
and most varied life form on Earth, accounting for more total biomass than any
other. Every gram of good, productive topsoil contains some 10 billion bacteria,
comprised of thousands of colonies with millions of organisms in each colony.
They occupy every known habitat. Most of them are harmlessindeed, most
have no interaction with people at all; a few cause inconvenience in the form of
bad taste and odor or a tendency to cause clogging of fine-filters; and a very few
can cause disease. The most common odor-causing types merit special
mention: they are filamentous bacteria resembling mold under the microscope,
except that they are much thinner than molds, and are called actinomycetes.
Many of them produce antibiotics of the -mycin type, in addition to some of the
same earthy-musty-mildewy-moldy-fishy odor compounds produced by molds
and algae. Bacteria are so varied, it is difficult to know how to categorize them,
but the following types are important to know about:
-autotrophs: those that make their own food by photosynthesis or oxidation/
reduction of minerals
-heterotrophs: those that must consume organic matter from the
environment to live
-HPC organisms: those detected by the Heterotrophic Plate Count
procedure
-pathogens: those capable of causing disease
-opportunistic pathogens: those capable of causing disease only if given
an unusual opportunityopen wounds, burns, defective immunity, etc.
-coliforms: those that resemble (biochemically) Escherichia coli, the
predominant commensal intestinal organism of mammalsindicators of
sewage contamination
-Total Coliforms: an important public health screening test to identify all
coliform organisms, both fecal and environmental, belonging to the genera
Escherichia, Klebsiella, Enterobacter, Citrobacter, and Serratia

22

-Fecal Coliforms: an important public health test to identify E. coli,


specifically
-Anaerobic bacteria: those that prefer or require the absence of dissolved
oxygen; they often produce septic smells, and some produce rotten egg
smell
-iron bacteria: partial autotrophs that oxidize dissolved iron to rust
-sulfur bacteria: anaerobic autotrophs that chemically reduce sulfate ion or
metabolize organics containing sulfur to produce the stink of rotten eggs, or
hydrogen sulfide, H2S, and also organic sulfides, all with offensive T&O.
Silver Filters or filters containing silver as an antibacterial agent are very
common and deserve special mention. The silver ion, Ag+, has a long history in
water treatment, going back to Roman times, and there is no doubt that it can be
an effective agent against many bacterial pathogens if the concentration and
contact time are sufficient. However, it is illegal in the U.S. to make that claim,
because it is not effective against all pathogens. Many bacteria are relatively
immune to silver, and it has no effect on viruses or cysts at all, so antimicrobial
claims can easily be dangerously misinterpreted. Thus, such products are
allowed to claim only that the silver is bacteriostatic, meaning that rapid,
uncontrolled bacterial growth inside the filter is inhibited. Further, it is illegal to
suggest that silver has any effect on harmful or pathogenic bacteria; a manufacturer may claim only that silver inhibits the growth of those harmless bacteria that
might produce bad taste and odor or premature clogging of filters. Unfortunately, silver ion is not very effective against many of those bacteria, so we are in
the ridiculous situation of being unable to make an accurate and truthful claim
because it could be misinterpreted, and unwilling to make the only permitted
claim because it is not very accurate. Everpure, Inc. now promotes the use of its
filters with silver in applications where they would be most useful (where they are
required by law in other countries, or where bacterial pathogens might be present
and users need all the help they can get), while using labeling language that
complies with U.S. law.
Bacteriological Testing: For some reason, the bacteria that inhabit drinking
water systems are difficult to detect by culturing them. More than 99% of those
shown to be present by staining them with dyes and counting them with the aid
of a microscope just dont grow into countable colonies when spread onto a
nutrient medium in a petri dish. They are called non-fermenters or viable but
nonculturable organisms. Part of the reason for the error in plate counts may be
that many are trapped in a clump of biofilm containing hundreds of individuals,
which grows out as a single colony. Or, it may be that their growth requirements
(nutrients, pH, temperature, atmosphere, salinity, etc.) are simply too exotic and
different from what is provided by the standard test conditions. However, the
main reason is believed to be the phenomenon of bacterial dormancy: each
specific strain of bacteria seems to require a rest of some weeks or months after

23

growing actively for a while. At any moment, the bacterial population of a municipal distribution system will be comprised mostly of three to seven specific
types, each following its own schedule of growth, decline, dormancy, and
regrowth, and almost all of them are dormant most of the time. That means that,
regardless of the specific plate count procedure used, the true bacterial numbers
are at least 100 times, perhaps as much as 1000 times higher. Thus, plate
counts are not very accurate or informative. It is folly to try to compare individual
plate count results; only controlled bacteriological testing programs, in which the
same trained person takes the samples in the same way, at the same place, at
the same time of day, etc., yield scientifically valid data that can be compared.
Even then, the variation is so great that the averages must differ by more than a
log (logarithm) or power of ten to be considered statistically different. For
example, an average influent plate count of, say, 350/mL and an average filter
effluent plate count of 3450/mL are not significantly different numbers,
statistically speaking.
Tests for coliform organisms are done on 100 mL volumes, either by filtering 100
mL through a membrane and then culturing the membrane, or by adding special
chemicals to a bottle containing 100 mL of sample. The special nutrient medium
and temperature are inhibitory to most non-coliform organisms, but many
unwanted HPC organisms may grow anyway, making the true coliforms difficult
or impossible to detect. When that happens, the lab analyst may reject the
sample with a TNTC notation (Too Numerous To Count) and request a new
one. In the U.S., any repeat coliform tests are to be analyzed using a PresenceAbsence test procedure, in which chemicals called MMO and MUG are
added to 100 mL of the sample. If there are any total coliform organisms in
the sample, their metabolism will turn the MMO yellow by the next day. If there
are any fecal coliform organisms (E. coli) in the sample, they will turn the MMO
yellow and also metabolize the MUG to something that fluoresces chartreuse
when viewed with UV illumination.
3. Algae used to be considered plants, but the world is no longer divided into
animal, vegetable, or mineral. Now, phytoplankton (algae) and zooplankton
(protozoa) are both placed in a new fifth kingdom called Protoctista , giving us
the present classification into bacteria, fungi, protoctists, plants, and animals.
Algae are often one-celled and microscopic in size. They produce three types of
problem, two of them serious. Several kinds of microscopic one-celled and
filamentous algae produce the musty-earthy-mildewy-moldy-fishy taste and
odor which ranks No. 2 in consumers complaints of tap water (after chlorine
T&O). Some of the same algal species are also capable of causing serious
clogging of filters. During seasonal algae blooms in reservoirs, their numbers
may be so enormous that the municipal filtration plant output is less than half of
actual productionthey use more water backwashing the filters than they put
into the distribution system. When they take short-cuts at the waterworks, pointof-use filters also clog quickly. The less serious problem is toxicity. Two or three
common species produce both nerve and liver toxins that can be deadly. This is
not a serious problem because the water does not become dangerous until after
24

it has become so smelly and unsightly as to be disgusting to a person. Thus,


pets and farm animals may be affected by algal toxins in ponds and puddles, but
hardly ever people. Algal toxins are not presently regulated, but they are being
watched carefully.
4. Molds are fungi. Molds in water are mostly microscopic and single-celled,
although they can grow into sheets of slime and larger structures like mushrooms if given the chance. They often contaminate water-using equipment via
airborne spores. They are a major source of the musty-earthy-mildewy-moldyfishy T&O compounds that are also produced by algae and filamentous
bacteria. With one exception, molds are harmless, and the only problems they
cause are moldy T&O and ugly slimes. The exception refers to people with
damaged immune systems, who can die from some fungal infections.
5. Protozoa and Larger Parasites often escape disinfection processes, and
physical removal by mechanical filtration is often the most cost-effective remedy.
Boiling for one minute will kill everything except bacterial spores, but that is the
most costly approach. No chemical disinfectant has been found reliable, and
standard ultra-violet systems are ineffective. It happens that the parasite most
difficult to kill or remove is also one of the most prevalentCryptosporidium
oocysts may be as small as 3 microns in diameter, but most are 4-7 um in size.
They can be found in virtually all surface raw water supplies worldwide, and no
municipal or regional waterworks anywhere can guarantee killing or removing
them all. Most strive for 99.9% reduction using multiple process barriers, but
many do not succeed. Thus, every person using a surface water supply needs a
fine-filter with 99.9% efficiency (or better) for reduction of 1-micron particles at
the point of use. So far, Cryptosporidium is the smallest, and 1-micron filtration
is good enough. All other known waterborne parasitesGiardia , Entamoeba,
and Cyclospora cysts, various round worms, tapeworms, flukes, and their eggs,
Schistosoma larvae, etc.are much larger.
B. Iron Water (including Manganese). Secondary MCL: 0.3 mg/L Fe + Mn
Iron and manganese may be in the water as it is pumped up from wells, or iron
pipes and other equipment may get oxidized (corroded) to form rust. Iron is
almost always in the ferrous form, the Fe+2 ion, when fresh from the well or
corrosion, but it is easily oxidized further, to the ferric form, Fe+3. Ferrous iron is
the one producing metallic taste. Ferric iron in water attracts hydroxide ions so
strongly that it will steal them from water if the pH is above about 5.0. The
reaction produces iron floc, which is a gooey, rust-colored mass.
Fe+3 + 3 H2O Fe(OH)3 (solid) + 3 H+

10+33

KEq =

The exceptionally large equilibrium constant indicates that the reaction will go
essentially to completion, and hardly even a single atom of Fe+3 in solution will

25

escape precipitation. This is the usual cause of consumer complaints of yellow


stains on porcelain and laundry. The usual remedy is either oxidation to Fe+3
followed by filtration to remove the iron floc, or removal by ion exchange water
softening if the concentration is not too high. Oxidation is usually done by
feeding liquid chlorine bleach with a chemical feed pump, followed by a contact
tank and filtration system to filter out the floc. Chlorination is preferred because
it also disinfects, but it can also be done with oxidizing media such as
manganese greensand and granular brass. Actual rust is a combination of the
oxides of both ferrous and ferric iron: FeO/Fe2O3, sometimes written Fe3O4.
Iron in water from wells with high organic content is sometimes combined with
huge tannin and lignin molecules from rotting vegetation, called color bodies, to
form an even more deeply-colored product called heme iron, which is difficult to
remove. (Heme refers to the blood pigment, which has an atom of ferrous iron at
the center of an organic complex. But the organic part of heme and hemoglobin
is nothing like tannins and lignins.) Some tannin molecules are large enough to
be called colloidal, and it is often possible to remove most heme iron by finefiltration in the sub-micron range. But if the molecular weight is too small, it must
be removed either by ultrafiltration or nano-filtration, or by chemical coagulation
followed by standard filtration.
Manganese usually occurs in combination with iron in well waters, and it makes
the stains on laundry and porcelain even darker. When Mn+2 is oxidized by
dissolved oxygen or chlorine, the result is manganese dioxide, MnO2, which is
dark brown in color. It is removed by the same methods used to remove iron.
C. Sulfur Water. Secondary MCL for odors: 3.0 TON
Sulfur Water refers to the stink of hydrogen sulfide, H2S, which is almost
always derived from bacterial activity. There are several species from two
different families: one family specializes in reducing sulfate ion, SO42, to sulfide,
S2, which then immediately acquires two hydrogen ions to become H2S. The
other family consumes organic matter containing sulfur and metabolizes the
sulfur to hydrogen sulfide. H2S makes the infamous rotten egg smell, and it is
also an acid which can cause rapid corrosion of all types of plumbing materials.
Therefore, it is always important to treat sulfur water, even if people get
accustomed to the smell. Sulfur and iron and the bacteria that oxidize or reduce
them often occur together, and when they do, the laundry and plumbing stains
are black, due to black ferrous sulfide, FeS. Such bacterial mixtures often
produce peculiar septic odors that are similar to that of hydrogen sulfide.
Analysis of water samples for hydrogen sulfide is possible, but since it is a
volatile gas, a special preservative must be used at the time of sampling.
However, the nose is a superior detector, and since the odor threshold is so low
that even the tiniest concentrations must be removed, an accurate analysis is
usually not needed. The only good remedy is chemical oxidation followed by

26

filtration to remove the elemental sulfur. Activated carbon can remove low levels
for a short time, but it is inefficient and has low capacity. New catalytic carbons
are an improvement, but a long contact time (slow flow rate) is still needed.
Chlorination is the preferred approach because it also disinfects, but oxidizing
media such as manganese greensand and granular brass may also give
satisfaction.
D. Acid Water. Secondary MCL: pH between 6.5 and 8.5
A water supply may be too acidic because of acid rain, contamination from acid
mine drainage, or it may simply be too pure (like distilled) and have insufficient
alkalinity to buffer the water against atmospheric acids. The secondary MCL
becomes mandatory in the U.S. only if it is necessary to control the corrosion of
lead- and copper-containing plumbing materials. (Other materials also corrode,
but they are less toxic.)
The remedy is to add alkalinity to the water, and there are two approaches. If it
is a private well that is being or is to be chlorinated, it is a simple matter to add a
solution of sodium carbonate (soda ash, Na2CO3), or sodium hydroxide
(caustic soda, NaOH), to the chlorine bleach. If there is no chemical pump, use
granular bed filters containing calcite (calcium carbonate: ground limestone or
marble) or magnesia (magnesium oxide) media, or a mixture of the two. These
dissolve slowly in the acid water, consuming excess acid and providing the water
with additional hardness and alkalinity.
E. Alkaline Water and Excessive Alkalinity. Secondary MCL: pH between 6.5
and 8.5
Alkaline water and excessive alkalinity are not necessarily the same thing. True
alkaline waters are rare and occur mostly in desert areas or regions with geologic
deposits of trona, bauxite, borax and other alkaline ores. Such waters are often
undrinkable due to high TDS or salinity, regardless of the pH. Excessive
alkalinity may be paired with high hardness, or it may be due to less extreme
levels of the same minerals that produce alkaline waters. The notion of
excessive alkalinity is relative, also. Generally, more than 250 mg/L alkalinity
can be expected to cause problems with lime scale, but levels as low as 80 mg/L
are preferred by some producers of post-mix beverages requiring carbonated
water for soft drinks. That is because any alkalinity in the water will consume
some of the acidity from acidulants such as citric acid in the syrup concentrate
that give the necessary tartness to the drinks. In addition, alkalinity in the water
fights against the carbonation process itself. Thus, commercial dealkalization
technology is very important.
All remedies to excessive alkalinity involve adding acid to the water in some way.
It is not advisable to add acid to the chlorine solutions being fed for disinfection
and/or oxidation, as is often done with alkaline additives used to treat acid

27

conditions, because that would produce toxic chlorine gas in the storage tank
before it could be fed. Acids must be dosed separately. One way is with a
simple pot-feeder cartridge similar to those used for many polyphosphate
feeders, except that the material being dissolved and fed is a food-grade
acidulant such as citric acid, tartaric acid, malic acid, etc. A portion of the
incoming water is directed to the feeder, which dissolves some of the acid to
form a concentrated solution which is then metered back into the main stream
through an orifice that is sized to produce the proper proportional feed.
The ion exchange approach is often preferable, especially if the water is also too
hard or has excessive TDS. Dealkalization by ion exchange involves the use of
a cationic ion exchange resin in hydrogen form (loaded up with H+ ion or
acid): any cations such as the hardness ions (Ca+2 and Mg+2), plus sodium,
Na+, potassium, K+, etc. are exchanged for H+ ion, which then immediately
neutralizes one H-equivalent of alkalinity. This may be by direct combination
with OH to make water, or the H+ may combine with one of the alkalinity ions
(carbonate or bicarbonate), driving that equilibrium one step to the left and
causing one molecule of CO2 to bubble away and be lost from the system.
Either way, this treatment approach represents a double-whammy: not only is
the excess alkalinity reduced; an equivalent amount of hardness or sodium ion is
also removed, and the overall TDS is reduced as well. A degree of added
control can be achieved by using a weakly-acidic cation resin (WAC resin)
instead of the strong-acid cation resin used for water softening. (See the next
section for a full discussion of strong and weak resins.) The result in this case
is that such a resin will remove only the amount of hardness that is balanced by
alkalinity. For example, if a water has 250 ppm of hardness but only 200 ppm of
alkalinity, WAC resin will remove only 200 ppm of the hardness, and the H+ that
is liberated from the resin will neutralize all of the alkalinity. If the alkalinity level
is greater than the hardness level, WAC resin will remove all of the hardness but
only an amount of alkalinity equal to the original hardness level. (This kind of
arithmetic is permitted only when the hardness and alkalinity values are
expressed as CaCO3.)
F. Hard Water (not regulated)
Hardness in water is the sum of all ions that react with soap to produce soap
scum or bathtub ring, and also inhibit lathering. The problem ions are all
metals with more than one + chargemostly calcium (Ca+2) and magnesium
(Mg+2) in most water supplies, but also including zinc (Zn+2), copper (Cu+2),
manganese (Mn+2), and others if present. In addition to interfering with
cleaning, hardness also combines with alkalinity to form lime scale, which is
calcium carbonate, CaCO3. Scale is a problem because it forms hard deposits in
water lines and water-using equipment which can scratch valves, insulate the
equipment and interfere with heating or chilling operations, and cause clogging.
The standard treatment for hard water is by softening, or water conditioning
by sodium-cycle ion exchange. Hard water is directed through a bed of cationic

28

ion exchange resin in sodium form (with Na+ ions attached to the resin), which
exchanges the Na+ ions for hardness ions in the water.
However, softening is not always desirable, because the Na+ ions put into the
water may be unwanted. Even if the possible negative health effects of Na+ on
blood pressure are avoided by using K+ as the exchanged ion instead of Na+,
both the taste and the TDS effect of either one can make the treated water taste
salty and be unsuitable for brewing coffee or making soft drinks. The taste
threshold for Na+ is in the 100-150 ppm range for most people, and softening
water with a hardness of 217 ppm as CaCO3 will result in the addition of 100
ppm Na+ to the water. Thus, about 15 grains per gallon or 257 ppm as CaCO3
should be considered the maximum hardness that can be treated by softening
without producing a salty taste. Sometimes reverse osmosis is used to remove
excess sodium and other dissolved minerals with health effects.
The Langelier Index (LI) is a system for estimating or predicting the amount or
degree of problems with lime scale a particular water supply will cause. It is
based on calculating the pH at which the water would reach the saturation point
for calcium carbonate (pHs) using the data from a chemical analysis of the water.
Specifically, the LI is equal to the actual pH minus the calculated pHs. This
produces a number, usually between 3 and +3. If the LI is positive, the water
will deposit calcium carbonate; if the LI is negative, the water will dissolve
calcium carbonate. To calculate the pHs, you need to know the total dissolved
solids (TDS), the concentrations of calcium ion and total alkalinity, and the actual
pH of the water. You also need to decide what temperature youre interested in,
and it is useful to have an electronic calculator that will give you the logs
(logarigthms) of the concentrations of calcium and alkalinity. A short log table is
given below, along with tables for special constants derived from the temperature
and the TDS.
pHs = A + B log(Ca+2) log(alkalinity)
where A = constant derived from temperature
B = constant derived from TDS
Note that a negative LI is sometimes misused to predict the corrosiveness of a
water to metal plumbing materials. The LI does have some influence on
corrosion, but it is only one of many factors, and it is too simplistic an answer to
that question.
Temp. Constant
C
A
log
0
2.60
4
2.50

TDS
mg/L

Constant
B

0
100

9.70
9.77

29

Ca+2 or alk.
mg/L as CaCO3
10
20

1.00
1.30

8
12
16
20
25
30
40
50
60
70
80

2.40
2.30
2.20
2.10
2.00
1.90
1.70
1.55
1.40
1.25
1.15

200
400
800
1000

9.83
9.86
9.89
9.90

30
40
50
60
70
80
100
200
300
400
500
600
700
800
900
1000

1.48
1.60
1.70
1.78
1.84
1.90
2.00
2.30
2.48
2.60
2.70
2.78
2.84
2.90
2.95
3.00

Example LI Calculation
Given: pH = 8.40; TDS = 400 ppm ; Ca+2 = 200 ppm as CaCO3; Alk. = 300 ppm
as CaCO3 How bad will scaling be in an ice maker? (0C)
LI = pH [A + B log Ca+2 log alk.]

Interpretation of LI Values
+3 = very severe scaling tendency
+2 = severe scaling tendency
+1 = moderate scaling tendency
0 = no scaling tendency
-1 = slightly corrosive to lime scale
-2 = moderately corrosive to scale
-3 = severely corrosive to lime scale

= 8.40 [2.60 + 9.86 2.30 2.48]


= 8.40 7.68
= +0.72

When softening is not an option, the scale-forming potential of the water can be
limited by reducing the pH or the alkalinity, which are discussed above. But
often the simplest and most cost-effective treatment to inhibit lime scale formation and deposition is Everpures proprietary InsurIceTM System, which is the
combination of fine-filtration and low-level polyphosphate feed. (Regardless of
the name, the InsurIce approach also works very well for coffee brewing. The
conditions inside steamers are too extreme, so they are treated by dealkalization.) Polyphosphate treatment is very important and needs its own discussion.
Polyphosphates are polymers of the phosphate ion, PO43. The smallest one is
pyro-phosphate, P2O74; the next is called tripolyphosphate, P3O105, and so on,
up to large molecules of 20 or 30 phosphates linked together in chains, having
very large electrical charges. They are useful for inhibiting lime scale in two
ways. The first is direct, mechanical interference with the growth of calcium
carbonate crystals. The shape of the PO43 group is somewhat similar to the
shape of CO32, and this similarity enables a phosphate to take the place of
carbonate on the surface of a crystal of growing calcium carbonate. However, it

30

is just enough unlike carbonate to prevent any further crystal growth on top of it.
Chemists call this type of mechanism competitive inhibition: lime scale
crystal growth is inhibited when polyphosphate ions compete with carbonate
ions for the same Ca++ sites on the surface of growing scale; when the polyphosphate wins, scaling is stopped at that point for several hours or days,
depending on temperature. At boiling water temperature, polyphosphates break
apart or revert to single PO4 units, or ortho-phosphate in less than an hour.
If the hardness level is very high, it may cause precipitation of calcium orthophosphate, thus negating the original reason for using a polyphosphate.
The other mechanism of lime scale inhibition by polyphosphate feed is called
electro-static dispersion. There are millions of tiny particles of all kinds in
waternot just dirt, but also tiny scale particles that use a speck of dirt as a
nucleus to crystallize upon and growand the vast majority of them carry a
negative electrical charge. That means they all naturally repel one another to
some extent, because, just as in magnets, opposites attract and likes repel. But
there is a lot of violent activity in water on a microscopic scale: molecules and
particles in water are involved in some million-billion-trillion or an estimated 1027
collisions per second. Sometimes these particles collide with enough momentum to overcome the repulsion, and then they stick together to form a larger
particle. That is one of the ways scale and other kinds of sediment grow: by
agglomeration. Fortunately, the repelling force can be magnified by dosing the
water with dissolved ions having many negative charges, and that is exactly what
polyphosphates do. The smallest polyphosphate has a -4 charge; many have
charges of -20 or -30 or even more. They cluster around particles and add their
charges to make them repel each other thousands of times more strongly than
before, with the result that very few are able to overcome the repelling force and
become agglomerated into larger masses that deposit as scale or sludge.
The concentration of polyphosphates needed to be effective is less than 10 ppm,
and it does not depend on the hardness level. Neither the competitive inhibition
mechanism nor the electrostatic dispersion mechanism operates chemically,
with one molecule reacting with another molecule one-on-one. Instead, the
presence of small amounts of polyphosphate creates a non-scaling environment
by physical means, and the overall effect is called threshold treatment. Higher
polyphosphate concentrations of 20-100 ppm usually exceed the threshold of
calcium polyphosphate precipitation, producing massive amounts of phosphate
scale and sludge. Still higher concentrationsup to 500 ppm or morecreate a
soft water by sequestering every hardness ion with a polyphosphate molecule
as a complex ion which remains dissolved. Examples of this approach are
high-phosphate laundry detergents and bath salts, but it is never used for
drinking water because such high levels of polyphosphates can cause diarrhea
in sensitive persons.
If the water has been pre-filtered by an efficient fine-filter, there will be very few
particles present at all, and the enhanced repelling force will be even more
effective. Everpure research has shown that fine-filtration alone can reduce

31

scale by half or more, just because scale includes a lot of dirt, iron floc, and other
particulate matter in addition to calcium carbonate crystals. Pairing efficient finefiltration with threshold polyphosphate treatment can reliably reduce scaling, and
therefore maintenance costs for ice-makers and coffee brewers, by more than
80%. This patented technology is embodied in Everpure InsurIceTM products for
ice makers and is effective for water supplies having up to about 15 grains per
gallon (GPG) or 250 ppm hardness as CaCO3. Above 20 GPG hardness,
dealkalization usually gives better results.
G. Brackish Water and Excessive Total Dissolved Solids (Secondary MCL:
500 ppm; 1000 ppm in California)
Water can acquire excessive dissolved mineral content by evaporation, as in the
case of Colorado River water being directed through the desert in aqueducts to
Southern California; by flowing through deposits of soluble salts, as in the true
alkaline waters that are full of sodium sulfate or sodium carbonate; and by sea
water intrusion, as is seen in all coastal areas. These waters are not merely
unpalatable; they are actually unhealthful until the excess dissolved material is
removed. In an extreme case, such as drinking sea water, it can be fatal. That
is because of osmotic pressure, which deserves its own discussion.
Osmotic Pressure is a kind of fluid pressure that develops when solutions of
different strength are separated by a semi-permeable membranesuch as the
membranes inside our bodies and surrounding all of our cells. [Interestingly,
osmotic pressure is one of the few phenomena that depends only on the total
number of particles (including atoms, ions and molecules) in solution, and not
their size, or type, or chemistry. A tiny hydrogen ion exerts the same effect as a
huge protein molecule or microscopic particle of scale. The only other
phenomena with this characteristic are freezing point, boiling point, and vapor
pressure.] A semi-permeable membrane may be of the osmosis type (which
allows only water to pass through) or the dialysis type (which allows only
dissolved ions to pass through), or it may actually be a living membrane able to
choose specific materials. All are based on diffusion, which is a fundamental
physical process in which atoms and molecules that are concentrated in one
place tend to spread out and equalize the concentrationlike a drop of perfume
permeating a room. The drive to equalize concentrations is fundamental in
nature, and this is the force behind osmotic pressure.
When solutions of different strength are separated by an osmotic membrane, the
system attempts to equalize the concentrations, but the only mechanism
permitted is to allow water to move one way or the other (thats the definition of
an osmotic membrane). The only way movement of water can equalize the
concentrations is by moving through the membrane in the direction of the more
concentrated solution, as if in an attempt to dilute it. (If water went the other way
the concentrated side would only get more concentrated.) Thus, to look at it
another way, concentrated solutions have the ability to draw water across

32

membranes toward and into them. A concentrated solution exerts a real force
upon a dilute one if theyre connected by a semi-permeable membrane, and the
magnitude of the force is a direct function of the TDS. Therefore, if we drink
something that is more concentrated than our bodily fluids, like sea water, it
draws moisture out of our tissues and into the intestines, causing diarrhea and
making us more dehydrated than before. The only way to reverse the process is
to apply pressure to the concentrated sidea pressure greater than the osmotic
pressure the solution already hasand that is reverse osmosis. Water with a
TDS of 2000 ppm exerts an osmotic pressure of about 20 psi, and that is therefore the usual limit of water quality for RO systems that use only the 40-60 psi
line pressure provided by the water works. Water with more than 2000 ppm
TDS requires pumps and high-pressure housings for RO to be cost-effective.
The ion exchange process for reducing TDS is called demineralization. It
requires two types of ion exchange media: a cationic resin or zeolite in
hydrogen form (regenerated with acid, so that the exchange sites are loaded
with H+ ions); and an anionic resin in hydroxide form (regenerated with strong
base, so that the exchange sites are loaded with OH ions). Put them together
and they exchange everything with an electrical charge for either H+ or OH,
which then combine to form water. However, molecules that are not ionized will
not be removed. Examples: organic solvents classed as volatile organic
chemicals (VOCs) such as trichloroethylene and benzene, most insecticides and
herbicides, most of the organic taste & odor compounds, and a host of other
organic chemicals both natural and synthetic. Demineralization leaves in a lot
that both reverse osmosis and distillation can remove.
H. Turbidity (Primary Standard: less than 1.0 NTU generally, or 0.5 NTU where
direct filtration is used, 95% of the time during disinfection; maximum of 5.0
NTU at any time)
Turbidity is cloudiness or haziness in water, caused by a dispersion or scattering
of light by dissolved particles that are the same size as the wavelengths of light
used to illuminate them. (The visible part of the light spectrum includes light with
wavelengths from about 0.40 microns for violet light to about 0.77 microns for red
light.) It is measured by analyzing the scattered light intensity at a right angle or
90 to the path of the test light beam. This technique is called nephelometry,
and that is the source of the N in the NTU unit of turbidity.
The particles that cause turbidity may be clear or opaque, light or dark colored,
crystalline or amorphous, made of various chemical compositions. Some
particles may refelct light; others may absorb or refract the light, and it is therefore impossible to convert turbidity values to concentrations of sediment in mg/L.
However, most turbidity is just dirt and dust, composed of alumina or silica or
alumino-silicatethe most prevalent minerals in the Earths crust and the stuff of
nearly all rocks except limestone. Thus, an analysis of turbid water nearly
always shows aluminum, and its difficult to distinguish between this naturally-

33

occurring aluminum and aluminum floc (from the coagulation step in the
treatment train used by large water works) that has gotten past the filters. That
is unfortunate, because blobs of floc may well contain dangerous levels of toxic
or infectious sediment. (That is the purpose of flocculation/coagulationto
concentrate turbidity and dirt by agglomeration so they can be removed more
efficiently by the large granular bed filters.)
Turbidity is the only contaminant that must be determined daily (every four hours
at a minimum; preferably, continuously in real time) by waterworks operators,
because low turbidity is necessary for effective disinfection. Pathogens are
easily shielded from the effects of disinfectants by particles and sediment in the
water. But the turbidity measured at the treatment plant bears no resemblance
to the turbidity that exists at the far reaches of the distribution system. There,
many decades of sedimentation, biological growths, and corrosion products
accumulate, become encrusted, and slough off again to produce a steady supply
of new particles which clog filters, score valves, and cause foul tastes and odors.
This is the turbidity that matters to our products.
I: Taste & Odor (Secondary Standard, Odor: <3 TON [Threshold Odor No.])
Most taste is actually odor, and it was found that a Threshold Odor Number of 3
(a 3 : 1 dilution with odor-free water makes the bad T&O disappear) is the
maximum most people would tolerate before abandoning a water supply
(perhaps in favor of a less safe supply). The predominant bad T&O in potable
water these days is that of chlorine in some form. Second is the very common
musty-earthy-moldy-mildewy-fishy T&O produced by certain algae, bacteria,
and molds. Other fairly common offenders are iron and sulfur compounds
already discussed above, other strange T&O occasionally produced by microorganisms, and the strong medicinal T&O of chlorophenols that sometimes
plagues systems using marginal disinfection practices. Each of these will be
discussed below.
Chlorine T&O is familiar to everyone, but there are important differences. All
odorous substances are, by definition, volatile molecules, and that means
elemental chlorine gas, or Cl2 when we smell Free Available Chlorine or FAC.
That includes the hypochlorous acid molecule, HOCl, which is part of the
equilibrium. These have a rather pure smell when compared with the heavy,
dull smell of monochloramine, which is a form of chlorine intentionally produced
by many water works trying to limit the production of THMs. And since monochloramine is much weaker as a disinfectant than FAC, it is often used at much
higher concentrations, making the T&O even worse. It is produced by adding
ammonia (NH3) to water that already has a residual of FAC, and if the wrong
proportions are used or the pH is off, dichloramine may also be produced.
Trichloramine or nitrogen trichloride exists only at very low pH, so it can be
ignored. These chloramines can be recognized as the smell of a poorlymanaged swimming pool, which has a steady input of ammonia from the

34

perspiration and urine of swimmers. If the chlorine level does not keep pace with
the ammonia influx, the pool area will stink of chloramines, which is much worse
than the smell of ordinary free chlorine.
The same is true of drinking water, where chloramines may also be unintended.
When raw water is first chlorinated, ammonia and several naturally-occurring
amines react first; if there isnt enough chlorine to oxidize them completely,
disinfection suffers. As more chlorine is added, the oxidation of the nitrogen in
the amines progresses from mono- to dichloro- to trichloramine, and finally to
nitrogen gas, which bubbles away and is lost. The last reaction is known as the
breakpoint reaction because that is the one that destroys the last of the
chlorine demand and permits FAC to persist so that actual disinfection can
begin. The chemical reactions for these chlorine species are all given below:
Free Available Chlorine
Cl2 + H2O HOCl + H+ + Cl OCl + 2 H+ + Cl
chlorine
hypochlorous
hypochlorite
acid
ion
Combined Chlorine
Cl2 + NH3
NH2Cl
(sometimes intentionally produced,
chlorine ammonia monochloramine
sometimes not)
+ Cl2 NHCl2 (never intentional; very strong smell)
dichloramine
+ Cl2

NCl3
nitrogen
trichloride
+ Cl2

(only at pH < 3)

N2 + N2O + HCl
nitrogen nitrous
gas
oxide
(the Breakpoint Reaction)

Earthy-musty-moldy-mildewy-fishy T&O is well known to everybody, even if they


dont associate it with water. It is surprising that molds are not the only source,
but algae and certain bacteria also produce some of the same chemical
compounds. The differences in smell may simply be due to differences in
concentration and mixtures. For example, one of the compounds that smells
musty in the concentrations found in water smells likeis the aroma of green
bell peppers at lower concentrations. Theyre exactly the same molecule. These
compounds get into drinking water most often from algae blooms in the source
water, which are usually seasonal. Filamentous bacteria called actinomycetes
are also common causes, but their growths in reservoirs are more continuous
and seldom cause unexpected trouble. Actual mold growth is less common and
occurs mostly in long-abandoned plumbing or dead-end mains that have not
received fresh, chlorinated water for a long time.

35

Further to the discussion of sulfur water (which see, above): One of the
bacterial families that is able to reduce sulfate ion (SO42) to sulfide ion and
hydrogen sulfide (S2 and H2S) (rotten egg smell) can also produce spores,
which are tough survival forms similar to protozoan cysts, only smaller. The
bacteria are anaerobic, meaning they cannot tolerate dissolved oxygen, and
they grow only in unused, dead-end mains or other nooks and crannies in the
distribution system where all the oxygen has dissipated. They produce spores to
survive exposure to oxygen if it ever returns. When water treatment equipment
with good mechanical filtration ability is used on supplies carrying these spores,
the spores can become lodged in the media. If the equipment is not used
enough to keep the water inside them fresh and oxygenated, the spores can
grow into active bacteria and begin producing rotten egg smell. It is usually best
to replace or re-bed such media, but it may be possible to vend them often
enough to kill any of the new bacteria before they can make new spores.
Other anaerobic bacteria can make a different kind of foul T&O when oxygen is
absent, but they do not produce spores and are easier to get rid of. The smell
they produce is usually characterized as septic, referring to odors from sewage
that are different from pure hydrogen sulfide. It is not necessarily sewage
bacteria that are responsible for this problem in filters; many anaerobic bacteria
can make the same odorous compounds.
Chlorophenols are intensely offensive byproducts of disinfection that occur only
when the concentration of disinfectant is marginalthat is, barely sufficient to kill
pathogens, but without leaving much of a continuing residual to fight any subsequent contamination or chemical impurities. The T&O is strong, bitter, and
iodine-like and is usually described as medicinal. There are several; the worst
is called 2,4-dichlorophenol and has an odor threshold in the low parts-per-trillion
range. The chemical phenol is a six-carbon ring (a benzene ring) with an -OH
group attached somewhere. The huge molecules of tannins and lignins that give
the brown-yellow color to swamp water (and also to tea) are highly phenolic in
nature, meaning there are a great many benzene rings with one or several OH
groups attached, all linked up into an irregular polymer chain with a molecular
weight in the millions. These tannin and lignin molecules are the main source of
THMs and other disinfection byproducts when the disinfectant (chlorine, chlorine
dioxide, or ozone) chops them into little pieces and then chlorinates the
fragments. The fragments that are, or contain, a phenol quickly become
chlorophenols, which may easily be further oxidized (to destruction), but only if
there is sufficient chlorine or ozone. A chlorophenol problem can also arise if
water-using equipment contains phenolic materials such as PPO (polyphenylene-oxide)-based plastic or rubber compounds that let phenol and
derivatives leach into the treated water, and the residual chlorine level is very
low. The foul T&O may occur only randomly or rarely and can even plague
equipment already protected by filtration systems. All it takes is a few ppt (parts
per trillion) each of chlorine and a phenol, both of which are undetectable at that

36

level by any standard analytical method.


J. Color (Secondary Standard: 15 Color Units)
Color in water refers to the well-known yellowish-brownish tint caused by the
presence of humus, a complex and variable collection of organic molecules
derived from rotting wood and other plant materials containing cellulose, humic
acids, tannins and lignins. It is common in surface waters wherever there is a
lot of dead vegetation, as in southern swamps, northern bogs, and areas with
logging/paper pulp operations. The final product is the result of centuries of
microbial transformations, and every molecule is different, but they are all large
polymers of benzene rings with hydroxyl, carboxyl, and methoxy groups such as
this illustrative but hypothetical example proposed by Christman and Ghassemi
[Chemical Nature of Organic Color in Water, JAWWA 58:6, 722-741,1966]:

Color is measured by comparison with a special mixture of cobalt and platinum


salts, called the Hazen Standard, which has that tint. 15 Color Units is very
palejust barely discernable without side-by-side comparisons. Some can be
removed by activated carbon, but only with difficulty. Treatment with lime or
other source of calcium ion will precipitate much of it, for removal by finefiltration. Flocculation and coagulation with alum is even better. RO, UF and
NF membranes remove color very well.
Color and Taste & Odor : Many odorous molecules are produced as fragments
of color molecules after they are attacked by chemical disinfecants. In addition
to the phenols, these include chemicals with names like geraniol (which smells
like geraniums), pinene (from pine trees), camphor, cinnamic alcohol, vanillin,
etc., which can produce some very exotic mixtures of odors.
Color and Sediment : The vanillin that comes from humus is the same as the
vanillin from oak barrels that is so prized in aged wines and spirits. Those
products also contain other high molecular weight, humus-like molecules that
give an amber hue to brandy and whiskey. The largest of them will precipitate
with iron, manganese, calcium, zinc, and other metal ions, which means that
highballs made with very hard water may produce ugly sediment in the glass.
Cheap liquors whose color comes from charred sugar (caramel) instead of aging
in wood do not produce such sediments.
K. Contamination by Toxic Organics (Various Specific Primary MCLs)

37

Currently there are 57 organic chemicals out of the 83 contaminants regulated


under the U.S. Safe Drinking Water Act, with MCLs ranging from 10 mg/L for
xylene to only 0.00000003 mg/L for dioxin. It is beyond the scope of this Short
Course to discuss each one, or even to try to classify them all in terms of toxicity
or solubility, but it is important to know about the sources of many of them and
whether they are adsorbable to activated carbon.
Agricultural Pesticides: Half (27) of them are insecticides and herbicides, and
most of those are chlorinated hydrocarbons which are quite insoluble, and
therefore highly adsorbable to particles and to activated carbon. Those
containing nitrogen (amines) and phosphorus (phosphates) are more soluble and
less adsorbable, and those containing several oxygen atoms (alcohols [hydroxides] and organic acids with COOH groups) are not well adsorbed to anything
and are very difficult to remove, except by RO.
Industrial Solvents and Byproducts: There are 22 of these. Most industrial
solvents are small, volatile, low-molecular-weight molecules that readily
penetrate RO membranes and adsorb only poorly to activated carbon. Nine
more byproducts of plastics and rubber manufacturing are variable in carbon
adsorption: some are adsorbed extremely well, and others extremely poorly.
But most of these are larger than the solvents and can be removed by RO
membranes.
Disinfection Byproducts (DBPs): This refers mostly to byproducts of chlorine
disinfection, acting on naturally-occurring humus and color bodies, but ozone
can produce the same chemicals because it is a powerful oxidizer that can
change chloride ion to chlorine. At present, only one chemical family is
regulatedthe THMs or trihalomethanes, as exemplified by chloroform, CHCl3.
The MCL is 100 ppb total THMs (TTHMs), but that is expected to change
significantly soon. In addition, two more families of DBPs, the halogenated
acetic acids (HAAs) and acetic acid nitriles (HANs), will be regulated in the
coming years. All are presumed to be carcinogens, and THMs within the
presently permitted levels have also been implicated in human miscarriages.
SECTION III
TYPES OF WATER TREATMENT AND WATER TREATMENT EQUIPMENT
A. Mechanical Filtration and Filtration Equipment
Anything that blocks a fluids path and reduces the cross-sectional area through
which solid objects can pass is a filter. Two logs across a stream will filter out
dead cows. But there is no need to go that farthe word filter comes from
felt, a mat of wool that is a very good filter. In fact, the golden fleece of ancient
Greece was real, not myth: gold dust does adhere to wool, and mats of it used in
those early ore sluices were one of the first industrial uses of filtration.

38

Felt is a good example of one of the two basic types of filter, which is the depth
filter: a mass of many random layers of media forcing the water into a tortuous
path. Solids get entangled and are removed from the flow. The other filter type
is surface filter, such as a fish net or a window screen: a single layer of meshwork which stops everything larger than a certain size, and lets everything
smaller than that size pass. Almost all filters in commerce are depth filters.
Another way to categorize filters is by the size of the particles they can remove:
Designation
coarse filtration
fine-filtration
ultra-filtration
nano-filtration
hyper-filtration

Particle Size
30 um and up
0.1 um and up
500,000 MW and up
1000 MW and up
atoms

Significance
visible dirt
colloids, cysts
proteins, viruses
membrane softening
reverse osmosis

Fine-filtration down to fractions of a micron is the same as micro-filtration, and it


represents the best mechanical filtration that can be accomplished with ordinary
materials and hydraulic designs. These include ceramic candle filters, pressedcarbon block filters, precoat filters, sintered glass/metal/plastic powder filters,
and microporous polymeric membrane filters (usually pleated to maximize
surface area). Removing particles smaller than 0.1 micron at reasonable
pressures and useful flow rates requires abandoning the fundamental concept of
mechanical filtration by physical interference with the flow, and adopting an
entirely new and different paradigm: membrane filtration. Not to be confused
with microporous polymeric membrane filters, the semi-permeable membranes
used for ultrafiltration, nanofiltration, and reverse osmosis systems have no
pores and do not physically remove particles; they exclude them chemically
because the membrane material is formulated to be extremely hydrophilic
(water-loving). Water dissolves into the membrane and makes its way to the
other side by diffusion, while most everything that is not water is passively left
behind. Ultrafilters are generally designed much like conventional mechanical
filters, and those without a continuous waste stream to get rid of accumulations
are subject to rapid clogging. However, RO and nanofilter membrane modules
are designed with a continuous waste stream and tangential flow geometry
which permit use for years without clogging.
A special note on colloids: these are the largest molecules or the smallest
particlesso small that they never settle due to gravity. They are kept in
suspension forever by Brownian motion, the random vibration of microscopic
particles in water caused by collisions with water molecules and other molecules
dissolved in the water. Common colloidal suspensions are blood (both the blood
cells and the protein molecules are colloids), tea or swamp water (tannins and
lignins are colloids), and tap water containing micro-organisms and very fine
turbidity. It is notable that the upper end of the colloidal range coincides with the

39

sizes of the wavelengths of visible light: violet light has dimensions of about 0.44
um; red light, about 0.77 um. This size range also overlaps with that of the
largest particles that can be removed by physio-chemical adsorption rather than
by mechanical filtration. Thus, a mechanical filter efficient in the colloidal range
can benefit from simultaneous adsorption of even smaller particles if the filter
medium is, or contains, an effective adsorbent such as activated carbon. This is
the advantage of precoat carbon filters and carbon block fine-filters made from
powdered activated carbon: they filter mechanically down to the limit of that
technology, and then there is at least the hypothetical opportunity (depending on
flow rate) to remove everything else that is adsorbable, right down to the level of
individual atoms and molecules.
Pressure Drop: Anything put in the way of flowing water causes a resistance to
the flow (by definition), slowing it down and creating a back-pressure known as
the differential pressure, delta-P or P, or pressure drop. It is simply the
dynamic (flowing) line pressure provided by the municipal waterworks or other
pressure source, minus the gauge pressure at the outlet of the filter (or whatever). It is equal to the sum of all the many little resistances along the way from
point A to point B: every centimeter of pipe or tubing, every fitting and valve with
a reduced orifice, every particle or fiber of filter medium, even every change of
direction with an ell or tee adds to the overall pressure drop of a system.
Hydraulic Capacity: The importance or significance of pressure drop is that it
determines the hydraulic life (capacity before becoming plugged with sediment)
of filters. One end-point of mechanical filtration is the appearance of either
particles of the filter media or the filtered solids in the effluent. The force of the
flowing water carries particles toward the effluent, where they eventually break
through if they can. If the pressure and flow are very regular, this can occur on
schedule, and the hydraulic life is somewhat predictable. Usually, however, flow
surges or water hammers cause the media to shift or flex, so that particles
previously removed from the flow (filtered out) are dislodged and stream toward
the outlet suddenly. This is known as channeling and dumping, and once a
filter has had a channeling and dumping event, it is usually finished as an
effective mechanical filter. Some media may be able to repair themselves
momentarily and thus last a long time with repeating cycles of filtering and
dumping, until the medium is either backwashed and rebedded or replaced.
Channeling and dumping is both dangerous and disgusting, and it should not be
permitted. It is the job of filter designers to assure that filtration will not break
down (allow channeling and dumping) until the P exceeds a certain level,
usually 40 psi (276 kPa or 2.72 atm). That is because the minimum line
pressure in the U.S. is usually about 40 psi, which means that when the P is
also 40 psi, there may be very little flow of water. All or nearly all of the pressure
provided is consumed by the system, with little or nothing left over to provide
push to make the water flow. Thus, a P of 40 psi is usually considered the
hydraulic end of life. It should be easy to see that in any system designed to
transport water, a filter in the line will probably be the major contributor to

40

pressure drop. And it is self-evident that:


a), finer-grained media provide better mechanical filtration than coarse
media, and
b), finer-grained media also produce greater pressure drop than coarse
media.
Thus, it should be easy to see that there is a trend from coarse filters with low
pressure drop to fine-filters with higher pressure drop, and that there is a practical limit to the thickness or depth of filter media because of pressure drop. For
granular media, the practical limit is about 30 inches (76 cm) of bed depth: beds
of granular carbon, ion exchange resins or filter sand deeper than that can be
expected to have excessive initial P at the usual flows and pressures. It should
also be easy to see that, as the filter media particle size becomes increasingly
smaller, the pressure drop for a given media depth increases. Or to put it the
other way, when the media particle size is decreased, the media depth must
decrease if the same initial P is to be maintained. This trend is taken to the
logical limit in precoat filters, which are Everpures specialty.
Everpure precoat filters use finely-powdered filter media (mostly powdered
activated carbon, but also including diatomaceous earth) hydraulically deposited
on and retained by a filtration barrier, which is a fabric. Cartridges are shipped
with the Micro-Pure media mix dry and powdery in the bottom. When a new
cartridge is first activated, the incoming water makes a slurry of the media, which
begins to deposit on the fabric septum as soon as the pressure vessel fills with
water and the water pressure begins to force water through. The original design
uses a relatively coarse, woven polypropylene filament fabric that lets some of
the Micro-Pure powder through as initial black water during a run-in or activation, which is required for several minutes. Soon, the particles of media form
interlocking structures analogous to the construction of arches and domes,
creating a stable filtration layer called the precoat cake. Since the 1970s a
much finer non-woven, paper-like polyethylene fabric has enabled production of
precoat filters with no need for an initial activation to purge the initial filter media
fines. These precoat filters all have a media depth of only a few millimeters
less than a quarter-inchspread out over a large, pleated septum with several
square feet of filtration surface area. Their mechanical filtration efficiency is
NSF-Certified as >99.9% at the level of 0.5 um. The large surface area
compensates for the relatively high pressure drop per unit of area, enabling
acceptable overall Ps and flow rates.
The importance of filtration surface area is shown by the hydraulic equation:

where

q2
Pt = K
Ct
A2
Pt is pressure drop at time t (set at 40 psig)
K is a constant specific to the conditions
q is flow rate in gal/min. (held constant)
A is filtration surface area

41

C is concentration of particulate mater (held constant)


t is time in minutes
If the values of all the factors are supplied and the equation is solved for time,
which then converts into gallons or liters, it can be shown that the A2 term,
filtration surface area, so dominates the outcome that a much simpler rule of
thumb can be used to estimate the increase in hydraulic capacity attributable to
a given increase in surface area:
new capacity (original capacity) (filtration surface area increase)2 .
For example, if the area is doubled by using two filters instead of one, the simplified equation indicates that the pair should last approximately 22 or 4 times as
long as one. Three filters 32 or 9 times as long; four 42 or about 16 times
as many gallons before they plug up, and so on. (If this seems too fantastic to
be true, consider that the flow rate for each filter is divided by 2 or 3 or 4 in
multiple filter installations at the same time the area is multiplied, and flow rate,
q2, is also a squared term.) The significance of this relationship is seen in the
following comparison of the Everpure 12-in. precoat filter and comparably sized
(3 in. diam. X 12 in.) granular bed and spool or carbon block filters, all flowing at
0.5 gal/min.:
Filter Type
Granular bed
Spool/Block
12-in. Precoat

Filtration Surface Area


7.1 sq.in.
113 sq.in.
444 sq.in.

GPM/sq.ft.
10.2
0.64
0.16

Granular media always channel and dump before the pressure drop reaches 40
psig, so the hydraulic equation, above, and its consequences do not apply to
them. Also, many spool and block filters are far too coarse to resist breakdown
when the P exceeds 40 psig, and some of them may never plug up. But any
filter that can qualify to claim Cyst Reduction (document 99.9% efficiency at
removing live Cryptosporidium oocysts or 99.95% reduction of 3-4 um particles)
will quickly become a precoat filter as particles from the water accumulate on the
leading surface. Thus, it is proper to compare the expected hydraulic lifetimes of
the hypothetical 3 in. diam. X 12 in. cyst-removing carbon block filter and the
Everpure QC4 filter, using the rule of thumb. The fraction 444/113, representing
their surface areas expressed as a multiple, squared, gives the precoat 3.92 or
more than 15 times the hydraulic capacity of the block. However, in the real
world, these carbon blocks are made to fit standard pressure vessels and are
actually 2.5 in. diam. X 10 in., so the real-world numbers are (444/78.5)2 more
than 30 times greater capacity.
Carbon Block filters can be fabricated by molding or extruding a mixture of

42

activated carbon, other special adsorbents, filter aids, and binders into a stable,
unmovable structure of nearly any size, shape and porosity with a variety of
possible claims. Since the grains of media are held in place by binders, carbon
blocks are much less susceptible to hydraulic failure due to channelling and
dumping than ordinary granular media. However, that also assures that they
cannot avoid total clogging, eventually. The thickness of blocks allows the
entrapment and adsorption mechanisms of depth filtration to dominate the
removal of fine particles, so that they often achieve considerably greater
hydraulic capacity than that predicted by their filtration surface area and the
hydraulic equations discussed above. Thus, the 30 : 1 theoretical advantage of
precoat filters noted in the previous paragraph should be revised to some lesser
factor, say, 15-20 : 1.
Carbon block filters can also have excellent adsorptive ability and capacity, since
they may contain large amounts of media, sometimes exceeding that of GAC
filters of the same size. Thus, unlike precoat filters, they have significant
capacity for reduction of combined chlorine (monochloramine) and VOCs
(Volatile Organic Chemicals) such as THMs and industrial solvents, and it is a
simple matter to include special adsorbents for heavy metals in the media
mixture.
Coarse Pre-filters are used to remove most of the large suspended matter so
that fine-filters will last longer. They may be large granular media filters that are
backwashable, or smaller, less costly cartridge filters comprised of plastic foams,
wound string filters, or fibrous mats of various compositions. They need to be
inexpensive enough to permit replacement often. Almost all of them are
susceptible to channelling and dumping.
Multi-Media Filters use granular media of several (usually three) types that are
arranged in layers according to density and particle size. They are more efficient
than ordinary granular media bed-filters, but they must be backwashed more
often.
B. Adsorption and Adsorption Equipment
As described in Section I, adsorption is a two-step process (movement from the
bulk of the solution to a surface by diffusion, followed by weak chemical bonding
at the surface), and that takes time, which requires a minimum contact time
between the water and the adsorbent to be effective. Contact time can be
increased by using more adsorbent or by reducing the flow rate. The empty bed
contact time (EBCT) is an important engineering property of adsorbers, calculated by dividing the volume of the filter bed by the flow rate (e.g., liters divided by
liters per minute yields minutes of EBCT). Calculations involving it apply mostly
to designing large municipal filters with hundreds of cubic feet of media, and not
so much to the small cartridge filters used for point-of-use treatment. Scale-up
calculations for these are limited to simple, direct proportions with a maximum

43

multiplier or divisor of three. (e.g., if one liter of media lasts for 1000 L of test
water, then three liters of media should last for 3000 L. But larger amounts
should be tested again.) The granular adsorption media used by this industry
are limited to bed depths of 30 inches, or about a meter, because of the
pressure drop they produce at useful flow rates. Thus, there is a limit to the
number of small filter cartridges that can be plumbed in series to achieve better
adsorption performance.
C. Ion Exchange and Ion Exchange Equipment
The first step of ion exchange is identical to the first stage of adsorption
movement of contaminant molecules from the bulk of the solution to the surface
of the media by diffusion and Brownian motionbut the similarities end there.
While adsorption almost always involves only complete, electrically neutral (unionized) molecules, ion exchange occurs only with ionsmaterials having a + or
electrical charge. And while the second stage of adsorption (capturing the
contaminant and removing it from circulation) depends only on the chemical
affinity between the contaminant and the media, the capturing step for ion
exchange depends mostly on the electrical charges. And finally, after capture of
a contaminant, ion exchange is much more easily reversed than adsorption,
enabling many cycles of use and regeneration economically and conveniently.
Ion exchange depends on the existence of special media with inherent,
permanent electrical charges that attract and hold ions with the opposite charge.
Naturally-occurring and synthetic inorganic minerals with this property are called
zeolites, which are special crystalline compounds of silicate and aluminate.
Those chemical groups are both anions (ions with a negative electrical charge),
and therefore zeolites can exchange only cations (ions with a positive electrical
charge) such as Na+ or Ca+2. The ion exchange capacity of zeolites is not very
great, so synthetic organic ion exchange resins were developed to make water
softeners more marketable. Organic ion exchange resins can be formulated with
either anionic or cationic ingredients of many types, yielding products with
nearly every conceivable functionality and strength. Thus, there are resins
formulated specifically to exchange nitrate ion preferentially, or fluoride ion, or
uranium and plutonium from nuclear wastes, and so on. (These special resins
are rare, and therefore very expensive.)
Ion exchange is most efficient when the media is housed in cylindrical beds and
the water flows downward through it. The exchange of ions is very rapid, and
there is a zone or band of activity that slowly moves down the bed. Above the
zone, all of the active sites are exhausted, and below the zone, all are still in
completely regenerated form. The thickness or height of the exchange zone is
determined by flow rate: the slower the flow, the narrower the zone. Just as the
unit (area) flow rate in GPM/sq.ft. or L/min/m2 is the important factor for the
hydraulics of mechanical filtration, the unit (volume) flow rate in GPM/cu.ft. or
L/min/m3 is the relevant parameter for the chemical performance of media used

44

for adsorption, ion exchange, oxidation, etc. The usual unitary flow rate for ion
exchange systems is about 2 GPM/cu.ft. (0.214 L/min/m3 ) or 16 bed volumes
per hour.
1. Softening (Conditioning): Sodium cycle cation exchange is the most
common use of ion exchange, in which hardness (mostly ions of calcium and
magnesium) is exchanged for sodium ions to produce soft or conditioned
water. It is made possible by the availability of an ion exchange resin with
millions of active sites made of the chemical group called sulfonate, which can
be represented as R-SO3 [M+]. The R represents the connection to the polymer
backbone, and the M+ is the balancing cation (called a counter-ion, needed to
make an electrically neutral product that doesnt give a shock when it is touched)
at each of the millions of active sites in each of the millions of little resin beads.
Since the sulfonate is an anion, it can exchange only cations, and it happens that
this particular chemical arrangement produces a strong affinity for Ca+2 and
Mg+2, and a weak affinity for Na+. That means that, given a choice, the resin
would prefer to capture Ca+2 and Mg+2 rather than Na+. (There is no real
preference or choice. Its just a matter of stability and what lasts. Just as
dissolved molecules are involved in a million-billion-trillion or 1027 collisions per
second, attached contaminant molecules or ions are grabbed onto and let go
(sorbed and desorbed) millions or billions of times each second. If there is some
affinity between the two, then the contaminant will stay attached for a greater
percentage of the time, and we can say it has been adsorbed or exchanged, and
the fraction of time spent attached describes the strength of the adsorption or
exchange.) However, if there is no choice, or if the Na+ concentration is much,
much higher than anything else, the resin will load up with Na+ ion on all of the
millions of active sites. That is exactly what is done. The resin is initially treated
with about 5% (50,000 ppm) Na+ + Cl salt brine, and the 50,000 ppm Na+
completely overwhelms the effect of any other cations that might be in the water,
so the resin loads up with Na+ ion. In service, this fully regenerated resin is
then exposed to water containing Ca+2 and Mg+2 ions. The resin exercises its
preference and exchanges two Na+ ions for each Ca+2 or Mg+2 it encounters, as
long as there are any Na+ ions left attached to the resin. When theyre all gone,
the resin is said to be exhausted and ready for a new regeneration with a fresh
dosage of 50,000 ppm of salt brine. The resin still prefers to have Ca++ and
Mg++ attached rather than Na+, so the strong brine solution needs to flow through
the resin bed for 15-30 minutes, continuously. This assures that all of the
calcium and magnesium ions on the resin have the opportunity to be displaced
by the overwhelming presence of sodium ion. Such timing and the necessary
initial back-washing and final rinse of the bed are done automatically by the
water conditioner equipment package, so there is no need to go into detail here.
It is sometimes said that softening does not change the overall TDS (Total
dissolved Solids) of the water. That is not exactly true. The change is small, but
there will be a slight increase due to the atomic weights of the elements that are
involved, depending on the water supply. The A.W. of calcium is 40; that of

45

magnesium is 24; and sodium, 23. Thus, when one calcium is exchanged for
two sodium ions, their contribution to the TDS is 46, which is 15% more than the
40 they replaced. If it is magnesium that is exchanged for two sodium ions, the
change would be 48 versus 40, or 20% more. Most water supplies have considerably more calcium than magnesium, with the average composition being
probably about 2/3 calcium and 1/3 magnesium. That would produce an
average atomic weight of about 35, compared with 46 for the two sodiums.
Thus, the TDS of a hard water might increase as much as 30%, from, say, 350
mg/L to 450 mg/L after being softened.
Softener resins special affinity for Ca++ and Mg++ is actually for divalent cations, which means that any other member of their chemical family (the rest of
column 2 of the Periodic Table of the Elements: strontium, barium, and radium),
plus a great many other metal ions with two charges, will also be efficiently
removed by water conditioning. Removal of radium and its radioactivity is a very
important ability, as is removal of moderate concentrations of manganese and
iron. Softeners also remove lead, but that contaminant comes mostly from
plumbing materials that come after the softener.
2. Dealkalization is the addition of acid in some form, to drive the alkalinity
equilibrium to the left (see the Chemical Equilibrium discussion in Section I), so
that some of the alkalinity is lost as CO2. To do that with ion exchange requires
loading the resins (or zeolites) active sites with H+, which is the embodiment of
acidity. That is accomplished by regenerating the bed with a 5% acid solution
(50,000 ppm H+) instead of a salt brine. The resins active sites have a
preference for Ca++ and Mg++ and other divalent cations, but the high concentration of H+ ions overwhelms the preference and eventually removes all of the
calcium and magnesium if the regeneration lasts long enough.. During the
service cycle, the resin exercises its preference and readily exchanges each
Ca++ and Mg++ ion in the water for two H+ ions. These affect the alkalinity
equilibrium instantly and cause some to be lost as CO2 gas. Thus, the overall
result is removal of both hardness and alkalinity, and a reduction in TDS as well.
Dealkalization can be done with the same type of ion exchange resin as that
used for softening, called strong-acid or strongly acidic cation exchange resin,
but that is risky: when there is more hardness in the water than alkalinity (a
common condition), more H+ will be produced than there is alkalinity to be
neutralized, resulting in an acidic, corrosive, over-treated water. Therefore, it is
more prudent to use a different type of resin, called weak-acid or weakly acidic
cation exchange resin, sometimes referred to as WAC resin, for Weak Acid
Cation. The difference between them is the strength to overpower the effects of
weak acids, or buffers, in the water. (See the discussion of buffers in Section I.)
When a WAC resin is regenerated with acid and used to exchange H+ for hardness ions, it works only as long as alkalinity remains available to neutralize the
H+. If the water has more hardness than alkalinity, WAC resins in H+ form work
only up to the point of destroying all of the alkalinity, and then they stop. For

46

example, in treating a water supply with 300 ppm of hardness as CaCO3 but only
250 ppm of alkalinity as CaCO3, WAC resin would exchange only 250 of the 300
ppm, leaving 50 ppm of hardness and zero alkalinity, with no risk of over-treating
and creating a corrosive condition. If the water has more alkalinity than hardness, WAC resin will exchange all of the hardness and neutralize the same
amount of alkalinity.
There is also a third ion exchange process for dealkalization, and that is using a
strong-base or strongly-basic anion exchange resin in chloride form. It is
regenerated with NaCl salt brine just like softening; the difference is that the
resin exchanges two Cl ions for CO32 ions instead of two Na+ ions for Ca++ or
Mg++ ions. The system is set up and run exactly like a softener. This is hardly
ever done, because softener resin is among the cheapest and strong-base anion
resin the most costly media available.
3. Demineralization or deionization is the removal by ion exchange of all ions
in solution, both cations and anions. That requires use of both cation exchange
resin in H+ form and anion exchange resin in OH form. That means that the
cation resin is regenerated with acid (H+ ion), and the anion resin is regenerated
with strong base or OH ion. The cation resin exchanges all cations for H+ ion,
and the anion resin exchanges all anions for OH ion. The two combine to form
water, with nothing left over, so that the treated water is similar to distilled water
quality. It is not identical to distilled water, because the ion exchange process
does not remove electrically neutral molecules such as sugars, alcohols, many
pesticides, VOCs, and other un-ionized organic compounds which may be
present.
The two resin types may be combined in a mixed bed or in separate beds or
cartridges. It is difficult (but possible) to regenerate a mixed bed, because the
two resins must be separated first. This is done by fluidizing the bed (backwashing, or operating up-flow) so that the resin types separate by density. Once
separated, they are regenerated separately and then re-mixed. The amounts of
the resins used must be chemically balanced, so that they both become
exhausted at the same time.
Both the anion and cation resins may be either strong or weak-acting resins, but
the usual (most economical) combination is a strong-acid cation resin paired with
a weak-base anion resin. Since a weak-base anion resin cannot exchange the
anions of weak acids unless the conditions are already acidic, the strong-acid
cation resin in H+ form is always put first. To illustrate: a weak-base anion resin
in OH form will not exchange HCO3 and CO32 ions unless there is free H+
already available to neutralize the OH. (This is analogous to the situation with
WAC resin [see above, under Dealkalization], which will not exchange H+ for
hardness unless sufficient alkalinity is already available to balance it.) If the
water encounters the cation resin loaded with H + first, then the water will have
exchanged all cations for H+ already and be strongly acidic when it enters the

47

cartridge/bed containing the weak-base anion resin.


D. Oxidation filtration
There are two kinds of filter media which can be used in beds or cartridges to
oxidize ferrous iron, manganese and sulfide ion/hydrogen sulfide in water. (See
the Iron Water and Sulfur Water discussions in Section II.) One is a zeolite
called manganese greensand, either natural or synthetic, loaded up or
exchanged with Mn++ ion and then oxidized (regenerated) with chlorine or
potassium permanganate (KMnO4). The Mn++ on the zeolite becomes MnO2,
called manganese dioxide, which is a good oxidizing agent. When water
containing Fe++ or Mn++ or S2 flows through a bed of this manganese greensand
in MnO2 form, they are readily oxidized to Fe+++ or MnO2 or S, respectively,
which are all particles for a fine-filter to remove. At the same time, the MnO2 on
the zeolite is reduced back to its original Mn++ ionic form, which must then be reoxidized (regenerated) back to MnO2 again in a new cycle. (Carefuldont be
confused by the fact that were using manganese to remove manganese and
iron and hydrogen sulfide. The Mn++ in the water is oxidized to MnO2 and
removed by filtration later. The MnO2 on the zeolite gets reduced to Mn++ in the
process, but that Mn++ is supposed to stay attached and be recycled back to
MnO2.) When manganese greensand is used to oxidize Mn++ ion or Fe++ ion or
S2 ion in water, the reactions are:
Zeolite-MnO2 + Mn++ (dissolved) Zeolite-Mn++ +

+ Fe++ (dissolved)

+
2

+ S (dissolved)

MnO2
Fe+++ [Fe(OH)3 ]
S

All three end-productsthe MnO2, Fe(OH)3, and Sare filterable particles.


Also do not be confused because one of the possible regenerant chemicals is
another form of manganese, potassium permanganate. Its just a coincidence
that three different forms of manganese can be involved in this technology.
If a bed of manganese greensand is allowed to become completely exhausted,
the Mn++ attached by ion exchange to the zeolite may be lostexchangedand
require being regenerated by ion exchange before oxidizing it to MnO2 for
another cycle of oxidation filtration. Therefore, it is important that manganese
greensand installations be regenerated promptly on schedule, rather than waiting
until the oxidizing function begins to fail. By then, some of the MnO2 will already
be gone from the zeolite, and peak efficiency will be lost.
The other redox medium is copper-zinc alloy (brass). It was realized only
recently that one reason disinfectant chlorine dissipates so quickly in the
plumbing overnight is that the chlorine actively corrodes (oxidizes) the surface of
copper pipe and brass fittings, which then liberate ions of Cu+ or Cu++ or Zn++
into the water. (This is also the source of nearly all lead contamination in

48

watercorrosion of brass and solder containing 8% or more lead, used to make


the materials machinable and less brittle. Granular brass filter media is made
from pure alloy, so there is no lead hazard.) The presence of copper and zinc
atoms next to each other in an alloy allows an easy exchange of electrons, and
there is a natural tendency for electrons to flow from zinc to copper. If the two
were present as separate bars of zinc and copper metal connected by a wire, as
in the example in the discussion of Oxidation and Reduction in Section I, the flow
of electrons would do work, such as lighting a light bulb. In an alloy no wire is
needed, but electrons do not flow spontaneously because there is no place for
them to go. But if the alloy is immersed in water containing contaminants like
chlorine that can be reduced, the contaminants accept the electrons from the
zinc, which then becomes Zn++ ion in solution. Later, further down in the bed of
granules, the Zn++ ions tend to precipitate out as zinc oxide (ZnO) or hydroxide
[Zn(OH)2] and slowly coat the alloy, leaving it less and less effective. This can
occur slowly or rapidly, depending on the pH, the amount of chlorine, and the
level of dissolved oxygen in the water. Some Cu++ ions may be liberated, but
this again depends on the levels of chlorine, oxygen, and acidity. The net result
is a low but acceptable concentration of both Zn++ and Cu++ ions in the treated
water. The U.S. limits are 5 ppm for zinc (recommended) and 1.3 ppm for
copper (mandatory).
The exact electrochemical mechanism is not quite so clear for oxidation of Fe++
or H2S by granular brass filtration. However, the voltages for both of these two
oxidations lie between those for copper and zinc in the Electromotive Series
(which see, Section I), which means that electrons can go either way (cause
either oxidation or reduction) between them. The reduction of hydrogen sulfide
is said to proceed by oxidation of copper and zinc, which then precipitate the
sulfide as ZnS and CuS for later removal by mechanical fine-filtration.
E. Chlorination, Coagulation, and pH Adjustment
Chlorination, coagulation, and pH adjustment are very different activities, but
they are often accomplished simultaneously using the same equipment: a
chemical feed pump with its reservoir of concentrated chemicals, followed by a
contact/mixing tank. If the oxidant/ disinfectant is liquid chlorine bleach (5.2518% sodium hypochlorite, NaOCl), the coagulant is sodium aluminate (NaAlO2),
and the pH reagent is sodium hydroxide (NaOH, caustic soda) or sodium
carbonate (Na2CO3, soda ash), they can all be mixed together and fed with a
single pump.
The preferred process is a variation of standard practice known as superchlorination-dechlorination because a large excess of chlorine is used (about
10 times the usual dosage) and then removed before drinking. It is based on the
concept of the CT envelope, which states that the arithmetic product of one
combination of the chlorine concentration in ppm, C, times the contact time in
minutes, T, is equivalent in disinfection ability to any other combination that

49

produces the same total. For example, 30 minutes of contact time with 0.5 ppm
FAC is equivalent to 5 minutes at 3 ppm, because they both total a CT of 15,
which is the minimum for less-than-ideal conditions. Large waterworks have the
time and space to use low chlorine concentrations with long contact times, but it
is more economical for small systems to use relatively small contact tanks with
ten times as much chlorine as usual, and then waste 95% of it. At a flow rate of
5 gal/min., a 5-minute contact time would only require a 25-gallon tank (if the
mixing were perfect), while a 30-minute contact time would require a 150-gallon
tank. (Correction for mixing efficiency changes those figures to 89 gallons and
536 gallons. Details on page 55.)
Sodium hypochlorite solutions are made by bubbling chlorine gas up through a
solution of sodium hydroxide:
(not an equilibrium reaction)
Cl2 + NaOH H+ + Cl + NaOCl
The resulting solution is pH 10-12, depending on the local producer, and the free
chlorine is present as the sodium salt of hypochlorous acid, which is a weak
acid with a KEq = 107.5. However, NaOCl itself is stronga simple salt like
NaCl that is 100% ionized the instant it is dissolved:
NaOCl

Na+

H2O

+ OCl

(not an equilibrium reaction)

Only the acid, HOCl, is weak:


OCl

H+

KEq = 107.5

HOCl

That implies two things: first, that when the chlorine bleach is injected into the
water, it will instantly change from being all hypochlorite ion, OCl, in the bleach
to being mostly in the form of the un-ionized hypochlorous acid molecule, HOCl,
in the water. Second, hypochlorous acids KEq is close enough to the H+
concentration in neutral water to interfere with the pH of the water into which it is
being injected. Actually, because of the small amounts involved, its the other
way around: the pH of the water interferes with the hypochlorous acid equilibrium. This is extremely important, because only the HOCl molecule disinfects.
All forms of free chlorine are equally effective as oxidizing agents for iron and
manganese and hydrogen sulfide, and OCl ion can kill some viruses, but if the
pH is too high to allow most of the HOCl to exist as a neutral molecule that can
penetrate cell membranes, there will be little disinfection of bacteria. The 50/50
point in the dissociation of hypochlorous acid occurs at pH 7.4, and if the pH
rises to 8.5 the HOCl will be 95% dissociated and useless as a bactericide.
Therefore, it is important not to raise the pH any higher than necessary if the
chlorination system is expected to disinfect as well as oxidize.
Sodium Aluminate is a source of alum, or aluminum salts that form a fluffy
precipitate called floc when dissolved in water at ordinary water pH. If the
concentration is high enough, the pH is right, and there is enough gentle mixing

50

to promote agglomeration of the precipitate into larger clumps which entrap


particles of dirt and turbidity (called coagulation), the water can be greatly
clarified without the use of fine-filters. In large waterworks there is lots of space
and time, and after flocculation and coagulation the floc is allowed to settle, and
virtually all of the suspended particles sink to the bottom enmeshed in a solid,
jelly-like mass. The clear water is then sent to large granular media filters, but
the filters only have to remove the occasional clump of floc that gets stirred up.
Everything else has already been removed as sludge. Small, individual water
supply systems do not have the luxury of acres of space for water treatment, so
there is no settling phase, and the final filters are more important.

Aluminum is somewhat unusual in its ability to precipitate as a hydroxide and


then to re-dissolve as a complex ion with the addition of more strong base:
Al+++ + 3OH Al(OH)3 (solid aluminum hydroxide floc)
+ OH AlO2 (soluble aluminate ion) + H+ + H2O
Thus, if a solution of sodium aluminate (NaAlO2) is added to water with a
normal pH (near pH 7) or lower, the aluminate ion will convert to the Al(OH)3
form and precipitate, or flocculate as a fluffy, gooey solid material that entraps
colloids and particles. Sodium aluminate can be purchased as a 10% solution or
prepared by dissolving any aluminum salt in water and then adding NaOH until it
first precipitates and then redissolves.
Note that both the chlorine bleach (sodium hypochlorite) and the flocculant
(sodium aluminate) are produced as sodium hydroxide (NaOH) solutions. Both
chemicals will affect the pH only slightly, because they are added only at low
ppm levels. However, more NaOH can be added to either at any time. That
means that the final pH of any water supply that is treated with either or both
chemicals can be raised to any desired level by adding more NaOH, and they
can all be mixed together and fed as a single solution. As an alternative, sodium
carbonate (Na2CO3, soda ash) may be used to raise the pH instead of NaOH.
This is sometimes a good idea, since NaOH is so dangerous to use, while
sodium carbonate is relatively innocuous. However, since it is not as strongly
alkaline, Na2CO3 may not be strong enough to keep sodium aluminate from
precipitating too early, before it is fed, thus clogging the pump. This is seldom a
problem except when the brand of bleach locally available is deficient in NaOH.
So, if addition of sodium aluminate to chlorine bleach causes immediate precipitation, NaOH is the only recourse. Preparing such a mixed chemical feed
solution is unfortunately rather complicated. The best way is to treat each
ingredient separately in a jar test as described below, and then combine them
in the indicated ratio. The setting on the feeder pump is then determined by trial
and error.

51

Determining the Chlorine Dose: You will need a gallon jug, an ordinary bucket,
an eye-dropper, and a chlorine test kit that measures free chlorine (not total
chlorine). Put a gallon of the water to be treated into the bucket and start
adding chlorine bleach drop by drop, counting the drops. After 5 drops, mix and
wait 5 minutes, and then test for free chlorine with the test kit. The 5-minute wait
is to allow time for the chlorine to react with any iron, sulfide, ammonia, organics,
whatever, that may be present and acting as oxidant demand or chlorine
demand. The demand, if any, must be destroyed before any extra chlorine left
over (called the residual) can be made available to do any disinfection. If the
demand is not destroyed, the chlorine that would be measured with a total
chlorine test would include any combined chlorine consisting of monochloramine, dichloramine, and organic chloramines derived from amino acids, etc.,
which are all very poor disinfectants and also very weak oxidizers. (They are
very stable and last a long time in the pipes, which is one reason monochloramine is sometimes intentionally produced by large water works with large
distribution systems. But that is done only after effective disinfection with free
chlorine has already been completed.) The goal in the small chlorinator systems
used for private wells is a concentration of 3-5 ppm free chlorine. If its already
higher than that after 5 drops, dump the bucket and start over with a diluted
bleach solution. Repeat the process of adding drops of bleach, mixing, waiting 5
minutes, and testing as many times as it takes to find the total number of drops
of bleach needed to destroy the demand and produce a residual of 3-5 ppm free
chlorine after a five-minute contact time. Record the number of drops.
Determining the Sodium Aluminate Dose: Using the same gallon jug, bucket and
eye-dropper (cleaned up, of course!), add and count the drops of sodium
aluminate needed to coagulate all of the turbidity or color in the water. After
every 5 drops, stir very strongly for about 30 seconds, then back-stir for a
moment to quiet the solution, and then let it stand undisturbed for several
minutes, until the floc begins to settle. As soon as the clear solution above the
cloudiness can be seen, judge for yourself whether it is clear or color-less
enough. If not, add more sodium aluminate, mix violently, wait, and look again.
Repeat until you find the total number of drops needed per gallon to achieve the
effect you want.
Putting it All Together: The total number of drops of chlorine bleach and sodium
aluminate represent the ratio of the volumes of solution to mix together. For
example, if the chlorine dosage is 10 drops and the alum dosage is 20 drops,
mix the solutions in the ratio of 1:2. Then try out the mixed solution in the actual
installation: set the chemical feed pump at half-speed, turn on the water, and
take samples. Adjust the setting as needed to produce a free chlorine residual
of 3-5 ppm . But first, we need to discuss the rest of the system.
Designing a Chemical Feed System: The starting point is the desired or
expected peak flow rate. For most homes and small businesses, 5 gallons per
minute or 19 liters per minute is about right. The size of the contact tank or

52

mixing tank required is calculated by multiplying the flow rate times the desired
contact time and then dividing by a mixing efficiency factor: (5 gal/min.) (5 min.)
= 25-gal. tank; 28% = 89-gallon tank.
The efficiency of mixing in ordinary pressure tanks is very poor. Treatment
chemicals and treated water short-circuit from influent to effluent in contact
tanks quickly, providing much less actual contact time than expected. The best
that can be done is about 28% efficiency, and even that requires filling the
bottom third of an up-flow tank with pea-sized gravel. Thus, at a flow rate of 5
gal/min., a standard 82-gallon (310 L) galvanized steel tank with pea-gravel will
provide about 4.6 minutes of contact time; a 120-gal. (450 L) tank, 6.7 min.
F. Membrane Systems
This section is not about the microporous polymeric membranes used to make
pleated membrane micro-filters and also to collect and concentrate bacteria in
microbiology labs. Those are discussed in Sec. III-A, Mechanical Filtration.
Membrane systems are entirely different: they have no pores, and they do not
filter anything. Instead, they are formulated to be so extremely hydrophilic
(water-loving) that water dissolves into the membrane, and other materials
(chemicals, particles, microbes) are simply excluded by their chemical nature.
Exceptionsthings membrane systems dont remove wellinclude small, waterlike molecules such as alcohols, and other very small molecules, even if they
arent much like water, such as THMs and other VOCs. Once water molecules
are in the membrane, they find their way to the other side by diffusion and leave
everything else behind. The three major technologies using membrane systems
are reverse osmosis (RO), nanofiltration, and ultrafiltration. Several membrane
types are in use, and the most important ones are the cellulose acetate types,
the polyamide types, and the thin-film composite (TFC) types. The polyamide
membranes are readily damaged by disinfectant chlorine, but the cellulose
acetate membranes are not. TFCs are intermediate in chlorine sensitivity, and
both are often protected by activated carbon or granular brass prefilters.
Chlorine-resistant cellulose acetate membranes are used where the absence of
chlorine would allow degradation of the membrane by bacteria.
1. Ultrafiltration is the coarsest of the membrane technologies, generally being
limited to removing colloids the size of viruses and molecules the size of large
proteins and color bodies derived from rotting vegetation. Ultrafilters are
produced as long sheets of membrane that are rolled up into a cylinder. Alternating layers of netting provide space for water to penetrate into the interior of
the roll. This provides more surface area than the pleats often used for microfilters. They require good prefiltration by a micro-filter or fine-filter because they
clog quickly. Therefore, they do not find much use in water treatment. Some
food processing applications provide a waste stream that recirculates. This
helps delay clogging and enhances the life of the membrane.

53

2. Nanofiltration uses membranes that remove particles and molecules


hundreds of times smaller than ultrafilters, down to a molecular weight of about
1000. They are produced the same way as ultrafiltersrolled-up membrane
cylinders with netting spacersbut the design also includes a reject stream
that directs most of the water to waste before it can become so concentrated that
it causes clogging. This produces a tangential flow pattern that is less
conducive to clogging. That means there are three plumbing connections for
nanofilters: inlet, outlet product water, and outlet waste water. This is inefficient
in terms of wasting a lot of water, but highly efficient in terms of saving money on
premature replacement of expensive membrane cartridges.
Nanofiltration can be used alone to purify many water supplies without using any
other equipment or chemicals, because it can remove all particles of all sizes: all
colloids, all types of micro-organisms, nearly all pesticides (except for those that
are so small that they qualify as VOCs), and nearly all organic material that can
act as food for bacteria or serve as the precursors of THMs and other
disinfection byproducts. It can even remove 2/3 or more of ordinary hardness
and alkalinity, so that nanofiltration is often called membrane softening.
3. Reverse Osmosis is similar to nanofiltration, except that the membrane is a
thousand times more efficient, with the ability to remove 90% or more of individual atoms and ions. Ions with only one charge are removed less efficiently
than those with multiple charges. Thus, RO removes hardness and alkalinity
with 97-99% efficiency, while flouride (F), nitrate (NO3 ), cyanide (CN ), silver
(Ag+), etc. are removed only with about 95% efficiency. Mercury (Hg+) is the
most difficult, with efficiencies as low as 60% in some cases. As an exception,
some membranes are designed specifically for desalination of sea water, and
these remove Na+ and Cl with greater than 99% efficiency. RO unfortunately
wastes even more water than nanofiltration, often as much as 90%. Usually,
about 8 volumes are wasted to produce one volume of purified water.
Ultrafiltration and nanofiltration membranes are too coarse to develop much of
an osmotic pressure across them, (which see, Section II), but of course RO
membranes do. As a rule of thumb, one psi (6.9 kPa) of osmotic pressure is
produced by every 100 ppm of total dissolved solids. That means 2000 ppm of
TDS will produce about 20 psi (138 kPa) of osmotic back-pressure to overcome,
and that is the usual limit of water quality for RO systems that use only the
available water pressure to drive the process. Higher pressures, supplied by
pumps, allow RO systems to be both more efficient and more productive. Such
larger systems often have elaborate pretreatment to improve efficiency and
reduce maintenance. In addition to basic micro-filtration and chlorine removal,
this may include softening by ion exchange, injection of anti-scale chemicals,
and a use of a hot/cold mixing valve.
RO systems produce water so slowly that the product water is always collected
in tanks for dispensing later. Large systems with their own pumps usually have a
second pump to repressurize the product water, but the small systems that use

54

ordinary line pressure often use small bladder tanks that are pre-pressurized,
usually to 5 psi (34.5 kPa). These have only a single connection to serve as
both inlet and outlet. The RO system produces its trickle of water, which collects
in the tank as long as it can overcome the 5 psi pressure, or until the backpressure reaches 2/3 of the line pressure, and RO systems with a special
pressure-sensing valve shut themselves off. This saves water that otherwise
would run to waste while the tank is full. When water is drawn out for use, it is
propelled by the pressure in the tank, which is always at least 5 psi even when
empty. The reduction in pressure caused by using some of the water actuates
the pressure-sensing valve and allows systems to start up again and begin
producing more RO water.
Unfortunately, even RO membranes are unable to remove the small, lowmolecular-weight, volatile organic compounds (VOCs) such as THMs, solvents,
and some pesticides, and these must still be removed by adsorption to activated
carbon. Some systems are designed with the carbon adsorber immediately after
the membrane, subject to the slow trickle of product water, and others place the
adsorber between the bladder tank and the dispensing faucet. Either way, only
the product water is so treated. It would be a tremendous waste of carbon to
remove THMs, etc. from the influent water and then waste 90% of it as reject
water.
G. Ultra-Violet Irradiation
Electromagnetic radiation (light) covers an enormous range of wavelengths and
energies, from very weak but long radio waves many miles in length to extremely
powerful X-rays and gamma-rays with wavelengths only tiny fractions of a micron
in size. The size of the wave is inversely proportional to the energy it carries,
which means that light with the smaller wavelengths can do more damage. The
visible portion of the spectrum has wavelengths that range from about 0.40 um
or 400 nm for violet light to about 0.77 um or 770 nm for red light. Light with
wavelengths shorter than violet (but greater energy) is ultra-violet, and light with
wavelengths longer than red (but less energy) is infra-red. Thus, infra-red light is
simply heat, while visible light can give you a suntan, and UV light can cause
terrible burns and blindness.
It happens that UV light with a wavelength near 0.254 um or 254 nm is able to
produce resonance effects in the DNA of living cells which cause breaks and
kinks, leading to cell death. Therefore, UV irradiation is useful as a method of
disinfection for water. The energy is measured in watts (micro-watts) per cm2,
and the total dosage is the wattage multiplied by the time: uWsec/cm2. Thus, a
standard UV bulb that delivers 3800 uW/cm2 will produce a dosage of 38,000
uWsec/cm2 after ten seconds of exposure. That is the minimum dosage required
to meet the NSF International standard for Class A purification of raw waters
that may contain pathogenic bacteria and viruses. A lesser standard of 16,000
uWsec/cm2 can be NSF-Certified for controlling bacterial regrowth in the pipes of

55

systems in which the water has already been disinfected, and only non-pathogenic organisms are present. Standard UV systems are not powerful enough to
kill protozoan cysts and oocysts and the larger parasites, so these still require
physical removal. New pulsed-UV systems may prove to be effective against
cysts, but they are not yet approved by any health regulatory agency.

Advantages of UV treatment:
1. Treatment is essentially instantaneous. It is not difficult to provide a 10second contact time (an exposure vessel of less than a liter at a flow rate
of 1 gallon/min. (3.79 L/min.)) with modest resources.
2. Its all electrical, with no moving parts to break down, so the system can be
relatively simple in design and maintenance.
3. Because its all electrical, fail-safe measures are relatively easy to build in.
4. Water clarity is extremely important, but fine-filtration to remove turbidity
will also remove any protozoan cysts and larger parasites that would not
be killed.
Disadvantages of UV treatment:
1. There is no residual activity to protect the water against subsequent
contamination, so the overall system design may be complicated by the
need to place the UV last.
2. Routine maintenance to clean the optical surfaces is mandatory; the failsafe system should turn the system off if/when they get dirty.
3. Special meters for monitoring the wavelength and intensity of the UV light
are very costly, so complete systems cannot be inexpensive.
4. Treated water should be kept in the dark for 30 minutes, because ordinary
sunlight has the ability to activate repair enzymes in many bacteria, and as
many of 2/3 of those previously killed may be revived after only a
moments exposure.
5. There is a special design problem with small systems intended for
intermittent use: cold lamps require a warm-up to achieve peak efficiency,
but turning the system on and off repeatedly damages the lamp, and
leaving it on all the time heats the water and stimulates the growth of the
few organisms that survive.
6. UV treatment must not be used in recirculating systems because that can
produce a super-strain of UV-resistant bacteria that might be dangerous.
7. UV light is damaging to many plastics, so special shielding may be
important.

56

Das könnte Ihnen auch gefallen