Está en la página 1de 42

Reprinted from

2000. Reclamation of Drastically Disturbed Lands


American Society of Agronomy and American Society for Surface Mining and Reclamation.
Agronomy No. 41.
Chapter 6

ACID MINE DRAINAGE CONTROL AND TREATMENT

Jeffrey G. Skousen, Alan Sexstone and Paul F. Ziemkiewicz

6.1 INTRODUCTION

6.1.1 Acid Mine Drainage Formation

Acid mine drainage (AMD) forms when sulfide minerals are exposed to oxidizing
conditions in coal and metal mining, highway construction, and other large-scale excavations.
There are many types of sulfide minerals. Iron sulfides common in coal regions are
predominately pyrite and marcasite (FeS2), but other metals may be complexed with sulfides
forming chalcopyrite (CuFeS2), covellite (CuS), galena (PbS), and sphalerite (ZnS). Pyrite
commonly occurs with these other metal sulfides thereby causing AMD where Cu, Pb, and Zn
are mined.
Upon exposure to water and oxygen, sulfide minerals oxidize to form acidic, sulfate-rich
drainage. Metal composition and concentrations in AMD depend on the type and quantity of
sulfide minerals present. The drainage quality emanating from underground mines or backfills
of surface mines is dependent on the acid-producing (sulfide) and alkaline (carbonate) minerals
contained in the disturbed rock. In general, sulfide-rich and carbonate-poor materials produce
acidic drainage. In contrast, alkaline-rich materials, even with significant sulfide concentrations,
often produce alkaline conditions in water.
Acidity in AMD is comprised of mineral acidity (Fe, Al, Mn, and other metals depending
on the specific metal sulfide) and hydrogen ion acidity. Approximately 20,000 km of streams
and rivers in the United States are degraded by AMD. About 90% of the AMD reaching streams
originates in abandoned surface and deep mines. Since no company or individual claims
responsibility for reclaiming abandoned mine lands (AML), no treatment of the AMD occurs and
continual contamination of surface and groundwater resources results.
The oxidation of iron disulfides and subsequent conversion to acidity occur through
several reactions (Stumm and Morgan, 1970). The chemical equations are given in Chapter 5. If
any of the processes represented by the equations were slowed or altogether stopped, the
generation of AMD would also slow or cease. Removal of air and/or water from the system, two
of the three principal reactants, would stop pyrite from oxidizing. This occurs naturally where
pyrite-bearing rock occurs in water-saturated zones. Only small amounts of pyrite are oxidized
through weathering in undisturbed environments thereby generating only small amounts of
acidity. This small amount of acid is naturally diluted or neutralized by surrounding rocks.
However, when large volumes of pyritic material are exposed to oxidizing conditions, as in
1
mining, the pyrite reacts on a large scale and water moves the reaction products (Fe, sulfate, and
acidity) into underground and surface waters. Ferrous iron conversion to ferric iron is slow
under abiotic (or strictly chemical) conditions. But iron-oxidizing bacteria, Thiobacillus
ferrooxidans, greatly accelerate iron oxidation which, in turn, greatly speeds acid generation
(Waksman, 1922). Biotic catalysis can increase acid generation up to a million times (Leathen et
al., 1953).
Sulfur in coal and coal-bearing rock occurs as organic, sulfate, or pyritic forms (Chapter
4). Much of the sulfur in coal-bearing rock occurs as very small crystalline grains intimately
mixed with the organic constituents of coal. Sulfur may also be disseminated in nearby layers of
sandstone and shale (Temple and Koehler, 1954). The high sulfur content in coals and
associated strata deposited in marine settings suggests that bacterial reduction of sulfate in
seawater was the source of much of the pyrite found in coal.
Organic sulfur is complexed and inseparable from the coal, but is not chemically reactive
and has little to no effect on acid-producing potential. Sulfate sulfur is found only in minor
quantities in fresh coal and associated rocks. Sulfate is a reaction product of pyrite oxidation and
therefore is not responsible for acid production. Pyritic or sulfide sulfur is the predominant
species in coal and associated rocks and is the major acid-producer. Several types of pyritic
sulfur are known based on physical appearance (Chapter 5).
The equations show that 1000 Mg of rock containing 1% sulfur, all as pyrite, would yield
upon complete reaction 31.25 Mg of acidity. Therefore, 31.25 Mg of 100% CaCO3 would be
required to neutralize the acidity generated. When sulfur in the overburden or coal is exclusively
pyrite, the total sulfur content of the rock accurately quantifies the acid-producing potential.
When organic and/or sulfate sulfur are present in significant amounts, total sulfur measurements
overestimate the amount of acid that could be formed upon oxidation. Therefore, removal of
sulfates and organic sulfur may be necessary to increase the accuracy in predicting the acid-
producing potential of partially weathered materials (Chapter 4).
The rate of pyrite oxidation depends on numerous variables such as reactive surface area
of pyrite (Singer and Stumm, 1968), oxygen concentrations and solution pH (Smith and
Shumate, 1970), the presence of Thiobacillus bacteria (USEPA, 1971), and catalytic agents
(Caruccio et al., 1988). The possibility of identifying and quantifying the effects of these and
other controlling factors with all the various rock types in a field setting is unlikely. Further,
precise knowledge regarding oxidation rate is not needed for most real world situations when the
acid-producing potential has been accurately evaluated from total sulfur (Smith and Sobek,
1978).
The alkaline content of overburden determines whether there is enough inherent
neutralization potential (NP) to counteract the acid produced from pyrite oxidation. Of the many
types of alkaline compounds present in rocks, only carbonates and clays occur in sufficient
quantity to effectively neutalize AMD. A balance between the acid-producing potential and
neutralizing capacity of an overburden sample will indicate the ultimate acidity or alkalinity that
might be expected in the material upon complete weathering.
SMCRA requires coal operators and regulators to predict whether AMD may occur on a
potential mine site before disturbance. Accurate prediction of AMD before surface mining
requires a complete understanding of many components at a mine site. Three of the most
important factors are: 1) overburden geochemistry, 2) method and precision of overburden
handling and placement in the backfill during reclamation, and 3) the postmining hydrology of
2
the site.

6.1.2 Overburden Analyses

Premining analyses of soils, overburden, and the coal pavement are required by law to
ascertain the physical and chemical characteristics of the strata above and immediately below the
coal bed (Chapter 4). Overburden characterization provides important information about
overburden layers that are acid toxic, potentially acid-producing, neutral, or alkaline.
Overburden analysis for surface mining begins with Acid-Base Accounting (ABA). This
analytical technique provides a simple, relatively inexpensive, and consistent procedure to
evaluate overburdens. It balances potential acidity (based on total or pyritic sulfur content)
against total neutralizers (primarily carbonates) in an overburden sample. Samples containing
more acid potential than alkaline material are shown to be deficient in neutralizing materials
( maximum needed ), while the reverse situation is shown as "excess" neutralizing materials.
Rock layers that are marginal (having about equal proportions of each type of material)
can be subjected to leaching or weathering analyses (Skousen et al., 1987). Leaching/weathering
procedures were devised to predict the chemical responses of a rock to weathering. The
simulation tests are conducted over a predetermined time to estimate the rate of reaction of
minerals in the rock. Leaching methods consist of placing a rock sample (ground to a specified
particle size) into water-tight containers, flushing or flooding the containers with water, and
collecting and analyzing the effluent for pH, acidity, alkalinity, iron, sulfate, etc. Leaching
equipment may be varied and solvents may be recycled or heated during the leaching procedure.
Several computer software programs are also available to model hydrogeochemical systems to
predict AMD production. The models use overburden geochemical data, chemical and
biological oxidation rates, anticipated water infiltration rates, and environmental conditions.
These additional analyses sometimes supplement information given by ABA and can often help
evaluate how a particular rock type may react in a backfill. Identification of the chemical
production potential of overburden layers aids in developing overburden handling and placement
plans.

6.1.3 Overburden Handling and Placement

The prevailing approach to control AMD in eastern USA is to keep water away from
pyritic material. Once overburden materials have been identified and classified based on
overburden analyses, the overburden handling and placement plan for the particular site must be
carefully followed. Recommendations have focused on segregating and placing acid-producing
materials above the water table on top of a 1- to 2-m layer of nontoxic material on the pavement,
and then treating, compacting, and covering it with a clay cap or other type of sealant material to
reduce surface water infiltration into this material (Skousen et al., 1987). In areas where
limestone or other alkaline layers occur in the overburden, overburden materials may be blended
to neutralize the acidic materials. Limestone should be added to the topsoil to neutralize any
residual acidity and to raise the pH for vegetation establishment.

6.1.4 Postmining Hydrology

3
The hydrology of a backfill and its effect on AMD are very complex, but research on the
movement of water into and through a mine backfill has provided some information on
preventing and controlling AMD. Generally, the porosity and hydraulic conductivity of the
materials in a backfill are greater than those of the consolidated rock overburden that existed
before mining, and changes in flow patterns and rates should be expected after mining (Caruccio
and Geidel, 1989). Often, the fine-grained topsoil placed over the backfill conducts water more
slowly than the underlying coarse material, and thus the topsoil may determine the amount and
rate of water movement into the backfill (Guebert and Gardner, 1992). As water moves into
coarse materials in the backfill, it follows the path of least resistance. For example, water will
flow through more permeable acid sandstones and around calcareous shales. The water
continues downward until it encounters a barrier, the coal pavement, or other compacted or
slowly permeable layer. Water does not move uniformly through the backfill by a consistent
wetting front. The chemistry of the water emanating from the backfill will reflect only the rock
types encountered in the water flow path, not the geochemistry of the total overburden
(Ziemkiewicz and Skousen, 1992).
Diverting surface water above the site to decrease the amount of water entering the mined
area is highly recommended. Alternatively, pyritic material can be placed where it will be
rapidly and permanently inundated, thereby preventing oxidation of the acid-forming material.
Inundation is only suggested where a water table may be re-established to cover acid-producing
materials (such as below drainage deep mines), and has not been recommended for surface
mined lands or above drainage deep mines in the mountainous Appalachian USA region.
Upgradient of mined areas, water with low mineral acidity can be treated with limestone in the
water course to improve water quality.

6.2 AT-SOURCE CONTROL OF ACID MINE DRAINAGE

Acid mine drainage control technologies are measures that can be undertaken where
AMD formation has either already taken place or is anticipated to be a problem in the future.
At-source control methods treat the acid-producing rock directly and stop or retard the
production of acidity, whereas treatment methods add chemicals directly to acidified water
exiting the rock mass. Coal companies mining in acid-producing areas of the eastern USA must
often treat AMD, and they face the prospect of long-term to indefinite water treatment and its
attendant liabilities. It is obvious that cost-effective methods which prevent the formation of
AMD at its source would be preferable. Some control methods are most suitable for abandoned
mines and others are only practical on active operations. Other methods can be used in either
setting.
Some of the techniques described below are partially successful and have demonstrated
less than 100% control of acidity produced on-site, and are considered failures by some people.
While a technique that controls 80% of a site's acidity may not relieve a coal mine operator of
liability, the method may be suitable for an abandoned mine reclamation program or a watershed
restoration project. Removing a significant portion of the acid or metal load in a watershed by a
partially-effective control strategy may improve the health of a stream to a point of re-
introducing some fish species or re-establishing some designated uses of the stream.
Alternatively, the method may be combined with another partial control scheme to achieve
effluent limits. Since partial control methods are often the least costly, their use in combination
4
with other techniques is often financially attractive. For example, a partial control method such
as a dry barrier can be combined with alkaline amendment to reduce the amount of alkalinity
required to offset acidity.

6.2.1 Alkaline Amendment to Active Mines

Recent studies have indicated that certain types of alkaline amendments can successfully
control AMD from pyritic spoil and refuse (Brady et al., 1990; Perry and Brady, 1995; Rich and
Hutchison, 1990; Rose et al., 1995). Nearly all alkaline amendment schemes rely on ABA to
identify the required alkalinity for neutralization of pyritic materials. Alkaline amendment
methods are a modification of the concept of selective handling. Selective handling seeks to
blend acid-producing and acid-neutralizing rock units in the mining process to develop a neutral
rock mass. Selective handling may also create a postmining hydrologic regime that minimizes
the contact between acid-forming rock and groundwater, or it may isolate acid-producing rocks
from the rest of the backfill by use of barriers. In the eastern USA coal fields, the pit floor is
often rich in pyrite, so isolating it from groundwater may be necessary. Isolation methods can
include building highwall drains which move incoming groundwater away from the pit floor or
placing impermeable barriers on the pit floor. For example, acid-forming material can be
compacted or capped within the spoil (Meek, 1994). If insufficient alkalinity is available in the
spoil, then external sources of alkalinity may be imported (Skousen and Larew, 1994; Wiram
and Naumann, 1995).
Limestone is often the least expensive and most readily available source of alkalinity. It
has a NP of between 75 and 100%, and is safe and easy to handle. On the other hand, it has no
cementing properties and cannot be used as a barrier.
Fluidized Bed Combustion (FBC) ash is produced at power generating plants that burn
high sulfur coal or refuse in a FBC system. Sulfur dioxide emissions are controlled by injecting
limestone into the combustion bed. At combustion temperatures, the limestone calcines leaving
calcium oxide. About one-half of the CaO reacts with sulfur dioxide to form gypsum and the
rest remains unreacted. Therefore, FBC ashes generally have NPs of between 20 and 40%, and
they tend to harden into a cement after wetting (Skousen et al., 1997). Other power-generation
ashes, like flue gas desulfurization products and scrubber sludges, may also have significant NP,
which make them suitable alkaline amendment materials (Stehouwer et al., 1995).
Kiln dust, produced by lime and cement kilns, contains similar levels of CaO (15 to 30%)
as FBC ash, but also contains 50 to 70% unreacted limestone. Kiln dust absorbs moisture and
also hardens upon wetting (Rich and Hutchison, 1994). It is widely used as a stabilization and
barrier material.
Steel-making slags are locally available in large quantities at low cost. When fresh, steel
slags have NPs from 45 to 90%. Studies indicate that columns of steel slag maintain constant
hydraulic conductivity over time and produce highly (>1,000 mg/L) alkaline leachate. Steel slag
can be used as an alkaline amendment as well as a medium for alkaline recharge trenches. Slags
are produced by a number of processes, so care is needed to ensure that candidate slags are not
prone to leaching metal ions like Cr, Mn, and Ni.
Other alkaline materials may have higher NPs than limestone, but the source of the
material should be checked and a complete analysis should be done to evaluate NP and metal
content before use. Quicklime, kiln dust and hydrated lime all have higher activities than
5
limestone, though it is not clear that the kinetics of pyrite oxidation favor readily soluble sources
of alkalinity.
Phosphate rock has been used in some studies to control AMD. It may react with Fe
released during pyrite oxidation to form insoluble coatings (Evangelou, 1995), but phosphate
usually costs much more than other calcium-based amendments and is needed in about the same
amounts (Ziemkiewicz and Meek, 1994).
Accurately predicting the amount of a particular alkaline amendment to add in a certain
situation has significant cost and long-term liability implications. diPretoro and Rauch (1988)
found sites with >3% NP as calcium carbonate equivalent in overburden produced alkaline
drainages, while acidic drainage resulted at <1% NP. Brady and Hornberger (1989) suggest
threshold values of NP > 3% and S < 0.5% as guidelines for delineating alkaline-producing
strata. Brady et al. (1994) showed that 3% net NP in an overburden caused alkaline drainage
while less than 1% net NP produced acidic drainage from 38 mines in PA. They concluded that
mining practices (such as selective handling, and concurrent reclamation) enhanced the effect of
alkaline addition on reducing acidity. Further refinements (Perry and Brady, 1995) gave a value
of 21 Mg/1000 Mg net NP (or 2.1%) to produce alkaline drainage on 40 sites in Pennsylvania.
Alkaline addition of lime kiln dust at rates to neutralize the MPA was successful in producing
alkaline drainage from a Pennsylvania site after mining (Rose et al., 1995).
Brady and Hornberger (1990) conclude that NP from ABA shows the strongest
relationship with actual postmining water quality. This relationship is only qualitative (e.g. acid
vs. non-acid), and NP must significantly exceed MPA in order to produce alkaline water. If NP
and MPA are similar, AMD will most likely result.
Accurate, representative overburden analyses before mining are crucial in developing
mining plans and alkaline addition programs in acid-producing areas. Therefore, it is best to
generate cost-effective control strategies when acid problems are identified during premining
planning (Chapter 4). Errors in predicting postmining water quality from premining overburden
analyses include unrepresentative sampling of overburdens and inaccurate analyses (Rymer et
al., 1991; Wiram and Naumann, 1995), and non-homogeneous placement of spoils. For
example, Schueck (1990) reported AMD generation from a surface mine in Pennsylvania
resulted largely from buried refuse and pit cleanings within an otherwise neutral to alkaline spoil
matrix as identified by ABA.
Some spoils are composed of a mixture of acid-forming and alkaline rocks, while other
materials like refuse are dominated by acid-producing rocks with no NP. In spite of significant
alkalinity in overburdens, AMD originates from localized sites within the backfill. While
finding the path of least resistance to the downstream side of the backfill, infiltrating water is
influenced only by acid and alkaline rocks directly in its path. If water flows primarily through
permeable acid sandstones, AMD can result, and the water flows freely to the nearest stream
while the alkalinity in the pile remains unreacted. Unless contacted directly by acid water, most
of the spoil limestone will remain in solid form. Thus, the presence of alkalinity in the backfill
does not ensure that it will neutralize acidity. For efficient neutralization, the acid-forming and
alkaline material must be thoroughly mixed. Where insufficient alkalinity is present, it is
necessary to add alkaline material to the rock mass. If one relies on random spoil dumping, an
overwhelming supply of alkaline material would be needed. This probably accounts for the
reported field observations that two or more units of NP is required for each unit of MPA. If
there is a consensus on the subject, it would appear that material mixtures with an NP/MPA ratio
6
above 2 and an net NP above 30 Mg/1000 Mg (3% net NP) generate alkaline water. Even at
high NP/MPA ratios, acidity may be produced if the acid-producing materials are incorrectly
placed so that they intercept groundwater, or the alkaline material is added only to the surface
(Lusari and Erickson, 1985). Thorough mixing during materials handling can reduce the
required NP/MPA ratio considerably.
Addition of alkaline material to refuse is a relatively simple process. Since the refuse
leaves the preparation plant on a belt conveyor or in a slurry line, the alkalinity can be metered
directly into the refuse stream. Rich and Hutchison (1990 and 1994) reported a successful
operation where 2% lime kiln dust is added to refuse at a preparation plant in West Virginia.
Not only did the kiln dust prevent acid formation, it improved the strength of the refuse pile by
absorbing moisture from the filter cake, allowing easy access for large haulage trucks. Currently
at least eight preparation plants in the eastern USA are using this technique.
Adding alkalinity to spoil requires more care during mining and reclamation. The high
volume of spoil necessitates applying alkalinity only to those rock masses that need it (like coal
roof rock, partings and pavement). Surface mining operations can remove this rock with front
end loaders after the prime excavator exposes the coal and place the acid-producing material in
cells in the backfill for treatment with an alkaline amendment (Skousen and Larew, 1994). Since
acid-producing rock types are often a small proportion of the total spoil, such a procedure results
in efficient use of the alkaline material. Additional compaction is desirable during placement in
cells. Often the final pit floor contains considerable pyrite, much of it within 30 cm of the coal.
Two approaches can be taken: 1) remove the pyritic material and place it in cells for alkaline
amendment, or 2) seal the pit floor with a self-cementing material such as FBC ash.

6.2.2 Alkaline Recharge Trenches

Alkaline recharge trenches (Caruccio et al., 1984) are surface ditches filled with alkaline
materials. These trenches can minimize or eliminate acid seeps through an alkaline-loading
process by placing alkaline material into contact with surface infiltrating water. Alkaline
recharge trenches were constructed on top of an 8-ha coal refuse disposal site, which produced
AMD seepage (Nawrot et al., 1994). After installing the alkaline recharge pools, acidity
reductions of 25 to 90% were realized with concomitant 70 to 95% reductions in Fe and sulfate
in seepage water. The following conclusions and recommendations were made:
--Use highly soluble alkaline materials (e.g. CaO, Ca(OH)2 waste products),
--Maximize water volumes through the trench by directing surface flow into the pool,
--Use multiple alkaline recharge pools to increase chances of influencing groundwater flows,
--Construct infiltration paths into the backfill to improve alkaline diffusion and flushing,
--Allow sufficient time (possibly 3-5 annual cycles) for the effect to become apparent.

6.2.3 Bactericides

Anionic surfactants are used to control bacteria that catalyze the conversion of Fe+2 to
+3
Fe , which can thereby control pyrite oxidation. They are used in situations where immediate
control of AMD formation is important and work best on fresh, unoxidized sulfides.
Bactericides are often liquid amendments, which can be applied to refuse conveyor belts or
sprayed by trucks on cells of acid-producing materials in the backfill. Bactericides have also
7
been used at metal mines (Parisi et al., 1994). One example of surfactant application was done
in 1988 on a 4.5-ha refuse site in Pennsylvania. Surfactant was applied via a hydroseeder at
rates of 225 kg/ha initially, then successive amounts were made as fresh refuse was deposited.
Effluent from the pile showed a 79% reduction in acidity and a 82% decrease in Fe. Cost
savings at the AMD treatment plant were $300,000 per year (Rastogi, 1994). Surfactants, by
themselves, are not seen as a permanent solution to AMD. Eventually the compounds either
leach out of the rock mass or are decomposed. However, slow-release formulations are
commercially available and have been successfully used at regraded sites (Splittorf and Rastogi,
1995). Bactericides appear to work best when used in combination with other control methods,
and can be useful in preventing acid conditions in pyritic rock piles which remain open for
several years until the site is reclaimed.

6.2.4 Dry Barriers

Barriers are constructed from materials that retard the movement of water and oxygen
into areas containing acid-producing rock. Barriers can achieve substantial reductions in water
flow through piles, but generally do not control AMD completely. Grouts can be used to
separate acid-producing rock and groundwater. Injection of grout barriers may significantly
reduce the volume of groundwater moving through spoil. Gabr et al. (1994) characterized the
groundwater flow of an acid-producing reclaimed site where a 1.5-m thick wall was installed by
pumping a mixture of class F fly ash and portland cement grout into vertical boreholes near the
highwall. After two years, the grout wall reduced groundwater inflow from the highwall to the
spoil by 80%, resulting in one of two seeps completely drying up and substantially reducing the
flow of the other seep.
Plastic liners are rarely used in mining because covering the large volumes of waste with
a liner is usually too expensive. However, this method may be appropriate in settings where
isolation of small pods of acid-producing material is possible. At the Upshur Mining Complex
in West Virginia, Meek (1994) reported covering a 20-ha spoil pile with a 39-mil PVC liner.
This treatment reduced acid loads by 70%. At the Heath Steele Metal Mine in New Brunswick,
a soil cover was designed to exclude oxygen and water from the pile, and its performance was
evaluated for five years (Bell et al., 1994). The capillary barrier consisted of a 10-cm gravel
layer for erosion control, 30-cm gravel/sand layer as a evaporation barrier, 60-cm compacted till
(conductivity of 10-6 cm/sec), 30-cm sand, and pyritic waste rock. The barrier excluded 98% of
precipitation, and oxygen concentrations in the waste rock dropped from 20% initially to around
1%. In a similar study at the Waite Amulet Mine in Quebec, a barrier consisting of a 60-cm
thick layer of compacted clay between two 30-cm layers of sand was placed over pyritic tailings
(Yanful et al., 1994). After three years, the barrier reduced oxygen flux by 99% resulting in a
95% reduction in the rate of acid generation. The barrier excluded 96% of precipitation.
Evaluations of capillary barriers by these and other studies (Nicholson et al., 1989; Rasmuson
and Eriksson, 1986) indicate that this type of dry barrier can be an effective AMD control agent.

6.2.5 Wet Covers

Disposal of sulfide tailings under a water cover, such as a lake or fjord, is another way to
prevent acid generation by excluding oxygen to sulfides. Wet covers also include flooding of
8
above ground tailings ponds. Deposition of sulfide tailings under water has been used at various
mines in Canada. Fraser and Robertson (1994) studied four freshwater lakes used for
subaqueous tailings disposal and found the following:
--Reactivity of tailings was low since low dissolved metals were found in the pore water,
--Low metal dissolution from sediments indicated little impact of metals within the lakes,
--Biological communities existed in some of the lakes,
--Metal levels in fish were variable and depended on the history of the lake. Metal levels in fish
appeared to be elevated in some parts of the lake while others were typical of background
levels.
--Sulfide-rich Zn, Cu and Pb tailings were placed by slurry line into shallow (<8 m) Anderson
Lake in north-central Manitoba since 1979. Pedersen et al. (1994) studied the site and
concluded:
--Mine tailings were widely distributed on the lake floor,
--Sediments deposited in the lake had much organic matter,
--Anaerobic conditions developed at shallow depths,
--No Cd, Pb or Cu were being released from the submerged tailings,
--Very low amounts of Zn were dissolved into lake water from the tailings.
At Elliot Lake, Ontario Rio Algom Limited established a 65-ha flooded field demonstration site
in which water depths of 0.5 to 1.0 m were developed in a series of terraces and dikes (Dave and
Vivyurka, 1994). The first cell was completed and ready for flooding in October 1992. Over
50% of the sulfidic uranium tailing s surface was treated with limestone and the top 15 cm was
mixed. While flooding this cell, lime slurry was added to the inflow water for additional
neutralization. At the end of flooding in December 1992, the surface water pH was 7 with
acidity averaging 7 mg/L. This general water quality has been maintained for two years. For
comparison, inflow from a nearby uncontaminated lake had an average pH of 6.4 and acidity of 5
mg/L and alkalinity of 4 mg/L. The tailings pond was in a valley so only the engineered dam at
the downstream was exposed to air.

6.2.6 Alkaline Amendment to Abandoned Mines

Abandoned mines in the eastern USA generate more than 90% of the AMD in streams
and rivers, most of which comes from underground mines. Abandoned underground mines are
problematic because they are often partially caved and flooded, access is restricted, and reliable
mine maps are often unavailable. Abandoned surface mines comprise huge volumes of spoil of
unknown composition and hydrology. Rehandling and mixing alkalinity into the backfill is
generally prohibitively expensive. These problems on abandoned mines were considered so
intractable that they have only recently been addressed.
Filling underground mine voids with non-permeable materials is one of the best methods
to prevent AMD from abandoned underground mines. Since underground mine voids are
extensive (a 60-ha mine with a coal bed height of 1.5 m and a recovery rate of 65% would
contain about 600,000 m3 of voids), the fill material and the placement method must have very
low unit costs. Mixtures of class F fly ash and 3-5% portland cement are used to control
subsidence in residential areas. The slurries are injected through vertical boreholes at between 8
and 16 m centers. Research and demonstration projects are exploring both pneumatic (the use of
air pressure) and slurry injection methods for placing FBC ash in abandoned underground mines
9
(Burnett et al., 1995). Preliminary results indicate that pneumatic methods can extend the
borehole spacing to about 30 m at costs substantially less than those of slurry methods. On
reclaimed surface mines still producing AMD, researchers in Pennsylvania saw small
improvements in water quality after injecting coal combustion residues into buried pods of
pyritic materials (Kim and Ackman, 1994; Schueck et al., 1994).

6.2.7 Remining and Reclamation

Remining is returning to abandoned surface or underground mines for further coal


removal. Where AMD occurs, remining reduces acid loads by: 1) decreasing infiltration rates,
2) covering acid-producing materials, and 3) removing the remaining coal which is the source of
most of the pyrite. Hawkins (1994) studied 57 discharges from 24 remined sites in
Pennsylvania, and found contaminant loadings (e.g. flows and metal concentrations) were either
reduced or unchanged after remining and reclamation. Short-term loads were sometimes
increased during the first six months after remining and reclamation. Reduction in loads resulted
from decreased flow rather than large changes in concentrations. An evaluation of ten remining
sites in Pennsylvania and West Virginia showed eight of the sites to produce a net profit from
coal sales. All the sites were reclaimed to current standards, thereby eliminating highwalls,
covering refuse, and revegetating the entire areas. All sites also had improved water quality and
some completely eliminated AMD coming from the site (Skousen et al., 1997). Faulkner and
Skousen (1995) found significant reductions in acid loads after land reclamation, and the acid
load reductions were due both to reductions in water flow from the site and reductions in acid
concentration in the water.
Remining has been combined successfully with alkaline addition and special handling to
change water quality from acid to alkaline at specific sites. Skousen and Larew (1994) described
the surface remining of an underground mine in Preston County, West Virginia. Alkaline
overburden from an adjacent mine was imported at a rate of 15,000 Mg/ha to the remining job at
a cost of $0.55/Mg of coal removed. An average of about 18,000 Mg/ha coal was removed,
making the cost of hauling the alkaline material to the site around $9,900/ha. Water quality from
the deep mine prior to remining averaged pH 3.7 and 75 mg/L acidity, while after remining and
reclamation, pH was above 7.0 with no acidity. Based on average premining water flows and
analyses of the deep mine discharge on the site, the chemical cost for treating AMD on this site
would have been around $200,000 over a 20-year period. The alkaline addition had a one-time
cost of $45,000.

6.3 CHEMICAL TREATMENT OF ACID MINE DRAINAGE

Since the passage of SMCRA in 1977, coal mine operators have been required to meet
environmental land reclamation performance standards established by federal and state
regulatory programs. Operators must also meet water quality standards established in the Clean
Water Act of 1972 (CWA), which regulates discharges into waters of the USA. In addition to a
surface mining permit required by SMCRA, each mining operation must be issued a National
Pollution Discharge Elimination System (NPDES) permit under CWA. Allowable pollutant
discharge levels are usually determined by the USEPA s technology-based standards, or the
discharge levels may be based on the more stringent water quality-based standards. Water
10
quality standards may apply where discharges are being released into streams with designated
uses.
If AMD problems develop during mining or after reclamation, a plan to treat the
discharge must be developed. Treatment of AMD includes neutralization of acidity and
precipitation of metal ions to meet the relevant effluent limits. In most cases, a variety of
alternative treatment methods can be employed to meet the limits specified.
NPDES permits on surface mines usually require monitoring of pH, total suspended
solids (TSS), and Fe and Mn concentrations. Other parameters may be requested by the
regulatory authority in a particular mining situation. In order for an operator to make a selection
of an AMD treatment system, one must determine (in addition to the above parameters) the flow
rate, sulfate concentration, and Fe+2 concentration in the AMD. The receiving stream's
designated use and seasonal fluctuations in flow rate are also important. After evaluating these
variables over a period of time, the operator can consider the economics of different chemicals
and alternative AMD treatment systems. Most AMD chemical treatment systems consist of an
inflow pipe or ditch, a storage tank or bin holding the treatment chemical, a means of controlling
its application rate, a settling pond to capture precipitated metal oxyhydroxides, and a discharge
point. The latter is the point at which NPDES compliance is monitored.

6.3.1 Chemicals Available to Treat AMD

Six chemicals are used to treat AMD (Table 6.1). Each chemical has characteristics that
make it more or less appropriate for a specific condition. The best choice among alternatives
depends on both technical and economic factors. The technical factors include acidity levels,
flow, and the types and concentrations of metals in the water. The economic factors include
prices of reagents, labor, machinery and equipment, the number of years that treatment will be
needed, and the interest rate.
Enough alkalinity must be added to raise water pH so insoluble metal hydroxides will
form and settle out of the water. The pH required to precipitate most metals from water ranges
from pH 6 to 9 (except Fe+3, which precipitates at pH >3.5). The types and amounts of metals in
the water therefore heavily influence the selection of an AMD treatment system. Ferrous iron
converts to a solid bluish-green Fe(OH)2 at pH >8.5. In the presence of oxygen, Fe+2 oxidizes to
Fe+3, and Fe(OH)3 forms a yellowish-orange solid (commonly called yellow boy) which
precipitates at pH >3.5. In oxygen-poor AMD where Fe is primarily in the Fe+2 form, enough
alkalinity must be added to raise the solution pH to 8.5 before Fe(OH)2 precipitates. A more
efficient way of treating high Fe+2 AMD is to first aerate the water (also outgassing CO2),
causing the Fe to convert from Fe+2 to Fe+3, and then adding a neutralizing chemical to raise the
pH to 6 or 7 to form Fe(OH)3. Aeration after chemical addition is also beneficial because it
greatly reduces the amount of neutralizing reagent necessary to precipitate Fe from AMD.
Aluminum hydroxide precipitates at pH > 5.0 but also enters solution again at a pH of 9.0.
Manganese precipitation is variable due to its many oxidation states, but will generally
precipitate at a pH of 9.0 to 9.5. As this discussion demonstrates, the appropriate treatment
chemical can depend on both the oxidation state and concentrations of metals in the AMD
(USEPA, 1983). Interactions among metals also influence the rate and degree to which metals
precipitate. For example, Fe precipitation will largely remove Mn from the water at pH 8 due to
co-precipitation, but only if the Fe concentration in the water is much greater than the Mn
11
content (about 4 times more or greater). If the Fe:Mn ratio is less than 4, Mn is not removed by
co-precipitation and a solution pH of >9 is necessary to remove it from solution.

6.3.1.1 Limestone

Limestone has been used for decades to raise pH and precipitate metals in AMD. It has
the lowest material cost and is the safest and easiest to handle of the AMD chemicals.
Unfortunately, its successful application has been limited due to its low solubility and tendency
to develop an external coating, or armor, of Fe(OH)3 when added to AMD. In cases where pH
is low and mineral acidity is also relatively low (low metal concentrations), finely-ground
limestone may be dumped in streams directly or the limestone may be ground by water-powered
rotating drums and metered into the stream. These applications have recently been tried in West
Virginia in AMD-impacted streams with great success. Limestone has also been used to treat
AMD in anaerobic (anoxic limestone drains) and aerobic environments (open limestone
channels). These latter two techniques are discussed later in this chapter under passive treatment
systems.

Table 6.1. Chemical compounds used in AMD treatment.


____________________________________________________________________________________________
Common Name Chemical Name Formula Conversion Neutralization 2000 Cost3
1 2
Factor Efficiency $ per Mg or L
Bulk <Bulk
____________________________________________________________________________________________
Limestone Calcium carbonate CaCO3 1.00 30% $ 11 $ 16
Hydrated Lime Calcium hydroxide Ca(OH)2 0.74 90% $ 66 $110
Pebble Quicklime Calcium oxide CaO 0.56 90% $ 88 $264
Soda Ash Sodium carbonate Na2CO3 1.06 60% $220 $350
Caustic Soda (solid) Sodium hydroxide NaOH 0.80 100% $750 $970
20% Liquid Caustic Sodium hydroxide NaOH 784 100% $0.06 $0.16
50% Liquid Caustic Sodium hydroxide NaOH 256 100% $0.29 $0.33
Ammonia Anhydrous ammonia NH3 0.34 100% $330 $750
____________________________________________________________________________________________
1
The conversion factor may be multiplied by the estimated Mg acid/yr to get Mg of chemical needed for
neutralization per year. For liquid caustic, the conversion factor gives L needed for neutralization.
2
Neutralization Efficiency estimates the relative effectiveness of the chemical in neutralizing AMD acidity. For
example, if 100 Mg of acid/yr was the amount of acid to be neutralized, then it can be estimated that 82
Mg of hydrated lime would be needed to neutralize the acidity in the water (100(0.74)/0.90).
3
Price of chemical depends on the quantity being delivered. Bulk means delivery of chemical in a large truck,
whereas <Bulk means purchased in small quantities. Liquid caustic prices are for L. Others in Mg.

6.3.1.2 Hydrated Lime

Hydrated lime is a commonly-used chemical for treating AMD. It is sold as a powder


that tends to be hydrophobic, and extensive mechanical mixing is required to disperse it in water.
Hydrated lime is particularly useful and cost effective in large flow, high acidity situations

12
where a lime treatment plant with a mixer/aerator is constructed to help dispense and mix the
chemical with the water (Skousen and Ziemkiewicz, 1996). Hydrated lime has limited
effectiveness if a very high pH is required to remove ions such as Mn. Operators of lime
treatment systems often increase lime application as Mn levels increase in the water. However,
due to the kinetics of lime dissolution, increasing the lime rate increases the volume of unreacted
lime that enters the metal floc settling pond.
Hydrated lime can be purchased in 20-kg bags or in bulk. Bulk lime is preferred by
operators due to cost and handling advantages. It can be delivered by barge, truck, or train to
many sites and handled pneumatically. Proper storage of Ca(OH)2 is important in order to
maintain its flow characteristics and thus ensure efficient use. The appropriate silo volume
depends on the daily lime requirement, but should contain enough to cover periodic unexpected
delays in delivery. The length of time that the system will be in operation is important because
the large initial capital expenditure can be amortized over time.

6.3.1.3 Pebble Quicklime

Pebble quicklime (CaO) has been recently used in conjunction with the Aquafix Water
Treatment System utilizing a water wheel concept (Jenkins and Skousen, 1993). The amount of
chemical applied is dictated by the movement of the water wheel, which causes a screw feeder to
dispense the chemical. The hopper and feeder can be installed in less than an hour. This system
was initially used for small and/or periodic flows of high acidity because CaO is very reactive.
Recently, however, water wheels have been attached to large bins or silos for high flow/high
acidity situations. Preliminary tests show an average of 75% cost savings over NaOH systems
and about 20 to 40% savings over NH3 systems.

6.3.1.4 Soda Ash

Soda ash (Na2CO3) is generally used to treat AMD in remote areas with low flow and
low amounts of acidity and metals. Selection of Na2CO3 for treating AMD is usually based on
convenience rather than chemical cost. Soda ash comes as solid briquettes, and is gravity fed
into water by the use of bins or barrels. The number of briquettes to be used each day is
determined by the rate of flow and quality of the water to be treated. One problem with the bin
system is that the briquettes absorb moisture, causing them to expand and stick to the corners of
the bin. This hinders the briquettes from dropping into the AMD stream. For short-term
treatment at isolated sites, some operators use a much simpler system employing a wooden box
or barrel with holes that allows water inflow and outflow. The operator simply fills the barrel
with briquettes on a regular basis and places the barrel in the flowing water. This system offers
less control of the amount of chemical used.

6.3.1.5 Caustic Soda

Caustic soda (NaOH) is often used in remote locations (e.g., where electricity is
unavailable), and in low flow, high acidity situations. It is commonly the chemical of choice if
Mn concentrations in the AMD are high. The system can be gravity fed by dripping liquid
NaOH directly into the AMD. Caustic is very soluble in water, disperses rapidly, and raises the
13
pH of the water quickly. Caustic should be applied at the surface of ponded water because the
chemical is more dense than water. The major drawbacks of using liquid NaOH for AMD
treatment are high cost and dangers in handling.
Liquid NaOH can freeze during winter months, but there are several options available to
deal with the freezing problem. These include burying the NaOH tank, installing a tank heater,
switching from a 50 percent to a 20 percent solution, adding a small amount of antifreeze to the
solution, and utilizing solid NaOH. Switching from a 50 percent to a 20 percent NaOH solution
lowers the freezing point from 0o C to about -37o C. Antifreeze can be added at 1 L per 25 L of
50% NaOH to lower the freezing point from zero to -40o C. Solid NaOH, which may be
delivered in a 30-kg drum or in bags, is reported by some operators to be cheaper and easier to
handle than liquid NaOH.
Tanks housing NaOH can range in volume from 2,000 to 30,000 L. Large tanks are
usually placed on a cement platform. The discharge line is fixed at the bottom of the tank and
transports the NaOH solution to the seep, ditch, or pond. The rate of flow is controlled by a gate
valve placed at the end of the discharge line.

6.3.1.6 Ammonia

Ammonia or sometimes called anhydrous (NH3 or NH4OH) is an extremely hazardous


chemical that must be handled carefully (Hilton, 1990). A gas at ambient temperatures, NH3 is
compressed and stored as a liquid but returns to the gaseous state when released. Ammonia is
extremely soluble in water and reacts rapidly. It behaves as a strong base and can easily raise the
pH of receiving water to 9.2. At pH 9.2, it buffers the solution to further pH increases, and
therefore very high amounts of NH3 must be added to elevate the pH beyond 9.2. Injection of
NH3 into AMD is one of the quickest ways to raise water pH. It should be injected near the
bottom of the pond or water inlet because NH3 is lighter than water. The most promising aspect
of using NH3 for AMD treatment is its cost, especially compared to NaOH. A cost reduction
figure of 50% to 70% is usually realized when NH3 is substituted for NaOH (Skousen et al.,
1990).
Major disadvantages of using NH3 include: 1) hazards associated with handling the
chemical, 2) uncertainty concerning nitrification, denitrification and acidification downstream,
and 3) the consequences of excessive application rates (Faulkner, 1990). Specialized training
and experience are important for the safe use of NH3. Companies using NH3 are required to
conduct additional analyses such as temperature, total NH3-N, and total acidity for discharge
water and receiving streams. While NH3 can be effective for Mn removal in many cases, this
requires careful monitoring and attention, and overapplication of NH3 often occurs. Therefore,
in situations where Mn is the ion of primary concern (low Fe, high Mn water), a different
chemical may be more appropriate (Faulkner, 1990).
6.3.2 Costs of Treating AMD

Costs have been developed for five AMD treatment chemicals under four sets of flow
and acid concentration conditions (Table 6.2). These conditions are: 1) 189 L/min (50 gpm) and
100 mg/L; 2) 3780 L/min (1000 gpm) and 100 mg/L; 3) 945 L/min (250 gpm) and 500 mg/L; 4)
3780 L/min (1000 gpm) and 2500 mg/L. These conditions represent a sufficiently wide range
for valid comparison of the treatment systems.
14
The costs for each technology were divided into installation costs and variable costs.
Installation costs include piping, system foundation, site preparation, equipment, hardware, and
labor. Labor costs were based on man hours at a wage scale of $27 an hour. Variable costs
include reagent cost, annual labor, and maintenance. The amount of reagent was computed using
acid neutralization formulas presented in Skousen and Ziemkiewicz (1996). Annual labor was
estimated man-hours to run the system for one year multiplied by $27 an hour. Other variable
costs include repair costs and electricity (Phipps et al., 1991).
The prices for the reagents, equipment, and labor were based on actual costs to operators
in West Virginia in May 1996. The price for each chemical is given in Table 6.1, and varies due
to whether the material was delivered in bulk or purchased in bags. The net present value (NPV)
is the value of the total treatment system, plus annual operating and chemical expenses over the
specified duration of treatment in 1996 USA dollars. A rate of 6% per year was used to devalue
the dollar during future years of the treatment period. The annual cost was obtained by
converting the total system cost (NPV) to an equivalent annual cost so that each system could be
compared. The parameters used in the analysis were entered in a spreadsheet.
Use of Na2CO3 has the highest labor requirements (10 hours per week) because the
dispensers must be filled by hand and inspected frequently (Table 6.2). Caustic has the highest
reagent cost per mole of acid-neutralizing capacity and Na2CO3 has the second highest.
Hydrated lime treatment systems have the highest installation costs of the five technologies
because of the need to construct a lime treatment plant and install a pond aerator. However, the
cost of Ca(OH)2 is very low. The combination of high installation costs and low reagent cost
make Ca(OH)2 systems particularly appropriate for long-term treatment of high flow/high acid
condition situations.
For a five-year treatment period, NH3 had the lowest annual cost for the low flow/low
acid situation (Table 6.2). Pebble quicklime was only about $160 per year more expensive than
the NH3 system but had lower reagent and higher installation costs. Caustic was third because of
its high labor and reagent costs, and Na2CO3 was fourth due to high labor costs. Hydrated lime
was the most expensive because of its high installation costs. With the intermediate flow and
acid cases, NH3 and CaO systems were by far the most cost effective, with Ca(OH)2 and Na2CO3
next. Caustic was the most expensive alternative at this intermediate flow and acidity condition.
In the highest flow/acidity category, the Ca(OH)2 and CaO systems are clearly the least costly
treatment systems, with annual costs of about $250,000 less than NH3, the next best alternative.

6.3.3 Other Aspects of AMD Treatment Technologies

While the primary AMD chemicals and applications have been discussed, particular
circumstances may require a different chemical, a combination of chemicals, and particular
management patterns to implement the most cost effective method, or to meet more stringent
effluent limits.
Other chemicals used sparingly in AMD treatment include flocculants or coagulants, which
increase particle settling efficiency (Table 6.3). These materials are usually limited to cases
where unique metal compositions require a specialized treatment system, or where aeration and
residence time in settling ponds are insufficient for complete metal precipitation. Coagulants
reduce the net electrical repulsive forces at particle surfaces, thereby promoting consolidation of
small particles into larger particles. Flocculation aggregates or combines particles by bridging
15
the space between particles with chemicals. Bridging occurs when segments of a polymer chain
absorb suspended particles creating larger particles (Skousen et al., 1993).
The most common coagulants and flocculants used in water treatment are alum and
Fe2(SO4)3. These materials are also called polyelectrolytes and produce highly-charged ions
when dissolved in water. Anionic polymers dissolve to form negatively-charged ions that are
used to remove positively-charged solids. The reverse occurs with cationic flocculants.
Polyampholytes are neutral, but when dissolved in water release both positively- and negatively-
charged ions. Flocculants may be added to water as a liquid, or more commonly, placed in water
as a gelatinous solid ("floc" logs).
Aeration is the process of introducing air into water. Oxidation occurs when oxygen in
air combines with metals in the water. Once oxidized, Fe and Mn can come out of solution at pH
values of 4 and 7, respectively. If the water is not or only partially oxidized, a solution pH of 8
to 10 is required for precipitation of these same metals. For this reason, aeration of water is a
major limiting factor in many water treatment systems. If aeration and oxidation were
incorporated or improved in the treatment system, chemical treatment efficiency would increase
and costs could be reduced substantially. Oxidants (Table 6.3) are sometimes used to aid in the
completion of the oxidation process to enhance metal hydroxide precipitation and reduce metal
floc volume. The hypochlorite products (bleach), H2O2, and KMnO4 are used in AMD situations
and have demonstrated very effective oxidation. Calcium peroxide has been shown to oxygenate
AMD as well as neutralize acidity (Lilly and Ziemkiewicz, 1992).
After chemical treatment, the treated water flows into sedimentation ponds so metals in
the water can precipitate. Dissolved metals precipitate from AMD as a loose, open-structured
mass of tiny grains called "floc". All chemicals currently used in AMD treatment cause the
formation of metal hydroxide sludge or floc. Sufficient residence time of the water, which is
dictated by pond size and depth, is important for adequate metal precipitation. Hilton (1993)
found pond size to be too small on most AMD treatment sites to result in complete treatment of
the water and precipitation of dissolved metals. The amount of metal floc generated by AMD
neutralization depends on the quality and quantity of water being treated, which in turn
determines how often the ponds must be cleaned. Knowing the chemical and AMD being
treated will help determine the general floc properties and will provide an estimate of the
stability of the various metal compounds in the floc.

Table 6.2. Costs in 2000 of five chemicals to treat acid mine drainage in West Virginia.
The analysis is based on a five-year operation period and includes chemical reagent costs,
installation and maintenance of equipment, and annual operating costs. The “Bulk”
chemical prices in Table 6.1 were used to calculate the reagent costs for only the 189 L/min
flow. The “Bulk” prices were used for higher flows. Neutralization efficiencies were not
included in the reagent cost calculation.
____________________________________________________________________________________________

Flow and Acidity Conditions

16
Flow (L/min) 189 3780 945 3780

Acidity (mg/L) 100 100 500 2500

Chemical

Soda Ash

reagent costs $3,731 $44,000 $58,300 $1,166,000


repair costs 0 0 0 0
annual labor 14,040 14,040 14,040 14,040
installation costs 229 229 229 229
salvage value 0 0 0 0
Net present value 75,052 244,679 245,774 4,911,804
Annual cost $17,817 $58,086 $58,346 $1,166,046

Ammonia

reagent costs $2,543 $22,440 $28,050 $561,000


repair costs 495 495 495 495
tank rental 480 1,200 1,200 1,200
annual labor 7,020 7,020 7,020 7,020
electricity 600 600 600 600
installation costs 1,936 6,357 6,357 6,357
salvage value 0 0 0 0
Net present value 48,547 139,117 162,749 2,407,725
Annual cost $11,525 $33,026 $38,636 $571,586

Caustic Soda (20% Liquid)

reagent costs $5,174 $79,341 $99,176 $1,983,520


repair costs 0 0 0 0
annual labor 7,020 7,020 7,020 7,020
installation costs 283 5,478 5,478 5,478
salvage value 0 0 0 0
Net present value 51,601 368,398 451,950 8,389,433
Annual cost $12,250 $87,457 $107,292 $1,991,636

Pebble Quicklime

reagent costs $1,478 $9,856 $12,320 $246,400


repair costs 500 2,500 2,500 10,000
annual labor 6,500 11,200 11,200 11,200
electricity 0 0 0 0
installation costs 16,000 80,000 80,000 120,000
salvage value 0 5,000 5,000 20,000
Net present value 49,192 162,412 172,790 1,127,220
Annual cost $11,678 $38,556 $41,020 $267,600

17
Hydrated Lime

reagent costs $814 $9,768 $12,210 $244,200


repair costs 1,000 3,100 3,500 10,500
annual labor 6,500 11,232 11,232 11,232
electricity 3,500 11,000 11,000 11,000
installation costs 58,400 102,000 106,000 200,000
salvage value 5,750 6,500 7,500 25,000
Net present value 94,120 228,310 242,809 1,313,970
Annual cost $22,344 $54,200 $57,642 $311,932

Table 6.3. Chemicals for acid neutralization, coagulation/flocculation, and oxidation


___________________________________________________________________________________________

CHEMICAL
NAME FORMULA COMMENTS

Acid Neutralization

Limestone CaCO3 Used in anoxic limestone drains and open limestone channels.

18
Hydrated Lime Ca(OH)2 Cost effective reagent, but requires mixing.
Pebble Quick Lime CaO Very reactive, needs metering equipment.
Soda Ash Briquettes Na2CO3 System for remote locations, but expensive.
Caustic Soda NaOH Very soluble, comes as a solid in drums, beads, or flakes,
or as a 20% or 50% liquid. Cheaper in the liquid form.
Ammonia NH3 or NH4OH Very reactive and soluble; also purchased as aqua ammonia.
Potassium Hydroxide KOH Similar to caustic.
Magnesium Hydroxide Mg(OH)2 Similar to hydrated lime.
Magna Lime MgO Similar to pebble quicklime.
Calcium Peroxide CaO2 Used as a neutralizer and oxidant; either powder or briquettes.
Kiln Dust CaO, Ca(OH)2 Waste product of limestone industry. Active ingredient is CaO
with various amounts of other constituents.
Fly Ash CaCO3,Ca(OH)2 Neutralization value varies with each product.

Coagulants/Flocculants

Alum (aluminum sulfate) Al2(SO4)3 Acidic material, forms Al(OH)3.


Copperas (ferrous sulfate) FeSO4 Acidic material, usually slower reacting than alum.
Ferric Sulfate Fe2(SO4)3 Ferric products react faster than ferrous.
Sodium Aluminate NaAlO2 Alkaline coagulant.

Anionic Flocculants Negatively-charged surface.


Cationic Flocculants Positively-charged surface.
Polyampholytes Both positive and negative charges on surface based on pH.

Oxidants

Calcium Hypochlorite Ca(ClO)2 Strong oxidant.


Sodium Hypochlorite NaClO Also a strong oxidant.
Calcium Peroxide CaO2 Trapzene, an acid neutralizer.
Hydrogen Peroxide H2O2 Strong oxidant.
Potassium permanganate KMnO4 Very effective, commonly used.

Ackman (1982) investigated the chemical and physical characteristics of AMD floc and
concluded that each floc varied depending on the nature of the AMD, the neutralization
chemical, and the mechanical mixing or aeration device used during chemical treatment. He
stated the most important physical property is the floc s settleability, which includes both the
settling rate and final floc volume. Furthermore, Ca(OH)2 and Na2CO3 produced granular, dense
flocs versus a more gelatinous, loose floc generated by NaOH or NH3. The chemical
compositions of flocs were generally composed of hydrated Fe+2 or Fe+3 oxyhydroxides, CaSO4,
Al(OH)3 A6H2O, CaCO3 and Ca(HCO3)2, with trace amounts of Si, PO4, Mn, Cu, and Zn.

19
Payette et al. (1991), using scanning electron microscope analyses, found that AMD
neutralized by Ca(OH)2 resulted in the formation of crystalline CaSO4 as well as various
amorphic metal hydroxides. Amorphic materials are those that have little or no structure and
order, and have a tendency to dissolve more quickly than crystalline materials under the same
conditions. Most minerals, on the other hand, show strong order, structure, and crystallinity.
Amorphic materials may also become ordered or crystalline with time. Payette et al. (1991)
showed that AMD floc was mostly amorphous at one hr after formation, while crystals were
observed in the floc 24 hrs after formation. In a series of experiments on floc generation and
stability, Brown et al. (1994a, b, c) found:

1. More floc was produced as the pH of the AMD solution was increased by chemical addition.
2. Each AMD source was unique in its reaction to four neutralization chemicals.
3. The amount of floc produced as a function of the amount of chemical added (termed its
"efficiency") remained about the same across all pH ranges for Ca(OH)2, NaOH, and Na2CO3.
Ammonia became less efficient at high pH.
4. Sodium carbonate was needed in the highest amount to raise water pH to 7.5 or greater.
5. Floc volumes were lowest with Na2CO3 and highest with Ca(OH)2 after one week of settling.
6. Greater settling time caused floc consolidation.
7. Flocs were composed of metals in ratios similar to the metal ratios of the AMD from which it
was generated.
8. Flocs were primarily amorphic (having no crystalline structure), except for Na2CO3 flocs.
9. Flocs collected from ponds on mined areas showed little similarity in composition to flocs
generated with the same AMD and chemical in the laboratory. The field flocs had soil particles
mixed with the chemical floc.
10. Aging of AMD flocs caused more stable flocs, thereby decreasing their likelihood of
releasing metals (Watzlaf and Casson, 1990). The greater stability of aged flocs remained even
after re-introducing the flocs into acidic solutions. Aging in a dry environment resulted in better
floc stability than flocs aged under water. Aging also caused floc consolidation.

Floc disposal options include: 1) leaving the floc submerged in a pond indefinitely, 2)
pumping or hauling floc from ponds to abandoned deep mines or to pits dug on surface mines,
and 3) dumping floc into refuse piles. Flocs pumped onto the surface of land and allowed to age
and dry is a good strategy for disposal. In its oxidized and dried condition, AMD flocs can
become crystalline and become part of the soil.
Lovett and Ziemkiewicz (1991) estimated NH3 chemical costs for a site in West Virginia
at $72,000 per year and floc handling costs at $486,000 per year. Based on a flow of 378 L/min
for this site, Brown et al. (1994b) estimated that this site generated approximately 77,900 m3 of
floc per year. Dividing $486,000 by 77,900 m3 of floc yields a cost of $6.25 per m3 for floc
handling and disposal on this site. Several mine operators observed that floc handling and
disposal may cost up to $15 per m3. Due to their high water content and the sheer volume of
material, floc handling costs frequently exceed chemical costs by several times.
Each AMD is unique and the chemical treatment of any particular AMD source is site
specific. Each AMD source should be tested with various chemicals by titration tests to evaluate
the most effective chemical for precipitation of the metals. The costs of each AMD treatment
system based on neutralization (in terms of the reagent cost, capital investment and maintenance
20
of the dispensing system) and floc disposal should be evaluated to determine the most cost
effective system.

6.4 PASSIVE TREATMENT OF ACID MINE DRAINAGE

Active chemical treatment of AMD to remove metals and acidity is often an expensive, long-
term proposition. In recent years, a variety of passive treatment systems have been developed
that do not require continuous chemical inputs and that take advantage of naturally-occurring
chemical and biological processes to cleanse contaminated mine waters. Passive technologies
include constructed wetlands, anoxic limestone drains (ALD), successive alkalinity-producing
systems (SAPS), limestone ponds, and open limestone channels (OLC). Natural wetlands are
characterized by water-saturated soils or sediments, with supporting vegetation adapted to
reducing conditions in their rhizosphere. Constructed wetlands are man-made ecosystems that
mimic their natural counterparts. Often they consist of shallow excavations filled with a flooded
gravel, soil, and organic matter to support wetland plants, such as Typha, Juncus, and Scirpus sp.
Treatment depends on dynamic biogeochemical interactions as contaminated water travels
through the constructed wetland. ALDs are abiotic systems consisting of buried limestone cells
that passively generate bicarbonate alkalinity as anoxic water flows through. SAPS combine
treatment concepts from both wetlands and ALDs. Oxygenated water is pre-treated by organic
matter removing O2 and Fe+3, and then the anoxic water flows through an ALD at the base of the
system. Limestone ponds are ponds built over the upwelling of a seep, which are filled with
limestone for treatment. OLCs are surface channels or ditches filled with limestone. Armoring
of the limestone with Fe reduces limestone dissolution by 20 to 50%, so longer channels and
more limestone is required for water treatment.
At their present stage of development, none of the passive systems can be reliably
implemented as a single permanent solution for most AMD problems to meet effluent limits.
Relative to chemical treatment, passive systems require longer retention times and greater space,
provide less certain treatment efficiency, and are subject to failure in the long-term. However, a
great many passive systems have realized successful short-term implementation in the field and
have substantially reduced water treatment costs at many mine sites (Faulkner and Skousen,
1994). Current research seeks to better understand the dynamically complex chemical and
biological mechanisms that occur within passive systems that are responsible for AMD
treatment. Selection of an appropriate passive system is based on water chemistry and flow
(Figure 6.1) and refinements in design are ongoing. As scientists and practitioners improve
treatment predictability and longevity of passive systems, they will play a more important role in
pollution abatement and environmental protection.

6.4.1 Constructed Wetlands

Research demonstrates that biological and chemical processes occurring in freshwater


wetlands improve the chemical composition of AMD introduced into these ecosystems. Huntsman
et al. (1978) and Wieder and Lang (1982) first noted amelioration of AMD following passage
through naturally occurring Sphagnum bogs in Ohio and West Virginia. Studies by Brooks et al.
(1985), Samuel et al. (1988), and Sencindiver and Bhumbla (1988) documented similar phenomena
in Typha marshes. Although evidence suggests that some wetland plants show long-term adaptation
21
to low pH and high metal concentrations (Keith Garbutt, unpublished data, West Virginia
University), AMD eventually degrades the quality of natural wetlands contrary to federal laws
designed for wetland protection and enhancement. Such regulations do not govern use of artificially
constructed wetlands for water treatment, which lead to the suggestion that these engineered systems
might provide passive low cost, low maintenance treatment of AMD (Kleinmann et al., 1983). Over
a thousand wetlands have since been constructed to receive AMD from both active mines and
abandoned mine lands. Highly variable treatment effectiveness and longevity have been observed
(Dietz et al., 1994; Hedin and Nairn, 1992; Hellier et al., 1994; Kleinmann et al., 1991; Wheeler et
al., 1991; Wieder, 1989). Consequently, the US Office of Surface Mining (OSM) will not currently
allow the use of constructed wetlands as the sole on-site method for AMD treatment (OSM, 1988).
However, properly constructed wetlands can reduce effluent concentrations of dissolved metals to
meet effluent limits and have been effectively utilized at many sites to reduce the costs of
conventional chemical treatments (Brodie, 1991).
Most AMD contains dissolved Fe, Mn, and Al. Constructed wetlands successfully
remove Fe, but Mn and Al removal has been more variable (Stark et al., 1996). Mechanisms of
metal retention within wetlands, listed in order of importance, include: 1) formation of metal
oxides and oxyhydroxides, 2) formation of metal sulfides, 3) organic complexation reactions, 4)
exchange with other cations on negatively-charged sites, and 5) direct uptake by living plants.
Other mechanisms may include physical filtration of suspended metal colloids and
adsorption/exchange of metals onto algal mats. Other beneficial reactions in wetlands include
generation of alkalinity due to microbial mineralization of dead organic matter, microbial
dissimilatory reduction of Fe oxyhydroxides and SO4, and dissolution of carbonates.
The way in which a wetland is constructed ultimately affects how water treatment occurs.
Two construction styles currently predominate: 1) "aerobic" wetlands consisting of Typha and
other wetland vegetation planted in shallow (<30cm), relatively impermeable sediments
comprised of soil, clay or mine spoil, and 2) "anaerobic" wetlands consisting of Typha and other
wetland vegetation planted into deep (>30cm), permeable sediments comprised of soil, peat
moss, spent mushroom compost, sawdust, straw/manure, hay bales, or a variety of other organic
mixtures, which are often underlain or admixed with limestone.

6.4.1.1 Aerobic Wetlands

Aerobic wetlands promote metal oxidation and hydrolysis, thereby causing precipitation
and physical retention of Fe, Al, and Mn oxyhydroxides. Successful metal removal depends on
dissolved metal concentrations, dissolved oxygen content, pH and net acidity of the mine water,
the presence of active microbial biomass, and detention time of the water in the wetland. The pH
and net acidity/alkalinity of the water are particularly important because pH influences both the
solubility of metal hydroxide precipitates and the kinetics of metal oxidation and hydrolysis.
Hydrolysis lowers pH, so alkalinity in the water buffers the pH and allows oxidation to continue.

22
Figure 6.1. Flowchart for selecting a passive AMD treatment system based on water
chemistry and flow (adapted from Hedin and Nairn, 1992).

Autooxidation reaction rates decrease a hundred-fold with each unit drop in pH. This
decrease is partially compensated for by microbial Fe oxidation at pH ranges from 1 to 4.
Following Fe oxidation, abiotic hydrolysis reactions precipitate Fe+3. Therefore, aerobic
wetlands are best used in conjunction with water that contains net alkalinity to neutralize metal
acidity. Abiotic Mn oxidation occurs at pH >8 while microorganisms are thought to catalyze this
reaction at pH > 6 (Wildeman et al., 1991). Manganese precipitation occurs much more slowly
than Fe and is sensitive to the presence of Fe+2, which causes chemical reduction of oxidized
Mn. Consequently in aerobic net alkaline water, Fe and Mn precipitate sequentially, not
simultaneously, with the practical result that several staged aerobic wetlands must be constructed
in series if removal of both metals is to be attempted.

23
Brodie and co-workers at the Tennessee Valley Authority (TVA) have reported
extensively on their use of staged aerobic wetlands to treat AMD generated by mine spoils, coal
slurries, and coal ash (Brodie et al., 1993). A typical staged design might include an ALD (see
next section) to passively add alkalinity to the source AMD, a settling basin to hold precipitated
Fe flocs, followed by two or three aerobic wetland cells to sequentially remove additional Fe and
Mn. TVA currently operates nine wetlands receiving moderate quality AMD (total Fe < 70
mg/L; total Mn < 17 mg/L; alkalinity 35 to 300 mg/L as CaCO3), which require no further post-
system treatment of water exiting the wetlands. Four TVA wetlands are associated with high Fe
(>170 mg/L) and low influent alkalinity with resultant production of high metal acidity and low
pH water. Two of these systems require NaOH treatment to comply with NPDES effluent limits,
while two others use ALDs for further treatment of the effluent. A final TVA wetland system
receives low Fe (<0.7 mg/L) and Mn (5.3 mg/L) and is ineffective in Mn removal. Based on
their experience with these systems since 1985, Brodie suggests that staged aerobic wetland
systems can accomodate Fe loads of up to 21 GDM (grams/day/m2) even in the absence of
excess alkalinity. Manganese loads up to 2 GDM can be accomodated if alkalinity is present.
Long-term successful treatment by a staged aerobic wetland also has been reported for
net alkaline water (total Fe 89 mg/L; alkalinity 88 mg/L as CaCO3) at the Simco constructed
wetland near Coshocton, OH (Stark et al., 1994). Analysis of 73 sites in Pennsylvania suggested
that constructed wetlands are the best available technology for many postmining ground water
seeps, particularly those of moderate pH (Hellier et al., 1994).

6.4.1.2 Anaerobic Wetlands

Anaerobic wetlands promote metal oxidation and hydrolysis in aerobic surface layers, but
primarily rely on chemical and microbial reduction reactions to precipitate metals and neutralize
acidity. The water infiltrates through a thick permeable organic subsurface sediment that
becomes anaerobic due to high biological oxygen demand. Several additional treatment
mechanisms function, including metal exchange and complexation reactions, formation and
precipitation of metal sulfides, microbially generated alkalinity due to reduction reactions, and
continuous formation of carbonate alkalinity due to limestone dissolution under anoxic
conditions. Since anaerobic wetlands produce alkalinity, their use can be extended to poor
quality, net acidic, low pH, high Fe, and high dissolved oxygen (>2 mg/L) AMD. However,
Wieder (1992) documents that the mechanism and efficiency of AMD treatment varies
seasonally and with wetland age. Microbial mechanisms of alkalinity production are likely to be
of critical importance to long-term AMD treatment. When wetlands receive high acid loads
(>300 mg/L), the pH sensitive microbial activities are eventually overwhelmed. Therefore, like
their aerobic counterparts, anaerobic wetlands are most successful when used to treat small
AMD flows of moderate water quality. At present, the sizing value for Fe removal in these
wetalnds is 10 GDM (Hedin and Nairn, 1992).
Sorption onto organic materials (such as peat and sawdust) can initially remove 50 to
80% of the metals in AMD (Brodie et al., 1988). Five different substrates (clay, topsoil, mine
spoil, acid wetland soil, and non-acid wetland soil) with vegetation were no different in Fe
removal than these materials without vegetation after one growing season (Brodie et al., 1988).
This finding suggests that a particular substrate material, as long as it has metal retention
capabilities, may not be as important as the conditions created by wetland plants and the
24
presence of an anaerobic zone. Kleinmann et al. (1991) suggested adsorption and complexation
by organic substrates may compensate for limited initial biological activity during the first few
months of operation in a new wetland system. In a laboratory simulation, Henrot and Wieder
(1990) observed that organic complexation in peat mesocosms was a significant retention
mechanism during initial application of AMD, but saturated at 12 mg Fe per gram dry peat.
They found little to no contribution by cation exchange reactions. A subsequent field study,
which examined five wetland substrate types over a 25-month period, also demonstrated that
exchange reactions were an insignificant sink for metal cations regardless of substrate type
(Wieder, 1993). Organic complexation saturated the substrate after only 1 to 7 months of AMD
input at 9 to 17 mg Fe per gram substrate. Although some natural inputs of organic matter occur
annually at plant senescence, the physical filtering capacity of a wetland will ultimately be finite
as all exchange and complexation sites become metal saturated. Substantial artificial inputs of
organic matter have been used as a successful strategy to temporarily renew this filtering
capacity, following an observed decline in wetland performance (Eger and Melchert, 1992;
Haffner, 1992; Stark et al., 1995).
Insoluble precipitates such as oxyhydroxides, carbonates, and sulfides represent a major
sink for metal retention in wetlands. Unlike exchange and complexation reactions, metal
precipitate retention has no theoretical maximum, but is limited in reality by the physical density
and volume of the material produced. For example, compared to diffuse amorphous Fe
oxyhydroxides, Fe mono and disulfides are compact and dense precipitates, which occlude less
wetland pore space volume per unit of Fe retained. The long-term stability of metal precipitates
is unclear. Ferric hydroxides can be reduced with time to Fe+2 by anaerobic Fe-reducing bacteria
in the wetland. Similarly, the use of Fe+3 as a pyrite oxidant under anaerobic conditions would
result in Fe+2 in wetland effluents. About 50 to 70% of the total Fe removed from AMD by
wetlands is found as Fe+3 hydroxides (Henrot and Wieder, 1990; Calabrese et al., 1991; Wieder,
1992). Oxyhydroxide formation depends both on the availability of dissolved oxygen and on the
inital oxidation state of Fe in the AMD. Oxygen must be present for Fe+2 oxidation, whether it
occurs abiotically or by Thiobacillus sp., so AMD often contains a mixture of Fe+2 and Fe+3.
However, Wieder (1993) reported significant retention of Fe+3 oxyhydroxides in anaerobic
wetlands due to the preponderance of dissolved Fe+3 in their source AMD. Once Fe+3 is present,
then oxygen is unnecessary for hydrolysis to the oxyhydroxide. When estimating oxyhydroxide
formation, attention to laboratory procedures are essential to avoid analytical artifacts.
Conventional aerobic extractions (Henrot and Wieder, 1990) can overestimate oxide formation
by 10 to 20% compared with similar extractions performed under anaerobic conditions
(Bhumbla et al., 1991; Calabrese et al., 1994).
Up to 30% of the Fe retained in wetlands may be found as reduced Fe and may be
combined with sulfides (Calabrese et al., 1991; McIntyre and Edenborn, 1990; Wieder, 1992).
Iron mono and disulfides form as a result of H2S formation by microbial sulfate reduction in the
presence of an oxidizable carbon source. In addition to its metal removal potential, sulfate
reduction consumes acidity and raises water pH (Hedin and Nairn, 1992; Rabenhorst et al.,
1992). In naturally occurring freshwater wetlands or lake sediments, sulfate reduction would not
be expected to be a predominate process since sulfate concentrations are usually low in these
ecosystems. High sulfate concentrations (200 to 2000 mg/L) are common in AMD and
availability of this anaerobic electron acceptor results in selective enrichment for sulfate-
reducing bacteria (SRB). Resulting activity of these bacteria in constructed wetlands has been
25
documented at levels comparable to coastal marine sediments (i.e. 250 nmoles per gram per
day)(Calabrese et al., 1991). Acidophilic strains are unknown, therefore constructed wetlands
must provide protected niches for SRB to flourish and persist. Wieder (1992) reported transitory
activity of SRB during the first year of wetland operation with very acidic AMD, but found the
bacteria to be distributed throughout the entire wetland, including near-surface sediments.
McIntyre and Edenborn (1990) and Calabrese et al. (1994) report persistence of SRB activity for
longer periods, particularly in deep sediments adjacent to zones of subsurface limestone
dissolution.
A variety of wetland substrates have been shown to provide sufficient dissolved organic
carbon to support SRB activity, including non-traditional but increasingly available materials
such as composted chicken litter (Gross et al., 1993). Wieder (1992) found greater rates of SRB
activity in straw/manure or mushroom compost wetland substrates than peat. Inorganic FeSx and
organic sulfur (produced by attachment of S-2 or HS- to organic matter) were dominant products
of sulfate reduction in Sphagnum/limestone and straw/manure vs Sphagnum alone and
mushroom compost wetlands, respectively (Wieder and Lang, 1982; Taddeo and Wieder, 1991).
Based on their inorganic S data, Calabrese et al. (1991) and Rabenhorst et al. (1992) estimate
that 30% of the total Fe retained was due to SRB activity. Bhumbla et al. (1991), working with
direct measurement of extractable Fe pools, estimated that 15% of the Fe retained could be
formed as sulfides.
Long-term retention of Fe sulfides and Fe oxyhydroxides in a wetland are not well
understood. Under continued anoxic conditions and in the absence of soluble Fe+3, pyrite should
remain stable. Indeed, Calabrese et al. (1994) changed the influent to their anaerobic wetland
from AMD to freshwater with no concomitant export of Fe+2. Since their effluent pH was
greater than pH 6 due to continued limestone dissolution, dissimilatory Fe reduction in surface
sediments may have been inhibited (Wieder, 1994). Dissimilatory Fe-reducting bacteria (IRB)
are important anaerobic microorganisms that occur in anoxic sediments (Lovely, 1993). These
bacteria use soluble carbon as an energy source and solid Fe+3 oxyhydroxide as an electron
acceptor, producing soluble Fe+2 and alkalinity as important reaction products. If SRB and IRB
compete for the same carbon source in the presence of excess sulfate and Fe+3 oxyhydroxide,
then IRB will competitively exclude SRB as the dominant microbial population (Chapelle and
Lovely, 1992). Since Fe+3 oxyhydroxides commonly accumulate in surface waters of AMD
wetlands adjacent to anoxic subsurface sediments, it is reasonable that such competition occurs.
Vile and Wieder (1993) suggested that nocturnal Fe+2 export resulted both from dissimilatory Fe
reduction and pyrite reoxidation. Both processes were probably inhibited during daylight by
oxygenic photosynthesis of cyanobacteria and green algae resulting in net Fe+3 export, since
influent AMD was 90% Fe+3. Vile and Wieder (1993) note that those SRB and IRB activities
produce significant alkalinity, but that hydrolysis of Fe+3 and pyrite reoxidation are sources of
acidity. In their systems, microbial mechanisms of alkalinity generation were overwhelmed after
two years of AMD exposure. These data suggest that microbial populations, thought to be as
numerous in non-AMD impacted wetlands as in their natural counterparts (Duncan and
Groffman, 1994), may gradually decrease in magnitude and diversity with time of AMD
treatment, though there are many systems that show no decline in performance after several
years. Some workers have indicated that wetland systems can be seeded with functional
microorganisms (Davison, 1993) to introduce or re-establish microbial activity. However,
experiments utilizing appropriate controls have not established the efficacy of this approach.
26
Experience with bioremediation of other wastes suggests that selection and enrichment of
naturally-occurring microbial populations is a superior, more cost-effective approach
(Alexander, 1993).
Calabrese et al. (1991) noted that their experimental wetlands continued to produce
circumneutral effluent high in soluble Fe even after interstitial water within the organic substrate
was significantly acidified. They suggested that the wetland had begun to function in a manner
analogous to an anoxic limestone drain where AMD contained no dissolved oxygen and
predominantly Fe+2 passively obtains carbonate alkalinity during passage through the limestone
bed underlying the wetland. Their data show that heterotrophic oxygen consumption,
dissimilatory Fe reduction and/or pyrite reoxidation continued to function in the wetland, since
influent water contained measurable dissolved oxygen and mixtures of both Fe+3 and Fe+2.
Relatively little work exists on the appropriate selection of plant species for wetlands, yet
this can have important implications for the long-term success of a project. In constructed
wetlands, higher plants serve several purposes including: substrate consolidation, metal
accumulation, physical filtration of metal precipitates, stimulation of microbial processes,
wildlife habitat, and aesthetics.
Wetland plant species vary in their ability to accumulate metals (Fernandes and
Henriques, 1990). Some reports document elevated tissue concentrations (Spratt and Wieder,
1988), while others show little metal accumulation (Folsom et al., 1981). On an annual basis,
uptake by Typha accounted for less than 1% of the Fe removed by volunteer wetlands treating
AMD (Sencindiver and Bhumbla, 1988). While it may be true that metal concentration and
accumulation in plant tissues may be small in any one year, plant tissue is a renewable resource.
Old tissue is senesced yearly and new tissue, with new sites for metal accumulation, is produced.
Thus, over the entire life of a wetland, plants may accumulate a significant portion of metals
received in a wetland.
Several studies report on the effects of different plant species in wetlands. Sphagnum
was the predominant wetland species found in AMD-treating wetlands and Sphagnum has a
well-documented capacity to accumulate Fe (Gerber et al., 1985; Wenerick et al., 1989).
However, Spratt and Wieder (1988) found that saturation of Sphagnum moss with Fe could occur
within one growing season. Some have indicated that metal retention over the long term is
limited in some wetlands because organic matter inputs by wetland plants is limited (Kleinmann,
1990). Many of the original constructed wetlands were vegetated with Sphagnum but few
remained effective. Cattails (Typha) have been found to have a greater environmental tolerance
than Sphagnum moss (Samuel et al., 1988). One of the reasons is cattails do not accumulate
metals into their tissues through uptake. Sencindiver and Bhumbla (1988) found that cattails
growing in AMD-fed wetlands did not have elevated concentrations of metals in plant tissues.
Algae and a few other wetland species have also received attention due to the observation that
enhanced metal removal was associated with algal blooms (Hedin, 1989; Kepler, 1988;
Pesavento and Stark, 1986). In Colorado, algal mixtures were found to aerobically remove Mn
from mine drainage (Duggan et al., 1992), presumably due to elevated pH resulting from algal
growth. Probably the most important role that wetland plants serve in AMD treatment systems
may be their ability to stimulate microbial processes. Kleinmann et al. (1991) explain that plants
provide sites for microbial attachment, release oxygen from their roots, and supply organic
matter for heterotrophs.
With any constructed wetland, it is important to allow the plants to establish for some
27
time with uncontaminated water before AMD is applied to the system. Plants can suffer
"transplantation shock", which can make them particularly susceptible to any environmental
shock. Immediate inundation with AMD will certainly slow growth and development and in
some cases may kill the entire wetland. If a wetland is built during the early to middle part of
the growing season, the plants should be given one to two months to establish prior to AMD
introduction. If a system is constructed late in the growing season, AMD flow should not begin
until the next growing season.
Current practice is to plant a fairly dense monoculture of Typha. As passive treatment
systems get bigger, the number of plants needed and the manpower required to plant them can
lead to prohibitive expense. So some planting designs that do not fill the wetland with plants at
its creation should be considered. Volunteer species are important components in some
constructed wetland systems, but these are not generally part of the initial matrix from which a
wetland will develop.
The choice of wetland species should be based on a knowledge of local conditions and
the relative ability of species to tolerate the conditions found in a particular watershed and their
ability to provide the functions listed above. For example, Typha angustifolia is extremely rare
in West Virginia, although it is common in surrounding states. Thus a wetland planted with T.
angustifolia in West Virginia is unlikely to flourish, while Typha latifolia does well. Similarly
T. angustifolia and T. latifolia have different niches with respect to water depth. In general, the
water depths usually employed in wetland systems are more suitable for T. latifolia than T.
angustifolia. It should also be noted that since AMD is extremely variable, a plant community
that thrives in one location may be completely inappropriate in a second nearby location if the
composition of the AMD is different.

6.4.2 Anoxic Limestone Drains

Anoxic limestone drains (ALDs) are buried cells or trenches of limestone into which
anoxic water is introduced. The limestone dissolves in the acid water, raises pH, and adds
alkalinity. Under anoxic conditions, the limestone does not coat or armor with Fe hydroxides
because Fe+2 does not precipitate as Fe(OH)2 at pH 6.0. ALDs were first described by
theTennessee Division of Water Pollution Control (TDWPC) (Turner and McCoy, 1990).
Tennessee Valley Authority (TVA) subsequently observed that AMD seeping through a coal
refuse dam was being treated passively by limestone contained in an old haul road buried under
the dam. Once the water containing excess alkalinity reaches aerobic conditions at the ground
surface, the metals oxidize and precipitate while the water remains near pH 6 (Brodie et al.,
1990). TVA and TDWPC began building ALDs in 1989. Originally, ALDs were used for pre-
treatment of water flowing into constructed wetlands. Brodie et al. (1993) reported that ALDs
improved the capability of wetlands to meet effluent limitations without chemical treatment.
Since 1990, ALDs have also been constructed as stand-alone systems, particularly where AMD
discharges from deep mine portals. Faulkner and Skousen (1994) reported both successes and
failures among 11 ALDs treating mine water in West Virginia. In all cases, water pH was raised
after ALD treatment but three of the sites had pH values <5.0, indicating that the ALD was not
fully functioning. When working correctly, the pH values of water in ALDs should achieve 6.0.
Water acidity in these drains decreased 50 to 80%, but Fe and Al concentrations in the outflow,
unfortunately, were also decreased. Ferric iron and Al will precipitate as hydroxides at this pH.
28
With Fe and Al decreases in outflow water, it is probable that some coating or clogging of
limestone is occurring inside the ALD.
Longevity of treatment is a major concern for ALDs, especially in terms of water flow
through the limestone. Eventual clogging of the limestone pore spaces with precipitated Al and
Fe hydroxides, and gypsum is predicted (Nairn et al., 1991). For optimum performance, no Fe+3,
dissolved oxygen (DO), or Al should be present in the AMD. Selection of the appropriate water
and environmental conditions is critical for long-term alkalinity generation in an ALD.
Since most AMD has mixed amounts of Fe+3 and Fe+2 and some DO, utilization of an
ALD under these conditions compromises the effectiveness and longevity of the system. Current
research involves pre-treatment of AMD with an anaerobic wetland to strip oxygen from the
water and to convert Fe+3 to Fe+2 (Kepler and McCleary, 1994; Skousen, 1995). The anoxic
water after passing through an anaerobic organic substrate wetland can then be directed
downward into underlying limestone or introduced into an ALD. There are still many questions
relative to the longevity of ALDs and the factors involved in limestone dissolution and metal
precipitation in ALD environments. Like wetlands, ALDs may be a solution for AMD treatment
for specific water conditions or for a finite period after which the system must be replenished or
replaced.

6.4.3 Successive Alkalinity-Producing Systems

Recently, successive alkalinity producing systems (SAPS) have been implemented in the
field (Kepler and McCleary, 1994). Acid water, from 1 to 3 m, is ponded over an organic
compost of 0.2 to 0.3 m, which is underlain by 0.5 to 1 m of limestone. Below the limestone is a
series of drainage pipes that convey the water into an aerobic pond where metals are precipitated.
The hydraulic head drives ponded water through the anaerobic organic compost, where oxygen
stripping as well as Fe and sulfate reduction can occur prior to water entry into the limestone.
Water with high metal loads can be successively cycled through additional SAPS. Iron and Al
clogging of limestone and pipes can be removed by flushing the system (Kepler and McCleary,
1997).

6.4.4 Limestone Ponds

Limestone ponds can be constructed over an AMD upwelling, seep or underground mine
discharge. Limestone is placed in the bottom of the pond and the water flows upward through
the limestone (Skousen et al., 1995). Based on the topography of the area and how the water
emanates from the ground, the pond can be built to pond water from 1 to 3 m deep, containing
0.3 to 1 m of limestone immediately overlying the seep. The pond should be sized and designed
to retain the water for 12 to 24 hrs for limestone dissolution, and to keep the seep and limestone
under water. If limestone coating occurs by Al or Fe hydroxides, the limestone in the pond
could be periodically disturbed with a backhoe to either uncover the limestone or to knock or
scrape off the precipitates. If the limestone is exhausted by dissolution, then a load of fresh
limestone can be dumped into the pond over the seep.

6.4.5 Open Limestone Channels

29
Open limestone channels (OLCs) are another means of introducing alkalinity to acid
water (Ziemkiewicz et al., 1994). Past assumptions have held that armored limestone (limestone
covered or coated with Fe or Al hydroxides) ceases to dissolve. Ziemkiewicz et al. (1997) found
armored limestone to be 50 to 90% effective in neutralizing acid compared to unarmored
limestone, and seven OLCs in the field reduced acidity in AMD by 4 to 62% compared to a 2%
acid reduction in a sandstone channel. Open limestone channels show promise for neutralizing
AMD in watershed restoration projects and AML reclamation project where one-time
installation costs are incurred, little to no maintenance is required, and systems do not have to
meet specific water quality standards. Long channels of limestone can be used to convey acid
water to a stream or other discharge point. Based on flows and acidity concentrations, cross
sections of stream channels (widths and heights) can be designed with calculated amounts of
limestone (which will become armored) to treat the water. Open limestone channels work best
where the channel is constructed on steep slopes (>20%) and where flow velocities keep metal
hydroxides in suspension, thereby limiting their precipitation and plugging of limestone pores in
the channel. Utilizing OLCs with other passive systems can maximize treatment and metal
removal. If constructed correctly, OLCs should be maintenance free and provide AMD
treatment for decades.

6.5 SUMMARY

Acid mine drainage occurs when geologic materials containing metal sulfides are
exposed to oxidizing conditions. Subsequent leaching of reaction products into surface waters
pollute over 20,000 km of streams in the USA alone. Mining companies must predict the
potential of creating AMD on their operations before disturbance by using overburden analyses.
On sites where a potential exists, special handling of overburden materials and quick coverage of
acid-producing materials in the backfill should be practiced. Alkaline addition with materials
such as kiln dust and FBC ash has shown favorable results in reducing or completely eliminating
AMD problems. Placing acid-producing materials under dry barriers effectively isolates these
materials from air and water. While not practiced much in the USA, placing acid-producing
materials under water has shown good success in other regions where complete inundation is
assured. Bactericides are useful in preventing AMD temporarily until a long-term control
technology like a barrier is emplaced. Since many sites have produced AMD for decades before
laws were passed, much effort has been concentrated on controlling AMD from abandoned mine
sites. Injection of alkaline materials into underground mines and buried pods of acid material in
surface mine backfills, remining of abandoned areas, and alkaline recharge trenches are AMD
control technologies being tested currently.
Chemicals used for treating AMD after formation are Ca(OH)2, CaO, NaOH, Na2CO3,
and NH3. Each chemical has advantages for certain water conditions and treatment. Caustic is
generally used when Mn is the primary element to be removed from the water. Under low flow
situations, all of the chemicals except Ca(OH)2 are cost effective. Under high flow situations,
Ca(OH)2 and CaO are clearly the most cost effective due to low reagent cost compared to the
other chemicals. Coagulants, flocculants, and oxidants are used in water treatment to meet more
stringent effluent limits and to aid in floc settling efficiency. Floc, the metal hydroxides
collected in ponds after chemical treatment, are disposed of in abandoned deep mines, refuse
piles, or left in collection ponds. Studies show that flocs are relatively stable materials and
30
metals contained therein do not resolubilize after disposal, especially if aged and dried.
Wetlands treat AMD by removing metals through: 1) formation of oxyhydroxides,
formation of metal sulfides, exchange and organic complexation reactions, and direct plant
uptake. Aerobic wetlands are used when water contains enough alkalinity to promote metal
precipitation. Anaerobic wetlands are designed when AMD contains no alkalinity, and alkalinity
is generated in these systems by microbial sulfate reduction and limestone dissolution in
anaerobic sediments. Anoxic limestone drains use limestone under anoxic conditions to generate
alkalinity. Under anoxia, limestone theoretically will not be coated or covered with Fe+3
hydroxides in the drain, decreasing the likelihood of clogging. Successive alkalinity-producing
systems pre-treat oxygenated AMD with organic matter to remove oxygen and Fe+3. The water
is then introduced into an ALD or limestone underneath the organic matter. Open limestone
channels use limestone in aerobic environments to treat AMD. Coating of limestone occurs but
the limestone continues to dissolve and the reduced effectiveness of limestone dissolution is
designed into the AMD treatment system. At present, most passive systems offer short-term
treatment possibilities after which the system must be recharged or replaced. Further, few
passive systems attain NPDES effluent limits and therefore are more practical for installation on
abandoned sites or watershed restoration projects where effluent limits do not apply and where
some removal of acid and metals will benefit stream restoration.

6.6 ACKNOWLEDGMENTS

The authors thank the editors, John Sencindiver, Keith Garbutt, and two anonymous
reviewers for helpful comments during the review process. Acid mine drainage research at West
Virginia University is supported by grants from the National Mine Land Reclamation Center, the
USDI Bureau of Mines, the West Virginia Division of Environmental Protection, and from funds
appropriated by the Hatch Act. This scientific paper #2544 is a contribution of the West
Virginia Agricultural and Forestry Experiment Station, Morgantown, WV.

6.7 REFERENCES

Ackman, T. 1982. Sludge disposal from acid mine drainage treatment. U.S. Bureau of Mines,
Report of Invest. 8672, Pittsburgh, PA. 38 pp.

Alexander, M. 1993. Biodegradation and bioremediation. Academic Press, San Diego, CA.

Bell, A.V., M.D. Riley, and E.G. Yanful. 1994. Evaluation of a composite soil cover to control
acid waste rock pile drainage. p. 113-121. In: International Land Reclamation and Mine
Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06B-94, Pittsburgh, PA.

Bhumbla, D.K., J.P. Calabrese, J.C. Sencindiver, A.J. Sexstone, and G.K. Bissonnette. 1991.
Attenuation of Fe and Mn in AMD applied to constructed wetlands. Agronomy Abstracts, p. 33.

Brady, K. and R.J. Hornberger. 1989. Mine drainage prediction and overburden analysis in
Pennsylvania. In: Proceedings, Ninth Annual West Virginia Surface Mine Drainage Task Force
Symposium, 25-26 April 1989, West Virginia University, Morgantown, WV.
31
Brady, K. and R.J. Hornberger. 1990. A manual for premining prediction of coal mine drainage
quality. Pennsylvania Department of Environmental Resources, Harrisburg, PA.

Brady, K., E.F. Perry, R.L. Beam, D.C. Bisko, M.D. Garner, and J.M. Tarantino. 1994.
Evaluation of acid-base accounting to predict the quality of drainage at surface coal mines in
Pennsylvania, U.S.A. p. 138-147. In: Internation Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06A-94. Pittsburgh, PA.

Brady, K., M.W. Smith, R.L. Beam, and C.A. Cravotta. 1990. Effectiveness of the use of
alkaline materials at surface coal mines in preventing or abating acid mine drainage: Part 2. Mine
site case studies. p. 227-241. In: Skousen et al. (eds.), Proceedings, 1990 Mining and
Reclamation Conference, 23-26 April 1990, West Virginia University, Morgantown, WV.

Brodie, G.A. 1991. Achieving compliance with staged, aerobic, constructed wetlands to treat
acid mine drainage. p. 151-174. In: Proceedings, 1991 American Society for Surface Mining
and Reclamation Conference, 14-17 May 1991, Durango, CO.

Brodie, G.A., C.R. Britt, H. Taylor, T. Tomaxzewski, and D. Turner. 1990. Passive anoxic
limestone drains to increase effectiveness of wetlands acid drainage treatment systems. In:
Proceedings, Twelfth National Association of Abandoned Mined Land Conference, 16-18 June
1990, Breckinridge, CO.

Brodie, G.A., C.R. Britt, T.M. Tomaszewski, and H.N. Taylor. 1993. Anoxic limestone drains
to enhance performance of aerobic acid drainage treatment wetlands: experiences of the
Tennessee Valley Authority. p. 129-145. In: G.A. Moshiri, (ed.), Constructed Wetlands for
Water Quality Improvement. Lewis Publishers, Boca Raton, FL. 632 pp.

Brodie, G.A., D.A. Hammer, and D.A. Tomljanovich. 1988. An evaluation of substrate types in
constructed wetlands acid drainage treatment systems, p. 389-398. In: Mine Drainage and
Surface Mine Reclamation, 19-21 April 1988, Vol. 1, Info. Circular 9183, U.S. Bureau of Mines,
Pittsburgh, PA.

Brooks, R., D.E. Samuel, and J.B. Hill. 1985. Wetlands and water management on mined lands.
Proceedings of a Conference, 23-24 Oct 1985, The Pennsylvania State University, University
Park, PA.

Brown, H., J. Skousen, and J. Renton. 1994a. Floc generation by chemical neutralization of acid
mine drainage. Green Lands 24(1): 33-51.

Brown, H., J. Skousen, and J. Renton. 1994b. Volume and composition of flocs from chemical
neutralization of acid mine drainage. Green Lands 24(2): 30-35.

Brown, H., J. Skousen, and J. Renton. 1994c. Stability of flocs produced by chemical
neutralization of acid mine drainage. Green Lands 24(3): 34-39.
32
Burnett, J.M., M. Burnett, P. Ziemkiewicz, and D.C. Black. 1995. Pneumatic backfilling of coal
combustion residues in underground mines. In: Proceedings, Sixteenth Annual Surface Mine
Drainage Task Force Symposium, 4-5 April 1995, West Virginia University, Morgantown, WV.

Calabrese, J.P., A.J. Sexstone, D.K. Bhumbla, G.K. Bissonnette, and J.C. Sencindiver. 1991.
Application of constructed cattail wetlands for the removal of iron from acid mine drainage. p.
559-575. In: Proceedings, Second International Conference on the Abatement of Acidic
Drainage, 16-18 Sept. 1991, Vol. 3, MEND, Montreal, Canada.

Calabrese, J.P., A.J. Sexstone, D.K. Bhumbla, J.G. Skousen, G.K. Bissonnette, and J.C.
Sencindiver. 1994. Long-term study of constructed model wetlands for treatment of acid mine
drainage. p. 406. In: International Land Reclamation and Mine Drainage Conference, 24-29
April 1994, USDI, Bureau of Mines SP 06B-94. Pittsburgh, PA.

Caruccio, F.T., L.R. Hossner, and G. Geidel. 1988. Pyritic materials: acid drainage, soil acidity,
and liming. p. 159-190. In: Reclamation of Surface-Mined Lands. Vol 1. CRC Press, Inc. Boca
Raton, FL. 250 pp.

Caruccio, F.T. and G. Geidel. 1989. Water management strategies in abating acid mine
drainage - Is water diversion really beneficial? In: 1989 Multinational Conference on Mine
Planning and Design. 16-17 Sept. 1989, University of Kentucky, Lexington, KY.

Caruccio, F.T., G. Geidel, and R. Williams. 1984. Induced alkaline recharge zones to mitigate
acid seeps. p. 43-47 In: Proceedings, National Symposium on Surface Mining, Hydrology,
Sedimentology and Reclamation, 7-10 Dec. 1984, Univ. of Kentucky, Lexington, KY.

Chapelle, F.H. and D.R. Lovely. 1992. Competitive exclusion of sulfate reduction by iron (III)-
reducing bacteria: A mechanism for producing discrete zones of high iron ground water. Ground
Water 30:29-36.

Dave, N.K. and A.J. Vivyurka. 1994. Water cover on acid generating uranium tailings-
laboratory and field studies. p. 297-306. In: International Land Reclamation and Mine
Drainage Conference, 24-29 April 1994. USDI, Bureau of Mines SP 06A-94, Pittsburgh, PA.

Davison, J. 1993. Successful acid mine drainage and heavy metal site bioremediation. p. 167-
170. In: G.A. Moshiri, (Ed.), Constructed Wetlands for Water Quality Improvement. Lewis
Publishers, Boca Raton, FL. 638 pp.

Dietz, J.M., R.G. Watts, D.M. Stidinger. 1994. Evaluation of acidic mine drainage treatment in
constructed wetland systems. p. 70-79. In: International Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06A-94, Pittsburgh, PA.

diPretoro, R.S. and H.W. Rauch. 1988. Use of acid-base accounts in premining prediction of
acid drainage potential: a new approach from northern West Virginia. p. 2-10. In: Mine
33
Drainage and Surface Mine Reclamation, 19-21 April 1988, USDI, Bureau of Mines Information
Circular 9183. Pittsburgh, PA.

Duggan, L.A., T.R. Wildeman, and D.M. Updegraff. 1992. The aerobic removal of manganese
from mine drainage by an algal mixture containing Cladophora. p. 241-248. In: Proceedings,
1992 American Society for Surface Mining and Reclamation Conference, 14-18 June 1992,
Duluth, MN.

Duncan, C.P. and P.M. Groffman. 1994. Comparing microbial parameters in natural and
constructed wetlands. J. Environ. Qual. 23: 298-305.

Eger, P. and G. Melchert. 1992. The design of a wetland treatment system to remove trace
metals from mine drainage. p. 98-107. In: Proceedings, 1992 American Society for Surface
Mining and Reclamation Conference, 14-18 June 1992, Duluth, MN.

Evangelou, V.P. 1995. Pyrite oxidation and its control. CRC Press, Boca Raton, FL. 293 pp.

Faulkner, B.B. 1990. Handbook for the use of ammonia in treating mine waters. West Virginia
Mining and Reclamation Association, Charleston, WV.

Faulkner, B.B. and J.G. Skousen. 1994. Treatment of acid mine drainage by passive treatment
systems. p. 250-257. In: International Land Reclamation and Mine Drainage Conference, 24-29
April 1994, USDI, Bureau of Mines, SP 06A-94, Pittsburgh, PA.

Fernandez, J.C. and F.F. Henriques. 1990. Metal levels in soil and cattails (Typha latifolia L.)
plants in a pyrite mine, Lousal Portugal. Int. J. of Environ. Studies 36:205-210.

Folsom, B.L., C.R. Lee, and D.J. Bates. 1981. Influence of disposal environment on availability
and plant uptake of heavy metals in dredged material. In: Tech. Report EL-81-12, U.S. Army
Corps of Engineers, Waterway Exp. Stat., ICE, Vicksburg, MS.

Fraser, W.W. and J.D. Robertson. 1994. Subaqueous disposal of reactive mine waste: an
overview and update of case studies-MEND/Canada. p. 250-259. In: International Land
Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines SP
06A-94. Pittsburgh, PA.

Gabr, M.A., J.J. Bowders, and M.S. Runner. 1994. Assessment of acid mine drainage
remediation schemes on groundwater flow regimes at a reclaimed mine site. p. 168-177. In:
International Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI,
Bureau of Mines SP 06B-94, Pittsburgh, PA.

Gerber, D.W., J.E. Burris, and R.W. Stone. 1985. Removal of dissolved iron and manganese
ions by a Sphagnum moss system. In: Wetlands and Water Management on Mined Lands, 23-
24 Oct. 1985, The Pennsylvania State University, University Park, PA.

Gross, M.A., S.J. Formica, L.C. Gandy, and J. Hestir. 1993. A comparison of local waste
34
materials for sulfate-reducing wetlands substrate. p. 179-185. In: G.A. Moshiri (ed.),
Constructed Wetlands for Water Quality Improvement. Lewis Publishers, Boca Raton, FL. 638
pp.

Guebert, M. and T. Gardner. 1992. Macropore flow effect on infiltration, throughflow, and
surface runoff on a reclaimed surface-mined watershed. p. 91-100. In: Proceedings,
International Symposium: Land Reclamation, 14-15 Dec. 1992, American Society of
Agricultural Engineers, Nashville, TN.

Haffner, W.M. 1992. Palmerton zinc superfund site constructed wetlands. p. 260-267. In:
Proceedings, 1992 American Society for Surface Mining and Reclamation, 14-18 June 1992,
Duluth, MN.

Hawkins, J.W. 1994. Assessment of contaminant load changes caused by remining abandoned
coal mines. p. 20-29. In: International Land Reclamation and Mine Drainage Conference, 24-29
April 1994, USDI, Bureau of Mines SP 06A-94, Pittsburgh, PA.

Hedin, R.S. 1989. Treatment of acid coal mine drainage with constructed wetlands. p. 349-362.
In: Wetlands Ecology, Productivity and Values: Emphasis on Pennsylvania. Pennsylvania
Academy of Science, Easton, PA.

Hedin, R.S. and R.W. Nairn. 1992. Passive treatment of coal mine drainage. Course Notes for
Workshop. U.S. Bureau of Mines, Pittsburgh, PA.

Hellier, W.W., E.F. Giovannitti, and P.T. Slack. 1994. Best professional judgment analysis for
constrcuted wetlands as a best available technology for the treatment of post-mining
groundwater seeps. p. 60-69. In: International Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06A-94. Pittsburgh, PA.

Henrot, J. and R.K. Wieder. 1990. Processes of iron and manganese retention in laboratory peat
microcosms subjected to acid mine drainage. J. Environ. Qual. 19:312-320.

Hilton, T. 1990. Handbook - Short Course for Taking A Responsible Environmental Approach
towards Treating Acid Mine Drainage with Anhydrous Ammonia. West Virginia Mining and
Reclamation Association, Charleston, WV.

Hilton, T. 1993. Technical information for fighting acid mine drainage. In: Proceedings,
Fourteenth West Virginia Surface Mine Drainage Task Force Symposium, 27-28 April 1993,
West Virginia University, Morgantown, WV.

Huntsman, B.E., J.B. Solch, and M.D. Porter. 1978. Utilization of a Sphagnum species
dominated bog for coal acid mine drainage abatement. Abstracts, 91st Annual Meeting Geologic
Society America, Ottawa, Ontario, Canada.

Jenkins, M. and J. Skousen. 1993. Acid mine drainage treatment with the Aquafix System. In:
Proceedings, Fourteenth Annual West Virginia Surface Mine Drainage Task Force Symposium,
35
27-28 April 1993, West Virginia University, Morgantown, WV.

Kepler, D.A. 1988. An overview of the role of algae in the treatment of acid mine drainage. p.
286-290. In: Mine Drainage and Surface Mine Reclamation, 19-21 April 1988, Info. Circular
No. 9183, USDI, Bureau of Mines, Pittsburgh, PA.

Kepler, D.A. and E. McCleary. 1994. Successive alkalinity-producing systems (SAPS) for the
treatment of acidic mine drainage. p. 194-204. In: International Land Reclamation and Mine
Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06A-94. Pittsburgh, PA.

Kepler, D.A. and E. McCleary. 1997. Passive aluminum treatment successes. In: Proceedings,
Eighteenth Annual West Virginia Surface Mine Drainage Task Force Symposium, 15-16 April
1997, West Virginia University, Morgantown, WV.

Kim, A.G. and T.E. Ackman. 1994. Disposing of coal combustion residues in inactive surface
mines: effects on water quality. p. 228-236. In: International Land Reclamation and Mine
Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06B-94, Pittsburgh, PA.

Kleinmann, R.L.P. 1990. Acid mine drainage in the United States. In: Proceedings, First
Midwestern Region Reclamation Conference, June 1990, Southern Illinois University,
Carbondale, IL.

Kleinmann, R.L.P., H. Edenborn, and R.S. Hedin. 1991. Biological treatment of mine water--an
overview. p. 27-42. In: Proceedings, Second International Conference on the Abatement of
Acidic Drainage, 16-18 Sept. 1991, MEND, Montreal, Canada.

Kleinmann, R.L.P., T. Tiernman, J.G. Selch, and R.L. Harris. 1983. A low cost low
maintenance treatment system for acid mine drainage using Sphagnum moss and limestone. p.
241-245. In: 1983 Symposium on Surface Mining, Hydrology, Sedimentology, and
Reclamation, 5-9 Dec. 1983, Univ. of Kentucky, Lexington, KY.

Leathen, W.W., S.A. Braley, and L.D. McIntyre. 1953. The role of bacteria in the formation of
acid from certain sulfuritic constituents associated with bituminous coal. Applied Microbiology
1: 61-68.

Lilly, R. and P. Ziemkiewicz. 1992. Manganese removal at a lower pH with calcium peroxide:
results of field trials. In: Proceedings, Thirteenth Annual West Virginia Surface Mine Drainage
Task Force Symposium, 8-9 April 1992, West Virginia University, Morgantown, WV.

Lovely, D.R. 1993. Dissimilatory metal reduction. Ann. Review Microbiol. 47:263-290.

Lovett, R. and P. Ziemkiewicz. 1991. Calcium peroxide for treatment of acid mine drainage. p.
35-46. In: Proceedings, Second International Conference on the Abatement of Acidic Drainage,
16-18 Sept. 1991, MEND, Montreal, Canada.

Lusardi, P.J. and P.M. Erickson. 1985. Assessment and reclamation of an abandoned acid-
producing strip mine in northern Clarion County, Pennsylvania. p. 313-321. In: 1985
36
Symposium on Surface Mining, Hydrology, and Reclamation, 6-10 Dec. 1985, Univ. of
Kentucky, Lexington, KY.

McIntyre, P.E. and H.M. Edenborn. 1990. The use of bacterial sulfate reduction in the treatment
of drainage from coal mines. p. 409-415. In: Proceedings, 1990 Mining and Reclamation
Conference, 23-26 April 1990, West Virginia University, Morgantown, WV.

Meek, F.A. 1994. Evaluation of acid prevention techniques used in surface mining. p. 41-48.
In: International Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI,
Bureau of Mines SP 06B-94, Pittsburgh, PA.

Nairn, R.W., R.S. Hedin, and G.R. Watzlaf. 1991. A preliminary review of the use of anoxic
limestone drains in the passive treatment of acid mine drainage. In: Proceedings, Twelfth
Annual West Virginia Surface Mine Drainage Task Force Symposium, 3-4 April 1991, West
Virginia University, Morgantown, WV.

Nawrot, J.R., P.S. Conley, and J.E. Sandusky. 1994. Concentrated alkaline recharge pools for
acid seep abatement: principles, design, construction and performance. p. 382-391. In:
International Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI,
Bureau of Mines SP 06A-94, Pittsburgh, PA.

Nicholson, R.V., R.W. Gillham, J.A. Cherry, and E.J. Reardon. 1989. Reduction of acid
generation in mine tailings through the use of moisture-retaining cove layers as oxygen barriers.
Can. Geotech. J. 26(1): 1-9.

Office of Surface Mining. 1988. Use of wetland treatment systems for coal mine drainage.
OSMRE Dir. TSR-10, U.S. OSMRE, Washington, D.C.

Parisi, D., J. Horneman, and V. Rastogi. 1994. Use of bactericides to control acid mine drainage
from surface operations. p. 319-325. In: International Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06B-94, Pittsburgh, PA.

Payette, C., W. Lam, C. Angle, and R. Mikula. 1991. Evaluation of improved lime
neutralization processes. p. 15-29. In: Proceedings, Second International Conference on the
Abatement of Acidic Drainage, 16-18 Sept. 1991, MEND, Montreal, Canada.

Pederson, T.F., J.J. McNee, B. Mueller, D.H. Flather, and C.A. Pelletier. 1994. Geochemistry
of submerged tailings in Anderson Lake, Manitoba: recent results. p. 288-296. In:
International Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI,
Bureau of Mines SP 06A-94, Pittsburgh, PA.

Perry, E.F. and K.B. Brady. 1995. Influence of neutralization potential on surface mine
drainage quality in Pennsylvania. In: Proceedings, Sixteenth Annual Surface Mine Drainage
Task Force Symposium, 4-5 April 1995, West Virginia University, Morgantown, WV.

Pesavento, B.G. and L.R. Stark. 1986. Utilization of wetlands and some aspects of aqueous
conditions. In: Proceedings, Seventh Annual West Virginia Surface Mine Drainage Task Force
37
Symposium, 1-2 April 1986, West Virginia University, Morgantown, WV.

Phipps, T.T., J.J. Fletcher, and J.G. Skousen. 1991. A methodology for evaluating the costs of
alternative AMD treatment systems. In: Proceedings, Twelfth Annual West Virginia Surface
Mine Drainage Task Force Symposium, 3-4 April 1991, West Virginia University, Morgantown,
WV.

Rabenhorst, M.C., B.R. James, and J.N. Shaw. 1992. Evaluation of potential wetland substrates
for optimizing sulfate reduction. p. 90-97. In: Proceedings, 1992 American Society for Surface
Mining and Reclamation, 14-18 June 1992, Duluth, MN.

Rasmuson, A. and J.C. Eriksson. 1986. Capillary barriers in covers for mine tailings. National
Swedish Environmental Protection Board Report 3307. Stockholm, Sweden.

Rastogi, V. 1994. Control of mine water quality in coal rejects disposal areas and coal
stockpiles with bactericides. Fifth International Mine Water Congress, Nottingham, U.K.

Rich, D.H. and K.R. Hutchison. 1994. Coal refuse disposal using engineering design and lime
chemistry. p. 392-399. In: International Land Reclamation and Mine Drainage Conference, 24-
29 April 1994, USDI, Bureau of Mines SP 06A-94, Pittsburgh, PA.

Rich, D.H. and K.R. Hutchison. 1990. Neutralization and stabilization of combined refuse using
lime kiln dust at High Power Mountain. p. 55-59. In: Proceedings, 1990 Mining and
Reclamation Conference, 23-26 April 1990, West Virginia University, Morgantown, WV.

Rose, A.W., L.B. Phelps, R.R. Parizek, and D.R. Evans. 1995. Effectiveness of lime kiln flue
dust in preventing acid mine drainage at the Kauffman surface coal mine, Clearfield County,
Pennsylvania. p. 159-171. In: Proceedings, 1995 National Meeting of the American Society for
Surface Mining and Reclamation, 3-8 June 1995, Gillette, WY.

Rymer, T.E., J.J. Renton, and P.F. Ziemkiewicz. 1991. Isolation of critical predictive acid
producing parameters from variable field data using advanced computer technology. p. 141-158.
In: Proceedings, Second International Conference on the Abatement of Acidic Drainage, 16-18
Sept. 1991, MEND, Montreal, Canada.

Samuel, D.E., J.C. Sencindiver, and H.W. Rauch. 1988. Water and soil parameters affecting
growth of cattails. p. 367-374. In: Mine Drainage and Surface Mine Reclamation, 19-21 April
1988, Info. Circular No. 9183, USDI, Bureau of Mines, Pittsburgh, PA.

Schueck, J. 1990. Using a magnetometer for investigating underground coal mine fires, burning
coal refuse banks and for locating AMD source areas in surface mines. p. 493-504. In:
Proceedings, 1990 Mining and Reclamation Conference, 23-26 April 1990, West Virginia
University, Morgantown, WV.

Schueck, J., T. Ackman, and B. Scheetz. 1994. Acid mine drainage abatement using fluidized
bed combustion ash grout after geophysical site characterization. p. 218-227. In: International
Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines
38
SP 06B-94. Pittsburgh, PA.

Sencindiver, J.C. and D.K. Bhumbla. 1988. Effects of cattails (Typha) on metal removal from
mine drainage. p. 359-366. In: Mine Drainage and Surface Mine Reclamation, 19-21 April
1988, Info. Circular No. 9183, USDI, Bureau of Mines, Pittsburgh, PA.

Singer, P.C. and W. Stumm. 1968. Kinetics of the oxidation of ferrous iron. p. 12-34. In: 2nd
Symp. on Coal Mine Drainage Research. Bituminous Coal Research, Inc., Monroeville, PA.

Skousen, J. 1995. Douglas abandoned mine land project: description of an innovative acid mine
drainage treatment system. Green Lands 25(1):29-38.

Skousen, J., D. Bhumbla, J. Gorman, and J. Sencindiver. 1997. Hydraulic conductivity of ash
mixtures and metal release upon leaching. p. 480-495. In: 1997 National Meeting of the
American Society for Surface Mining and Reclamation, 10-15 May 1997, Austin, TX.

Skousen, J.G., B.B. Faulkner, and P. Sterner. 1995. Passive treatment systems and
improvement of water quality. In: Proceedings, Fifteenth Annual West Virginia Surface Mine
Drainage Task Force Symposium, 4-5 April 1995, West Virginia University, Morgantown, WV.

Skousen, J.G., R. Hedin, and B.B. Faulkner. 1997. Water quality changes and costs of remining
in Pennsylvania and West Virginia. p. 64-73. In: 1997 National Meeting of the American
Society for Surface Mining and Reclamation, 10-15 May 1997, Austin, TX.

Skousen, J. and G. Larew. 1994. Alkaline overburden addition to acid-producing materials to


prevent acid mine drainage. p. 375-381. In: International Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06B-94. Pittsburgh, PA.

Skousen, J., R. Lilly, and T. Hilton. 1993. Special chemicals for treating acid mine drainage.
Green Lands 23(3): 34-41.

Skousen, J., K. Politan, T. Hilton, and A. Meek. 1990. Acid mine drainage treatment systems:
chemicals and costs. Green Lands 20(4): 31-37.

Skousen, J.G., J.C. Sencindiver, and R.M. Smith. 1987. A Review of Procedures for Surface
Mining and Reclamation in Areas with Acid-Producing Materials. EWRC 871, West Virginia
University, Morgantown, WV. 40 pp.

Skousen, J. and P. Ziemkiewicz. 1996. Acid Mine Drainage Control and Treatment. 2nd Ed.
National Research Center for Coal and Energy, National Mine Land Reclamation Center, West
Virginia University, Morgantown, WV. 362 pp.

Smith, E.E. and K.S. Shumate. 1970. Sulfide to sulfate reaction mechanism. Water Pollution
Control Res. Series 14010 FPS. USDI, FWQA. Washington, D.C.

Smith, R.M. and A.A. Sobek. 1978. Physical and chemical properties of overburdens, spoils,
wastes, and new soils. p. 149-172. In: Reclamation of Drastically Disturbed Lands. American
39
Society of Agronomy, Madison, WI.

Splittorf, D. and V. Rastogi. 1995. Ten year results from bactericide-treated and reclaimed mine
land. p. 471-478. In: Proceedings, 1995 American Society for Surface Mining and Reclamation
Annual Meeting, 3-8 June 1995, Gillette, WY.

Spratt, A.K. and R.K. Wieder. 1988. Growth responses and iron uptake in Sphagnum plants and
their relation to acid mine drainage treatment. p. 279-285. In: Mine Drainage and Surface Mine
Reclamation, 19-21 April 1988, Info. Circular No. 9183, USDI, Bureau of Mines, Pittsburgh,
PA.

Stark, L.R., F.M. Williams, S.E. Stevens, and D.P. Eddy. 1994. Retention and vegetative cover
at the Simco constructed wetland: an appraisal through year eight of operation. p. 89-98. In:
International Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI,
Bureau of Mines SP 06A-94. Pittsburgh, PA.

Stark, L.R., F.M. Williams, W.R. Wenerick, P.J. Wuest, and C.A. Urban. 1995. The effects of
carbon supplementation and plant species on iron retention in mesocosm treatment wetlands.
Wetlands 15:58-67.

Stark, L.R., F.M. Williams, W.R. Wenerick, P.J. Wuest, and C.A. Urban. 1996. The effects of
substrate type, surface water depth, and flow rate on manganese retention in mesocosm wetlands.
J. Environ. Qual. 25:97-106.

Stehouwer, R., P. Sutton, R. Fowler, and W. Dick. Minespoil amendment with dry flue gas
desulfurization by-products: element solubility and mobility. J. Environ. Qual. 24:165-174.

Stumm, W. and J.J. Morgan. 1970. Aquatic Chemistry. John Wiley and Sons, Inc., New York.

Taddeo, F.J. and R.K. Wieder. 1991. The accumulation of iron sulphides in wetlands
constructed for acid coal mine drainage (AMD) treatment. p. 529-548. In: Proceedings, Second
International Conference on the Abatement of Acidic Drainage, 16-18 Sept. 1991, MEND,
Montreal, Canada.

Temple, K.L. and W.A. Koehler. 1954. Drainage from bituminous coal mines. West Virginia
University Bulletin. Engineering Experiment Station Research Bull. 25, Morgantown, WV.

Turner, D. and D. McCoy. 1990. Anoxic alkaline drain treatment system, a low cost acid mine
drainage treatment alternative. p. 73-76. In: Proceedings, National Symposium on Mining, 14-
18 May 1990, Univ. of Kentucky, Lexington, KY.

US Environmental Protection Agency. 1983. Neutralization of Acid Mine Drainage, Design


Manual. USEPA-600/2-83-001, Cincinnati, OH. U.S. Govt. Printing Office, Washington, DC.

US Environmental Protection Agency. 1971. Inorganic sulfur oxidation by iron-oxidizing


bacteria. Water Pollution Control Res. Series 14010 DAY. USDI, FWQA. Washington, D.C.

40
Vile, M.A. and R.K. Wieder. 1993. Alkalinity generation by Fe(III) reduction versus sulfate
reduction in wetlands constructed for acid mine drainage treatment. Water, Air, and Soil
Pollution 69: 425-441.

Waksman, S.A. 1922. Microorganisms concerned in the oxidation of sulfur in the soil IV. A
solid medium for the isolation and cultivation of Thiobacillus thiooxidans. J. Bact. 7:605-608.

Watzlaf, G. and L. Casson. 1990. Chemical stability of manganese and iron in mine drainage
treatment sludge: effects of neutralization chemical, iron concentration, and sludge age. p. 3-9.
In: Proceedings, 1990 Mining and Reclamation Conference, 23-26 April 1990, West Virginia
University, Morgantown, WV.

Wenerick, W.R., S.E. Stevens, H.J. Webster, L.R. Stark, and E. DeVeau. 1989. Tolerance of
three wetland plant species to acid mine drainage: a greenhouse study. p. 801-807. In: D.
Hammer (ed.), Constructed Wetlands for Wastewater Treatment, Lewis Publishers, Chelsea, MI.
831 pp.

Wheeler, W.N., M. Kalin, and J.E. Cairns. 1991. The ecological response of a bog to acidic coal
mine drainage - deterioration and subsequent initiation of recovery. p. 449-464. In:
Proceedings, Second International Conference on the Abatement of Acidic Drainage, 16-18 Sept.
1991, MEND, Montreal, Canada.

Wieder, R.K. 1988. Determining the capacity for metal retention in man-made wetlands
constructed for treatment of coal mine drainage. p. 375-381. In: Mine Drainage and Surface
Mine Reclamation, 19-21 April 1988, Info. Circular No. 9183. USDI, Bureau of Mines,
Pittsburgh, PA.

Wieder, R.K. 1989. A survey of constructed wetlands for acid coal mine drainage treatment in
the eastern United States. Wetlands 9:299-315.

Wieder, R.K. 1992. The Kentucky wetlands project: A field study to evaluate man-made
wetlands for acid coal mine drainage treatment. Final Report to the U.S. Office of Surface
Mining, Villanova University, Villanova, PA.

Wieder, R.K. 1993. Ion input/output budgets for wetlands constructed for acid coal mine
drainage treatment. Water, Air, and Soil Pollution 71: 231-270.

Wieder, R.K. 1994. Diel changes in Fe(III) / Fe(II) in effluent from constructed wetlands
treating acid mine drainage. J. Environ. Qual. 23:730-738.

Wieder, R.K. and G.E. Lang. 1982. Modification of acid mine drainage in a freshwater wetland.
p. 43-53. In: Symposium on Wetlands of the Unglaciated Appalachian Region, Oct. 1982,
West Virginia University, Morgantown, WV.

Wildeman, T., J. Gusek, and G. Brodie. 1991. Wetland design for mining operations. Handbook
presented at the 1991 American Society for Surface Mining and Reclamation Conference, 14-17
May 1991, Durango, CO.
41
Wiram, V.P. and H.E. Naumann. 1995. Alkaline additions to the backfill: A key
mining/reclamation component to acid mine drainage prevention. In: Proceedings, Sixteenth
Annual Surface Mine Drainage Task Force Symposium, 4-5 April 1995, West Virginia
University, Morgantown, WV.

Yanful, E.G., B.A. Abbe, M. Woyshner, and L.C. St-Arnaud. 1994. Field and laboratory
performance of engineered covers on the Waite Amulet tailings. p. 138-147. In: International
Land Reclamation and Mine Drainage Conference, 24-29 April 1994, USDI, Bureau of Mines
SP 06B-94, Pittsburgh, PA.

Ziemkiewicz, P.F. and F.A. Meek. 1994. Long term behavior of acid forming rock: results of
11-year field studies. p. 49-56. In: International Land Reclamation and Mine Drainage
Conference, 24-29 April 1994, USDI, Bureau of Mines SP 06B-94, Pittsburgh, PA.

Ziemkiewicz, P.F. and J.G. Skousen. 1992. Prevention of acid mine drainage by alkaline
addition. Green Lands 22(2): 42-51.

Ziemkiewicz, P.F., J.G. Skousen, D.L. Brant, P.L. Sterner, and R.J. Lovett. 1997. Acid mine
drainage treatment with armored limestone in open limestone channels. J. Environ. Qual.
26:718-726.

Ziemkiewicz, P.F., J. Skousen, and R. Lovett. 1994. Open limestone channels for treating acid
mine drainage: a new look at an old idea. Green Lands 24(4):36-41.

42

También podría gustarte