Anda di halaman 1dari 71

CHEMM901 MSci Advanced Chemical Project

ACID MINE DRAINAGE REMEDIATION USING

LOW-COST, INDIGENOUS MATERIALS

- ALMASU MARE CASE STUDY -

Student: Cosmina I. Dabu

Supervisors: D.W. Lewis (Chemistry), L.C. Campos (CEGE)


Abstract
The Almasu Mare region from the Southern Apuseni Mountains is one of the main ore
mining areas in Romania. Mining causes formation of Acid Mine Drainage (AMD)
which leads to the degradation of river and ground waters in the area.

This study looks at AMD remediation: the use of low cost materials for the purposes
of pollutant adsorption and pH neutralization. The goal of the study is to provide a low
cost and environmentally friendly technique for the removal of AMD.

Synthetic AMD (SAMD) was prepared (Fe, Zn, Mn sulfates) for study. Volcanic Ash
(VA) and two different types of Fly Ash (EBFA and TFA) were used to capture heavy
metal ions and neutralize the pH of the SAMD.

Various analytical techniques were used to investigate the efficiency of the ashes in
SAMD clean up. Morphological changes in the ashes as a result of the treatment were
also observed. It was concluded that all the ashes studied provided an efficient means
for water treatment.

Designs for in situ implementation included filter beds and cubes of FA and VA,
synthesized with a cement binder. It was found that a combination of both ashes
provided an improvement on previously developed designs for in situ remediation.
Abbreviations
AMD = acid mine drainage

DIW = deionized water

VA = volcanic ash

FA = fly ash

EBFA and TFA = two different types of biomass fly ash

AC = activated carbon

C = cement

OPC = Ordinary Portland Cement

L/S= liquid to solid ratio

TDS = total dissolved solids

ICP-OES = inductively coupled plasma optical emission spectrometry

SEM-EDX = scanning electron microscopy with energy-dispersive X-ray


spectroscopy

IR = infrared spectroscopy

ATR = attenuated total reflectance infrared spectroscopy

XRD = X-ray diffraction


Acknowledgements
I would like to thank my supervisors Dr. Dewi Lewis and Dr. Luiza Campos without
which this work would not have been possible, for the patience and trust and for
allowing me to explore countless new directions in this study. I would like to
acknowledge Dr. Julia Stegemann for providing the biomass fly ash samples, Dr. Anna
Bogush for specifying the idea to use the fly ash in AMD remediation and Dr. J.K. Kim
for specifying the idea of mixed ash cube synthesis and assisting with the process. I
would also like to thank Dr. Anna Bogush, Dr. J.K.Kim, Dr. Judith Zhou, Miss Catherine
Unsworth and Mr. Ian Sturtevant from the CEGE department, for supporting me
throughout the project. I also thank Mr. Martin Vickers for providing assistance with
the XRD pattern interpretation and Dr. Steve Firth for providing assistance with Raman
spectra interpretation.

I would like to extend my gratitude to Prof. Gyuri Ilinca, Dr. Denisa Jianu, and Dr.
Barbara Soare from the University of Bucharest (UB), Faculty of Geology and
Geophysics for kindly providing me with information and guidance about the project
background, and indigenous volcanic tuffs. I am also grateful to Ecoind and Minerals
Zeo for kindly providing us with free ash samples and analyses. I would also like to
acknowledge my colleague from UB, Mr. Dragos Mitrica, as well as my mother and
my sister for their invaluable help in the acid mine drainage sample collection from
Almasu Mare.

Last, but not least, I thank my colleagues Miss Belle Taylor, Miss Maria Calleja and
Dr. Matthew Li, for the kind help and endless patience.

Declaration of contributions
Dr. Anna Bogush and Miss Catherine Unsworth performed the ICP-OES analysis for
all the liquid samples (section 3.5). Dr. Bogush also performed the SEM-EDX analysis
(section 3.2.1). Dr. Steve Firth provided assistance in running Raman spectroscopy
(section 3.4). Miss Pragna Kiri performed the XRD analysis (section 3.1).
1 Table of Contents
1. Introduction .................................................................................................................... 1
1.1. Aims........................................................................................................................ 1
1.2. Acid mine drainage – the nature of the problem ...................................................... 1
1.3. General scheme of conventional AMD treatment .................................................... 3
1.3.1. Waste water remediation plants ....................................................................... 3
1.3.2. Slow sand filtration ........................................................................................... 4
1.4. Selection of low-cost remediation materials ............................................................ 4
2 Materials, and methods for preliminary batch experiments ............................................. 7
2.1. Materials selection for this project ........................................................................... 7
2.1. Preparation of synthetic AMD (SAMD) .................................................................... 9
2.2. Batch experiments of EBFA, TFA and VA on SAMD ............................................... 9
3 Batch experiments results ............................................................................................ 10
3.1. Powder X-ray diffraction (XRD) ............................................................................. 10
3.2. Scanning electron microscopy with energy dispersive X-ray analysis (SEM-EDX)10
3.2.1. SEM-EDX results ........................................................................................... 11
3.3. ATR-IR spectroscopy analysis .............................................................................. 17
3.4. Raman spectroscopy analysis............................................................................... 18
3.5. Inductively coupled plasma emission spectrometry (ICP-OES) ............................. 19
4 In situ remediation........................................................................................................ 27
4.1. Improving the efficiency of AMD treatment using a combination of ashes ............. 27
4.2. Small column experiments .................................................................................... 29
4.2.1. First small column design: thick layers of TFA and VA ................................... 29
4.2.2. Second small column set-up: VA and multilayers of TFA ............................... 30
4.2.3. Final small column set-up: mixture of VA and TFA ......................................... 30
4.2.4. Results of small column experiments ............................................................. 31
4.3. Synthesis of ash and cement cubes – packing of materials for in situ application . 31
5 Conclusion ................................................................................................................... 33
6 Further work................................................................................................................. 34
7 Appendices .................................................................................................................. 41
7.1. Appendix 1. Sample collection of natural AMD, Almasu Mare, Romania ............... 41
7.2. Appendix 2: Water quality data on the collected AMD, analyzed by Ecoind .......... 45
7.3. Appendix 3 Particle size distribution of ash ........................................................... 46
7.4. Appendix 4. TDS and pH results of selected experiments ..................................... 47
7.5. Appendix 5. XRD analysis..................................................................................... 52
7.6. Appendix 6. IR-ATR spectra ................................................................................. 57
7.7. Appendix 7. Raman spectra .................................................................................. 60
7.8. Appendix 8. Maximum adsorption capacity of ashes as a function of pH .............. 63
7.9. Appendix 9. Radii of hydration and hydration enthalpies .... …………………………64
7.10. Appendix 10. Comparison of our VA sample to other microporous sorbents from
literature. ......................................................................................................................... 65
1. Introduction

1.1. Aims

Acid mine drainage (AMD) is polluted water with a high heavy metal content and a
very low pH. It is presently recognized as the most important problem associated with
mining, worldwide. AMD is produced when waste products from the industry of ore
mining and processing are oxidized by atmospheric oxygen, in the presence of water.
High levels of heavy metals in the rivers cause serious health problems, such as
cancer [1]. This project, carried out under supervision by staff in the Chemistry and
the Civil, Environmental and Geomatic Engineering (CEGE) departments, takes on
the case study of Almasu Mare (Western Romania), a severely AMD affected area;
we propose its treatment using cheap, locally available materials. Firstly, we want to
test out selected materials on a synthetic AMD (SAMD) sample to determine their
potential and efficiency for heavy metal adsorption and pH correction. Secondly, we
want to investigate their AMD remediation mechanisms. Thirdly, we set out to
determine the most suitable way of packing these materials for an in situ application.
Finally, we hope to create a functional design for on-site AMD treatment.

To achieve these goals, we will use a batch setting for preliminary experiments and
filter beds for subsequent experiments. Various ways of packing the ashes will also
be explored. The solid samples will be characterized by a variety of techniques (XRD,
SEM-EDX, IR spectroscopy, Raman spectroscopy) to determine the morphological
and compositional changes as a result of AMD treatment. The liquid samples resulting
from the treatment will be analyzed by ICP-OES to determine the remaining heavy
metal content and possible metal leachates.

1.2. Acid mine drainage – the nature of the problem

AMD is formed when material containing sulfide (generally FeS2 containing rocks) is
exposed to oxygen and water. Although AMD is a natural process, mining exposes
large quantities of pyrite ore to the atmosphere, thus enhancing the rate of AMD
production. The decomposition of pyrite is accelerated by natural bacteria that oxidize
it yielding ferrous iron and sulfate. The resulting water that comes in contact with these
oxidized products is characterized by a low pH (-3.5 - 5) and a high concentration of

1
heavy metals and other toxic elements - several orders of magnitude higher than the
maximum admitted limits for drinking water. Primary AMD sources include mine rock
and tailing dumps, mine workings (underground or open pit), or construction rocks
which are used in dams or roads. Secondary sources include treatment sludge
impoundments or emergency ponds [2].

Assuming pyrite as the oxidized mineral and oxygen as the oxidant, AMD generation
was described by [2] in a simplified manner as the reaction:

FeS2 +15/4 O2+7/2 H2OFe(OH)3(s)+2H2SO4

Natural AMD constitutes a complex system, with many variables, so the SAMD we
created had to be greatly simplified. Some of the precipitates, organics and
microorganisms present in the environment provide naturally reduce the existing AMD.
For example, various metal oxides such as Fe and Mn oxides are microporous media
which uptake heavy metal ions [3] pH buffering reactions occur as a result of the
dissolution of carbonate rich minerals, Al(OH)3 and aluminosilicates [4]. Furthermore,
simply aerating the water through a cascade-like setting promotes the oxidation of Fe
and Mn to insoluble hydroxides [5]. The final design of an in situ treatment should
make note of these factors and maximize their contribution.

AMD is a long-standing and widespread problem and as such a number of treatment


solutions already exist. Unfortunately, there are still a number of limitations associated
with these methods. These include sludge production and high costs of chemicals.
Economic feasibility becomes especially crucial for developing countries, where AMD
is extensive and the need for treatment is urgent.

The Apuseni mountains region of Romania, particularly the Southern part (Metaliferi
Mountains) where Almasu Mare is located, is known to be one of the oldest areas of
ore exploitation in Europe [6-8]. It is part of the Golden Quadrilateral and contains the
largest area of gold deposits in Europe, along with important quantities of Cu, Pb, Zn
and other minerals. Hundreds of years of exploitation have brought about very serious
AMD issues. In June 2013, the author collected samples from three points: the
entrances of two closed mines and a nearby river at the entrance of a small inhabited
village (details in 7.1 Appendix 1). We were told that people knew the surface water
was contaminated, and that the contamination had reached underground water, hence

2
they were collecting drinking water from a source several kilometers away. However,
the water was still used for livestock and crops irrigation. Ecoind1 in Romania did
preliminary tests on these samples in order for us to establish which pollutants were
most significant, and thus how best to model the SAMD. Tests showed the presence
of twelve different p and d-block metals (7.2 Appendix 2). Some of the metals were
found at concentrations 100 times the EU maximum admitted limit for river waters and
the average pH of the water samples was 2.7. At present, all the mines in the area are
non-operational yet none of them are safely closed down (to prevent AMD leakage)
nor is there any method of AMD remediation in place. The AMD leaks from the mines,
into creeks, and small rivers that go through villages, and eventually into large river
systems.

In order to develop a feasible in-situ treatment method, the current state of art
methodology is briefly reviewed.

1.3. General scheme of conventional AMD treatment

1.3.1. Waste water remediation plants

The recommended treatment processes for ore mining and dressing is lime
precipitation and settling, followed by impoundment and recycling. A general scheme
(provided in Figure 1-1) starts with passing water through grits to remove large solids
[9]. Chemical treatment is then applied for pH neutralization, followed by the addition
of a coagulant or a flocculant. Rapid mixing and flocculation are then employed to
group the resulting precipitates into flocs, which eventually form sludge. The sludge is
then pumped into a sedimentation tank, where it settles down at its bottom. Most of
the sludge is removed from there, and the remaining sludge is removed through
filtration. Additional processes such as ion-exchange or adsorption on activated
carbon (AC) are also quite common. The most common chemicals used as coagulants
are iron salts (FeCl 2, FeCl3, sulphates), aluminium salts (Al 2 (SO4)3,) polyaluminium

1
The National Institute for Research and Development for Industrial Ecology (Ecoind) is one of the
forefront Romanian institutions in the field of environmental research and services

3
chloride and NaAlO2). For pH correction lime, Na2CO3, NaOH and sulfide salts are
used and their quantity depends on the pH at which targeted metals precipitate [10].

Figure 1-1: General scheme of heavy metal contaminated waste water treatment [11]

1.3.2. Slow sand filtration

An interesting, low cost and effective way for waste water clean-up is using biosand
filters, which is a type of slow sand filtration. These have been used in the context of
heavy metal removal and other types of wastewaters. They reduce turbidity and colour
of the wastewater and remove chemical contaminants and microorganisms through a
single filtration process [12]. The downside is that this process cannot clean up an
entire river; a typical flow rate is limited to 15-20 L hr-1 but can be slower depending
on the height of the water column on the filter top and the size and type of the filter
components. Treatment usually involves diverting a part of the river to the filter, or
manually pouring water on the filter top, which is why this treatment is more suitable
for individual households.

1.4. Selection of low-cost remediation materials

Various low-cost materials have been used for AMD remediation. Limestone or lime
are used for chemical pH neutralization. Natural or man-made wetlands are used as
a set of complex processes occur; bioaccumulation, bacterial and abiotic oxidation,
sedimentation, neutralization and dissolution, as well as adsorption and ion-exchange
[13]. Carbon for metal adsorption is derived from a variety of waste products, such as
risk husks, coal, wood, coconut, peanut or almond shell, olive stones and so on.
Adsorption on waste or low-cost materials is one of the most promising ways for AMD

4
treatment. Some notable examples are chitosan (from fishery wastes), clay (has a
similar sorption capacity to zeolites), and industrial waste such as slurry [14, 15].

Biomass ash is a type of fly ash (FA), which contains additional biological starting
material. It mostly consists of the mineral constituents of calcinated litter, in oxide or
salt form. It is one of the cheapest materials for heavy metal removal, due to its high
availability and due to the fact that it doesn’t require any special pre-treatment or
reaction conditions.

One of the most utilized types of ash is ash derived from the combustion of coal.
Combustion processes usually produce a number of types of residues – FA, bottom
ash and boiler slag. Bottom ash and boiler slag are coarse (19-75 mm) and settle
down in the combustion chamber, whereas FA is composed of micron-sized particles
which are collected from dust collection systems [16]. It can be described as a ferro-
alumino-silicate with a significant amount of Ca, K and Na; the major compounds found
in FA are: SiO2, TiO2, Al2O3, Fe2O3, MnO, MgO, CaO, Na2O, K2O,P2O3, SO3.

This project is interested in calcareous (> 10%CaO) FA. CaO is what gives FA a high
alkalinity (pH = 9 -12). FA has a glassy matrix, consisting of 66-88% glass and a 2-3%
water-soluble fraction, which is primarily made of Ca 2+. Depending on the original
waste source that was calcinated, harmful trace metals such as Pb, Cd, Mn etc. can
be present in varying concentrations. FA has been studied as an adsorbent for heavy
metals under various conditions: contaminant type, pH, temperature and contact time
[17-19]. The adsorption of Cr(IV) by a mixture of FA and woolastonite (natural mineral
CaSiO3) was reported by Panday et al [20] , who found that 2.92 mg/g Cr6+ was
absorbed at pH = 2, assuming a Langmuir-type model. Vadapali et al [21] showed that
AMD can be neutralized by an alkaline South African FA using an L/S=3. They found
that maximum heavy metal removal occurred at L/S = 1.5 where 99.5% Fe and Mn
were removed. Zn, Cu, Ni and Pb were also successfully removed at various L/S
ratios, and Mg was removed entirely at pH > 11. The storage problem of FA has been
addressed through the proposition of using residual FA as raw material for
synthesizing zeolites, or as a backfilling material for mines [22].

Zeolites were one of the first materials used to treat AMD taking advantage of their
many useful properties, such as their ability to exchange some harmful compounds
into their structure. Numerous studies concerning heavy metal adsorption by natural
5
zeolites, and in particular by clinoptilolite, have been carried out for a variety of types
and concentrations of heavy metals [23] [24, 25] [26] [27]. Various reactivity series (in
exchange equilibrium) have been produced for clinoptilolite from isotherm data.
Medium levels of heavy metal contamination at pH ~ 6 produced the following series:
Co2+ > Cu2+ > Zn2+ > Mn2+ [24]. Gaikwad et al [28] used stilbite and two synthetic strong
acid ion exchange resins to test their efficiency in removing Cu(II) from synthetic AMD,
prepared with 200 mgL-1 CuSO4. Batch experiments showed the maximum adsorption
capacity of Cu(II) to occur in stilbite. 100% removal of Cu(II) was observed after 24
hrs contact time with zeolites, and around 80% in the case of both resins. Maximal
heavy metal adsorption in the natural zeolite clinoptilolite was found to occur at pH ~
6 [29].

Moreover, zeolites are generally very stable and can be left in the environment after
heavy metal uptake, or can be regenerated (usually with a solution of concentrated
brine). Zeolites have also been used as an element in biosand filters (BSFs), an
emerging technology developed using low-cost, locally available materials for the
removal of surface water pollutants. Mahlangu et al [30] report the development of a
bucket filter. It was comprised of a top biological layer (made of slime, micro-organisms
and sediments), fine sand, zeolites, coarse sand and gravel. This was found to provide
efficient removal of Ca, Mg, As and Fe ions.

Ordinary Portland Cement (OPC) is a mixture of di- and tricalcium silicate, tricalcium
aluminate, tetracalcium aluminoferrite and gypsum [31] which makes it a highly
alkaline material, suitable for AMD pH correction. There are various safety issues
associated with it, such as its exothermic settling process, which consequently makes
wet cement very caustic. Like FA, due to its high alkalinity, dry OPC is a serious irritant.
Release of CO2 during manufacturing as well as the consumption of a lot of fuel to
produce a temperature of about 1450°C required for calcination, contribute to its low
popularity as an AMD treatment material. OPC has been introduced as a secondary
material in AMD treatment; for example Gitari et al used it as a binder in a 6% OPC
mix with solid residues of fly ash obtained after AMD treatment [32]. This sludge was
subsequently used as a backfilling material for the mine. Evaluation over a long period
of time showed that OPC reduces the alkalinity of the FA residues, yet AMD treatment
remains effective.

6
2 Materials, and methods for preliminary batch experiments

This section discusses the materials selected for this project and why they were
chosen. We also outline the preparation of synthetic acid mine drainage (SAMD) and
preliminary batch experiments. The results from these experiments are presented in
the next section. Experiments regarding in situ implementation are each presented
separately in section 4.

2.1. Materials selection for this project

For this particular project the following materials were used:

Ash type Description


EBFA A type of biomass fly ash, derived from the co-combustion of poultry litter with
wood, horse bedding etc.

TFA A type of biomass fly ash, derived from straw combustion

VA Volcanic Ash rich in natural zeolite, clinoptilolite


*N.B. Particle size distribution details in 7.3 Appendix 3

Biomass ash represents about 19% of Romania’s primary energetic potential. Much
of this biomass ash comes from wood wastes, agricultural wastes from cereals, corn
or vine residue and urban and household wastes [33]. Thus biomass ash is a cheap,
readily available resource Romania. In this study, the biomass ash was used for two
purposes: pH remediation and heavy metal precipitation.

Romania has important natural volcanic tuff resources. Zeolite rich tuff (> 50%
clinoptilolite) can be found in a number of areas of the country [34]. Tuff from some of
these regions are already exploited; at present, the prices for raw crushed volcanic
tuff are low. Sample tuff was donated by Minerals Zeo contains about 60-70% zeolitic
crystalline phases [35]. Zeolites are able to remove heavy metal ions through two
mechanisms: adsorption and ion-exchange. Clinoptilolite is a particularly good
candidate for ion-exchange in an AMD context as it has large pores, a chemically
neutral structure and is unaltered in pH > 2. Prior to any experimental work, a plan
(see Figure 2-1) was devised to test a variety of scenarios and systematically work
from characterizing the materials to in situ implementation.

7
Context: Severely AMD affected area Almasu Mare, Romania

Why?
Collection of water samples from drainage sources in Almasu Mare

Preliminary report of metal contamination done in Romania -


showed what the most important heavy metals were

1.Are the ashes Determination of potential low-cost natural, indigenous materials


candidates for in situ
treatment? that can be used in AMD treatment. VA and FA were selected.
- no harmful leachates
or compounds formed
2. Is the treatment
efficient? Batch experiments conducted on synthetic AMD (SAMD), using VA
- is the pH neutralized, and two types of FA => assess the behaviour of materials and
are most of the heavy
metals removed, and determine efficiency of AMD treatment
is water potable?
3. What is the metal
removal mechanism Characterization of solid samples
and how is pH Characterization of liquid
before and after treatment:
neutralized? samples after treatment: ICP-
XRD, SEM-EDX, IR and Raman
OES
spectroscopy

What is the most In situ application


efficient ash
combination?
What is the best Batch experiments with Mixed filter Packing of FA and
way to pack the
ashes? two adsorbents => FA beds: FA, VA in cubes with
What is a feasible pre-treatment and VA VA, sand cement binder
treatment design?

Optimization of Creating new Design of in situ


What
parameters from ways of application technology
next?
existing materials and methodology
experiments packing
Fig. 2-1 Project outline

8
2.1. Preparation of synthetic AMD (SAMD)

SAMD was prepared using approximately 1.50g ZnSO 4, 0.75 g MnSO4, 6.50g FeSO4
which were dissolved in 5L deionized water. 1mL 95% H2SO4. H2SO4 was used to set
the solution to pH = 2.7.The total dissolved solids (TDS) were 652 mg L-1 2 .

2.2. Batch experiments of EBFA, TFA and VA on SAMD

Batch experiments were used as preliminary testing of the ashes for AMD treatment.
We wanted to gather information about the composition and morphology of the ashes
before treatment and then cross-reference this information with results after
experiment. It is important to know if these materials are candidates for in situ AMD
treatment, which implies they do not leach out any harmful elements when exposed to
acidic water nor do they form any harmful compounds as a result of the treatment. We
also wanted to detect the efficiency of the AMD treatment, hence test the pH of the
treated AMD and the equilibrium concentration of heavy metals and see if the water
quality reached drinking standards. Additionally we wanted to gain insight in the type
of heavy metal removal mechanism these ashes employ.

Batch experiments were conducted by placing a known amount of ash in 25 mL of


SAMD. The solution was placed in a sealed container and continuously mixed for 24
hrs in an automatic stirrer. For EBFA and TFA, 0.010, 0.035, 0.050, 0.075, 0.100 and
0.250 g were used respectively. For VA, 0.5, 1.0, 1.5, 2.0, 2.5 and 3.0 g were used.
pH, TDS (details and results of these measurements in 7.4 Appendix 4) and ICP-OES
were performed on the liquid samples after filtration through a 45 micron filter paper.

Solid samples were collected and characterized using XRD, SEM-EDX, IR and Raman
spectroscopy.

A small number of additional samples of EBFA and TFA were left to treat AMD for 1
hr. These results are only discussed in the ICP-OES results section, where they are
compared with the corresponding 24 hr results (extra details in 7.4 Appendix 4).

2
The World Health Organization puts drinking water quality at fair for 600-900 mgL-1 [52]

9
3 Batch experiments results

3.1. Powder X-ray diffraction (XRD)

All three ash samples were characterized before and after AMD treatment by XRD
(Bruker D8 Discover, Cu Kα radiation). The samples were step-scanned, integrated at
0.05° 2θ, over the range of 2-80°. XRD measurements were obtained using a thin film
of powder sample at room temperature.

EVA software was used, employing a search/match facility in order to determine the
most likely phases of minerals in the ash.

The EBFA sample before experiment showed crystalline portlandite (7.5 Appendix
5(A)) and an unidentified major crystalline phase, whilst the sample after experiment
didn’t show any crystalline phases (7.5 Appendix 5 (B)). TFA before experiment ( 7.5
Appendix 5 (C)) showed crystalline quartz and several other unidentified phases. TFA
after experiment (7.5 Appendix 5 (D)) shows a low crystalline quartz phase and several
unidentified phases. XRD data of VA before and after experiment showed the
presence of clinoptilolite (7.5 Appendix 5 (E)).

3.2. Scanning electron microscopy with energy dispersive X-ray analysis (SEM-
EDX)

SEM-EDX is a non-invasive technique that uses an electron beam to stimulate atoms


in the sample to give out characteristic X-rays, thus providing chemical
characterization of heterogeneous inorganic or organic samples. This project uses
SEM to determine the surface morphology of FA and VA samples before and after
experiments. EDX mapping is used to determine what elements are present on their
surfaces.

A Jeol JSM-6480LV high-performance, variable pressure analytical scanning electron


microscope equipped with and energy dispersive system and electron backscatter
diffraction using the Oxford Link system was used. Samples were loaded on Cu stubs
coated with a carbon graphite glue mixture and then carbon coated for 20 minutes. A
number of sample magnifications were used in order to best see the surface
morphology for each of the selected ash samples.

10
3.2.1. SEM-EDX results

This section will present the FA and VA surface morphology characterization, as well
as the EDX mapping of existing surface elements, before and after AMD treatment.
Samples before treatment, or the control samples, were left in distilled water for 24
hrs, and samples after treatment resulted from the above mentioned batch
experiments. The ash control sample is introduced first, followed by the corresponding
after treatment sample. We then compare the morphology and elements of these
samples to see if significant changes appear as a result of AMD treatment. Potential
existing compounds are also discussed. These findings can indicate the AMD
treatment mechanism; if AMD specific elements are found on the surface of the
samples, the mechanism is adsorption. If specific AMD elements are not found, but
ICP-OES analysis confirms their removal from the liquid sample, the mechanism is
sorption into the structure of the ash, or ion-exchange.

TFA characterization

A representative image of TFA control sample is presented (Figure 3-1) which shows
typical fly ash spheres at point 1, composed of amorphous silicates [36]. EDX mapping
of these spheres shows they are primarily Ca rich but contain a variable quantity of Fe
as well. Other notable elements found in the elemental spectrum of point 1 are P and
K, with traces of Al and Mg. Surface morphology didn’t show any crystalline phases,
so they probably exist as a mixture of amorphous compounds such as orthoclase,
apatite, olivine and so on. Point 2 shows a region with a high content of Fe (elemental
analysis not shown here), and point 3 is quartz, based on its morphology [36].

11
Figure 3-1: SEM image of TFA control sample. Elemental spectrum of point 1 (on FA typical amorphous
silicate sphere) is shown. Point 2 represents an iron rich particle. Point 3 is quartz.

SEM of the TFA after treatment sample indicated that the surface morphology
changed visibly compared to the sample before treatment (Figure 3-2, (A)). A number
of different formations appeared over the characteristic fly ash particles. The
characteristic FA silicate spheres still exist (Figure 3-2,(B)). The EDX mapping of point
1 in Figure 3-2 (C) as well as the octahedral, cauliflower-like morphology indicates the
presence of crystalline pyrite or jarosite based on and metal (oxy)hydroxides [37]. XRD
contains some unidentified phases (7.5 Appendix 5(D)), so they could correspond to
jarosite. Zn and S appear exclusively as a result of AMD treatment. Morphology of
formations at Point 2 in Figure 3-2(C) shows possible amorphous Mn, Fe, Zn
(oxy)hydroxides. Point 3 in the same figure shows a quartz particle.

In summary, there are definite morphological and elemental changes to the TFA after
AMD treatment. Ca rich regions appear in the control samples due to the mixing of
TFA with water, thus dissolving CaO. Samples after experiment are enriched with
AMD specific elements. The Ca, Fe, Mn, Zn are found as amorphous oxides,
oxyhydroxides and hydroxysulphates [32]. Minerals such as jarosite, wollastonite,
apatite and olivine are also a possibility. Based on this analysis, the mechanism of
TFA heavy metal removal is adsorption onto FA particle surfaces.

12
Figure 3-2 SEM images of TFA after AMD treatment. (A) shows morphological changes of the
sample as a results of AMD treatment. (B) Close-up of typical FA particle, elemental spectrum
of point 1 is shown. (C) Close-up of excrescence region. Point 2 is a region with amorphous
oxy(hydroxide) species. Point 1, with elemental spectrum included, shows crystalline pyrite or
jarosite based on and metal (oxy)hydroxides. Point 3 shows quartz.

EBFA characterization

EBFA control samples exhibit smaller particles than the corresponding TFA ones
(Figure 3-3 (A)). The morphology is different as well: the spherical formations are
present, but are less abundant and not well defined. However, the elemental
composition is very similar to TFA in that K, Ca, Mg, P, Si, Al, O are seen

13
(Figure 3-3 (A)). Areas of the sample not shown here have small amounts of Fe and
Mn. This EBFA also contains traces of S and Cl, but no harmful heavy metals
(confirmed by ICP-OES on broad range of metals).

Figure 3-3: SEM images of EBFA control samples. (A) Surface morphology of sample, (B) Close-up of
surface and elemental spectrum of point 1.

EBFA after AMD treatment does not have spherical particles, nor does it show any
likely crystalline phases (Figure 3-4 (A)). As with TFA, EDX mapping of the sample at
point 1 in Figure 3-4(B) shows an average increase in the abundance of AMD specific
elements Fe, Mn, Zn and S. This indicates the removal mechanism is adsorption.
Judging by the morphology, these are probably present in mixed amorphous phases,
predominantly metal (oxy) hydroxides [32].

14
Figure 3-4: (A) SEM image of EBFA after AMD treatment, (B) Close-up of surface and elemental spectrum
of point 1.

VA characterization

Figure 3-5(A) shows the structure of VA control sample. Figure 3-5 (B) point 1 shows
the typical tabular monoclinic crystal structure of clinoptilolite [35]. All the elements
present are constituents of this mineral. The atomic percent of the elemental spectrum
shows 26.80% Si and 5.51% Al, which works out as a 4.86 Si/Al ratio. This compares
very well to its theoretical value of Si/Al = 5 (derived from the pure clinoptilolite formula
[(Na, K, Ca)6Al6Si30O72. 24H2O]). Point 2 shows Mn rich particles. Other amorphous
phases of the same type of elemental composition were identified, as well as portions
of silica. These are likely to be amorphous as they do not show up in the XRD data
(7.5 Appendix 5 (E)).

The main morphological features are maintained in the VA ash sample after AMD
treatment (Figure 3-6). The appearance of amorphous formations over the clinoptilolite
crystals is also indicated. The bright spots in Figure 3-6(A) show Fe rich formations,
and so does point 1 in Figure 3-6 (B).

15
Figure 3-5: (A) SEM image of VA control sample, (B). Close-up of surface and elemental spectrum of point
1, showing clinoptilolite. Point 2 shows Mn rich formations.

Fe in Figure 3-6 appears exclusively as a result of AMD exposure and so does Zn.
However, the average weight percentage of Fe and Zn over all the mapped points in
VA is much smaller than the average weight percent over all the EBFA or TFA
samples. Additionally, S does not appear at all in VA after treatment. These findings
suggest that the main mechanism of heavy metal removal by volcanic ash is ion-
exchange within the zeolite, although complementary precipitation occurs, as shown
by the Fe-rich particles.

In summary we concluded that TFA, EBFA and VA are all good candidates for AMD
treatment. No harmful elements leach out in the presence of AMD, nor could we
determine the presence of any harmful compounds formed as a result of the treatment.
For TFA and EBFA, CaO dissolves to neutralize the pH. AMD specific elements are
exclusively present in the samples after treatment. The heavy metal removal
mechanism is adsorption; various amorphous and crystalline phases form on the
surface of the samples. No harmful trace metals were detected in VA either. Samples
before and after treatment showed the presence of clinoptilolite. VA’s main

16
mechanism of heavy metal removal is ion-exchange. The lack of S-containing species
points to the likelihood that SO42- was adsorbed along with Fe, Zn and Mn. A study
done by Oliveira et al [38] supports this conclusion; a natural zeolite pretreated with
Ba2+ showed a significant uptake of sulfate ions.

Figure 3-6 (A) Surface morphology of VA after AMD treatment. (B) Close-up of surface and elemental
spectrum of point 1.

3.3. ATR-IR spectroscopy analysis

Infrared spectroscopy measures the intensity of the adsorption and wavelength of


light in the mid-infrared region by a sample. ATR allows for a superior method of
powder sample collection, especially for samples that have low concentrations of
particular compounds, as in this case where silica and carbon are predominant and
dominate the spectrum. A Miniature FTIR Spectrometer (Bruker Optics Alpha) with an

17
integrated single reflection diamond tip was used. OPUS software was used to plot
the spectra.

Two VA samples after AMD treatment were plotted against a control sample (7.6
Appendix 6 (A)). No changes were detected with this technique. The broad bands at
3600 - 3400 cm-1 and the peak at 1630 cm-1 correspond to water inside the zeolite.
The peak at 1100 cm-1 corresponds to Si-O-Si and Si-O-(Al) asymmetric stretching
vibrations [39].

Two samples of TFA after experiment were plotted against a control sample ((7.6
Appendix 6 (B)). The bands at 1450 cm-1 and 850 cm-1 are due to M-OH vibration and
so is the one at 850 cm-1. The peak at 3660 cm-1 corresponds to an OH stretch, or a
water stretch from quartz. The broad peak at 3400 cm-1 is due to the presence of water
[39]. The split peak at 1100 cm-1 is attributed to O-Si-O stretching vibrations but it could
point out to the association of O with other compounds. No changes in the functional
groups as a result of AMD could be detected.

Three samples of EBFA were plotted against a control one (7.6 Appendix 6 (C)). The
IR peaks look very similar to those of TFA, in all the samples. No change can be
attributed to AMD treatment.

3.4. Raman spectroscopy analysis

Raman scattering is a type of molecular spectroscopy which gives information on the


structure and properties of molecules from their vibrational transitions. Raman spectra
can include fluorescence spectra as well.

The fine black particles of EBFA were identified as carbon. TFA had different coloured
particles, the black ones gave rise to a typical carbon spectrum (Figure 3-7). The G
band (or tangential band), found at 1582 cm-1 is characteristic of graphite [40] [41].
The D band occurs at about 1350 cm-1 and points to a disorder in the graphene
structure; it is a defect band that comes from the hybridized vibrational modes of
graphene edge. A broad band around 500 cm-1 shows the existence of amorphous
carbon.

18
2000
Raman Intensity (A.U.)

1000

0
0 1000 2000 3000
Raman shift (cm-1)

Figure 3-7: Representative spectrum of carbon in TFA after experiment – black particle, laser wavelength
785 nm, 5% power. N.B. The graph presents simple data points; the data is very noisy and as such appears
as a thick line.

The Raman spectrum of the TFA white particle showed the presence of crystalline
quartz at the characteristic peaks of 265 and 463 cm-1 (7.7 Appendix 7 (A)), which is
confirmed by the XRD. The presence of CaCO3 (calcite) is also suggested through the
presence of a sharp peak at 1086 cm-1 and several others in the region 150 -
300 cm-1 [40]. The VA sample before experiment (7.7 Appendix 7 (B)) shows a
spectrum characteristic of Clinoptilolite-Na [42]. The spectrum of the VA dark brown
particle after experiment (7.7 Appendix 7 (C)) has a low S/N ratio; hence it is hard to
definitely characterize the peaks. However, the peaks around 215 and 320 cm-1 and
the ones at 480 and 640 cm-1 can indicate the presence of Fe-oxides and Fe-sulfates
[40].

3.5. Inductively coupled plasma emission spectrometry (ICP-OES)

ICP-OES is an analytical method based on optical emission spectrometry. It is


primarily used for the detection of trace metals [43]. ICP-OES was used to determine
the concentration of Fe2+, Zn2+ and Mn2+ that were originally present in the AMD
sample and the remaining amount after the samples were exposed to the ashes. It
was also used to determine the quantity of Ca 2+ and Mg2+ that leached out of the
ashes. Analysis was done on a Varian 730 ICP-AES (axial configuration) equipped
19
with an autosampler, at the UCL Cross-Faculty Elemental Analysis Facility located at
the UCL Earth Science department. A 5% error is expected for these measurements,
which was adopted for the results presented in the subsequent graphs.

This analysis coupled with pH measurements provides us with data that assess the
efficiency of the ashes for AMD remediation. We are interested to see the kind of
adsorption isotherms the ashes obeyed. Adsorption isotherms describe the
relationship between the equilibrium concentration (Ce) of the adsorbed heavy metal
and the quantity of heavy metal adsorbed per unit of adsorptive material (or maximum
adsorption capacity qe). The particular shapes of the isotherms indicate some of the
material’s characteristics and can indicate the heavy metal removal mechanisms. The
resulting isotherms of the ashes were compared to model isotherms (Figure 3-8). We
are expecting type I isotherm, which is typical for microporous solids, or type II, which
indicates fine, non-porous solids [44].

Figure 3-8: The six standard IUPAC isotherms [44]

We are also interested in the percent removal of heavy metals as a function of ash
mass and the maximum adsorption capacity of the ashes as a function of solution pH.
All of these results are presented below.

EBFA AMD remediation

EBFA shows typical type II adsorption isotherms (Figure 3-9). Most of the heavy metal
uptake from the solution is shown to happen within the first hour of reaction. The pH
results don’t show any change with time either; an optimal pH = 7 is obtained for
0.100 g EBFA (7.4 Appendix 7 (A)).

20
180

160

140

120
qe (mg/g)

100

80

60

40 Fe
Mn
20 Zn

0
0 50 100 150 200 250
Ce (mg/L)

Figure 3-9: EBFA heavy metal absorption isotherms

ICP-OES results obtained after 24 hrs treatment showed 100% removal of Fe2+ and
Zn2+ with 0.250 g EBFA but only about 65% of Mn2+ (Figure 3-10). Mn generally
seemed to have much lower removal efficiency than Fe or Zn.

EBFA also leads to a significant increase in the equilibrium concentration of Ca with


increase in ash mass (7.4 Appendix 7 (A)). This indicates that the pH of the solution
AMD is mainly increased through the dissolution of the soluble alkaline fractions,
primarily Ca oxides. EDX confirms this, as does IR spectroscopy. Dissolution of
various hydroxide materials such as Fe or Al hydroxides may have aided the pH
increase. These act as buffers in the pH ranges 2.5-3.5 and 4.0-4.5 respectively.

In an in situ situation, we may want to use 0.100 g EBFA (liquid to solid ratio,
L/S = 250), since it shows the steepest increase in removal efficiency. It removed
almost all of the Fe and Mn and 60% Zn. Maximum adsorption capacity of this material
occurs at pH ~ 4 for all the heavy metals (7.8 Appendix 8).

21
100

80
Percent removal (%)

60

40

Fe
20
Mn
Zn
0
0.0 0.1 0.2 0.3
Mass (g)

Figure 3-10: Percent removal of heavy metals (y-axis) as a function of EBFA mass (x-axis)

We may thus conclude that EBFA removes the studied heavy metals from AMD
efficiently in a short amount of time, t ≤ 1 hr at L/S = 250, and fully neutralizes the pH.
For in situ remediation, it is worth noting that the formation of Fe oxides at a high pH
facilitated the subsequent precipitation of Mn and Zn species. Fe oxides are
microporous sorbents with a large area surface area of adsorption and a high affinity
for cation adsorption.

TFA AMD remediation

Similar to EBFA, TFA shows a type II adsorption isotherm (Figure 3-11). 0.250 g of
TFA (L/S = 100) shows 100% removal of all the heavy metals, after 24 hrs of
experiment (Figure 3-12). The corresponding pH of the sample was 9.5. TFA’s Mn
removal is more time dependent than EBFA but performs very close to the results
below even after 1hr (7.4 Appendix 7 (B)). Hence, AMD treatment using this ash
should be performed with L/S = 100, at a reaction time, t ≤ 1 hr. Maximum adsorption
capacity occurs around pH ~ 4 .5 for all the metals (7.8 Appendix 8).

22
300

250

200
qe (mg/g)

150

100

Fe
50 Mn
Zn
0
0 50 100 150 200
Ce (mg/L)

Figure 3-11: TFA heavy metal adsorption isotherms

100

80
Percent removal (%)

60

40

20
Fe
Mn
Zn
0
0.0 0.1 0.2 0.3
Mass (g)

Figure 3-12: Percent removal of heavy metals (y-axis) as a function of TFA mass (x-axis)

For both the FA types, an extra step should be inserted in the in situ remediation
process to remove some of the dissolved solids as well as the rest of the heavy metals,
while keeping the pH constant. Zeolites (from VA) would be a very good solution, since
they act as a pH buffer and have a high affinity for ion uptake (heavy metals and Ca),

23
hence decreasing the total dissolved solids. They also act as a flexible layer to trap
the resulting sludge.

VA AMD remediation

In terms of the chemistry involved, the zeolite is an oxidizing contact surface which
undergoes ion-exchange with the heavy metal ions (and in situ acts like a physical
filtration medium as well) [45]. In the case of VA, the main mechanism of pH
remediation is likely the dissolution of aluminosilicates, whereby H+ is consumed and
H4SiO4 is produced, and ion-exchange of H+ into the clinoptilolite [4]. This ion has a
very small radius of hydration (rh), as well as a relatively low standard molar hydration
enthalpy (7.9 Appendix 9) and is present in a high concentration at the starting AMD
pH, hence its diffusion is favoured. Zeolites exhibit an amphoteric behavior, hence can
act as solid buffers in aqueous media.

VA obeyed the type I adsorption isotherm (as it produced a straight line in a log-log
plot of the Freundlich equation), which is specific to microporous solids. The main
removal mechanism was ion-exchange, which implies sorption of metal ions within the
microporous solid and leaching of Ca, Na and K ions from within the framework. The
Freundlich isotherm was used for calculations. This model assumes that the uptake of
molecules occurs on a heterogeneous surface, where the active sites have a
heterogeneous energetic distribution [46]. This particular type of isotherm was chosen
because a number of studies in literature showed the behaviour of clinoptilolite to obey
this model well [47-49].
1/𝑛
Freundlich equation: 𝑞𝑒 = 𝐾𝐶𝑒

Where qe = amount of heavy metal ions adsorbed on adsorbent (mg/g)


K = Freundlich equilibrium constant; it is a sorption capacity measure
Ce = equilibrium concentration of heavy metal ions
n = a constant related to sorption intensity

Taking the log of both sides produces a linear plot of log (qe) vs log (Ce ),where log(K)
is the intercept and 1/n is the slope (Figure 3-13).
𝑙𝑜𝑔(𝑞𝑒 ) = log(𝐾 ) + 1⁄𝑛 𝑙𝑜𝑔(𝐶𝑒 )

24
These can be directly compared with other materials to produce a relative adsorption
capacity measurement of the VA.

Figure 3-13: Freundlich isotherm log-log plot

The correlation coefficients R2 are very close to the ideal value of 1 (Table 1:
Freundlich constants in the VA - heavy metal ions system. The slopes are 1/n << 1,
which indicates that the clinoptilolite adsorptive capacity is only reduced by a small
amount at lower equilibrium concentrations. Values of n between 2 and 10 are
indicators of good adsorption, hence all the metal ions were adsorbed well onto the
clinoptilolite.
Table 1: Freundlich constants in the VA - heavy metal ions system

Metal K (mg/g) n Correlation


coefficient (R2)

Fe 2+ 3.174 4.831 0.9891

Mn 2+ 0.354 2.616 0.9957

Zn 2+ 0.837 2.837 0.9965

The difference in adsorption capacity, given by K, is related to a number of factors,


such as the original concentration of ions in solution, cation radii of hydration, and

25
hydration enthalpies [23]. Our VA performed well compared to other microporous
sorbents (7.10 Appendix 10).

Based on the factors listed above, we can explain the trends in heavy metal adsorption
by our clinoptilolite. The heavy metal concentration present in the solution acts as a
driving force for the ion exchange of s-group metals from within the clinoptilolite.
Additionally, the higher polarizability of the heavy metals compared to the original Na +,
Ca2+ and Mg2+ causes a stronger binding to the zeolite framework. Ion exchange is
favoured for ions with increasing charge and from those with the same valency, the
higher atomic number ones. The smaller particles are adsorbed faster and in higher
quantities; they can diffuse through the internal structure of the zeolite easier. For the
heavy metals used, the hydration radii series is rh Fe2+ < rh Zn2+ < rh Mn2+. This is
consistent with the Freundlich isotherm related findings. The charge density on the
metal ion makes a difference in terms of how well it can interact with the adsorption
sites; higher charge density usually means a larger hydration sphere and hence a
higher hydration energy [50]. Hydration enthalpies (ΔhydH°) (7.9 Appendix 9) show the
series: Zn2+< Fe2+ < Mn2+.The hydration energy difference between the Fe and the Zn
ions is quite small though, so the other effects (heavy metal concentration and radii of
hydration) were probably preeminent. The maximum adsorption efficiency occurred at
pH ~ 5.4 for all three metals (7.8 Appendix 8).

A very good removal efficiency (> 85% of all the heavy metals) occurs for 1.5 g VA
(L/S = 17) (Figure 3-14). The resulting solution pH was 5.7. However, ion-exchange is
a process which requires a long contact time, and so for in situ treatment we would
need to slow down the water flow rate considerably. A viable solution is to pre-treat
AMD with ash and then treat it with VA.

26
100

Percent removal (%) 80

60

40

Fe
20
Mn
Zn
0
0 1 2 3
Mass (g)

Figure 3-14: Percent removal of heavy metals (y-axis) as a function of VA mass (x-axis)

4 In situ remediation

This section presents a number of steps taken to achieve the final goal of the project
– applying the ashes in situ for AMD treatment in Almasu Mare. In the following
designs, we have considered on-site conditions. We focused on realistic AMD contact
time with the materials, water flow rate through the media and ways of packing the ash
so that it is easy to handle.

4.1. Improving the efficiency of AMD treatment using a combination of ashes

In this experiment, only Fe related results were regarded. The previous batch
experiment EBFA results showed ~ 60% Fe2+ removal, after 24 hrs of AMD treatment.
VA at L/S = 17 and 24 hrs of experiment also showed ~ 95% Fe2+ removal. Here, we
used a combination of EBFA and VA to increase the efficiency of the process when
contact time is much shorter than 24 hrs. We chose to pre-treat the samples with EBFA
and then treat them with VA, which would ensure the removal of the remaining heavy
metal ions, as well as provide a sponge-like medium to trap sludge formed by the
EBFA treatment. A control run was only run with VA.

27
10 samples of 1.5g VA were run in a batch setting in 25 mL AMD. A sample was
filtered out every 10 minutes for the first 5 samples, then every 30 min, up to 200 min.
Figure 4-1 shows the kinetics of the VA metal removal. This was the control run to
compare with VA metal adsorption behavior after EBFA pre-treatment. Another 10
samples were pre-treated with 0.05 g (L/S = 500) EBFA for 1 hr, after which 1.5 g of
VA was added to each sample and the above procedure repeated.

Kinetics of this reaction are shown in Figure 4-1. The adsorption doesn’t reach
equilibrium after 200 min (which would be shown by a constant qe with time), which
means that there are more adsorption sites available for ion-exchange in the zeolite
for both the control experiment and the pre-treated one.

The pre-treated sample showed 70% removal within the first 10 min of the experiment
and a very good removal of occurs after 80 min (~85%); this is ~30% higher than for
the control sample. Experiments still need to be run to find the balance between the
amount of FA in the pre-treatment step and the amount of VA, as well as the optimal
contact time, however using EBFA and VA in combination looks promising.

80

70

60
Percent removal (%)

50

40

30

20
VA control
10 EBFA pre treatment

0
0 50 100 150 200
Time (min)

Figure 4-1: Fe percent removal with time in VA control samples and EBFA pre-treated samples

28
4.2. Small column experiments

A silica glass column of height h = 25 cm and a diameter of d = 2.0 cm was used for
these experiments. The column had an inlet connected through a thin plastic hose to
a 2L glass vessel containing the water source, and an outlet connected to a peristaltic
pump, with a variable flow rate (details in Figure 4-2).

Figure 4-2: Experimental setup of small column experiments

4.2.1. First small column design: thick layers of TFA and VA

The first experiment was done using 3.0 g of VA and 0.5 g TFA. The VA (1.0 cm layer)
was added to the column first, and the TFA (1.0 cm layer) over it. A thin layer of coarse
sand (0.4 cm) was put between the VA and the FA. A sand layer (1.0 cm) was also
placed over the TFA in order to keep the smaller particles from flowing into the water
(Figure 4-3). The pump was set at its maximum velocity of 220 rotations per minute.
The flow rate was determined by measuring the time taken to filter 10 mL of water
through the column. It was found to be 0.01 L/min. After several minutes, the filter
clogged up; FA was made out of such fine powder that the pressure created by the
vacuum pump compressed the particles together until it created a vacuum in the filter
and no water was able to pass through.

29
Figure 4-3 First column setup, thick layers of VA, TFA and sand

4.2.2. Second small column set-up: VA and multilayers of TFA

A revised filter set-up was improvised; multiple FA layers (each layer was 1.0 cm in
height) with thin layers of coarse sand (1.0 cm) and a thicker layer of VA at the base
(2.0 cm) (Figure 4-4). A flow of 0.003 L/min was observed for 3-4 min, after which the
filter clogged again.

Figure 4-4 Second column set-up using thin layers of sand, FA and VA

4.2.3. Final small column set-up: mixture of VA and TFA

A final set-up was attempted. A layer of VA was placed first; as this is largely formed
of 1.5 mm in diameter particles, it is unlikely that this material cause the blockage. A
layer of coarse sand mixed with FA was added over the zeolite, and an additional thin
layer of sand was placed over the FA mixture.

Not fluidizing the column and its small diameter induced a wall effect; the water flowing
from the top found a path of least resistance through the filter. Also, due to the different

30
densities of the FA and the coarse sand, the sand tended to settle down first, leaving
the FA above and causing clogging. However, a sample of 25 mL was able to be
collected after 5 min of run time (pH=11.9). ICP-OE analysis showed a 91% removal
of Fe, and ~100% Mn and Zn. The leached out Ca was 67 mgL-1. An extra sample of
about 10 mL was collected after 10 min of running (pH = 11.6).

4.2.4. Results of small column experiments

The conclusion drawn from all the filter experiments was that this particular setting of
zeolites and fly ash – placing them together in their raw form, with no other support –
is not effective in an in situ remediation unit. In a larger setting, such as a column with
a diameter several times larger than the small column, there should be less clogging
occurring, since the same amount of material will be spread over a larger surface area.
There would still be the problem of fluidizing the filter every time it is changed, which
cannot be done in situ, and decreases from the efficiency of the materials used for
treatment as well. These experiments did demonstrate the efficiency of the materials
for metal and pH remediation.

4.3. Synthesis of ash and cement cubes – packing of materials for in situ application

A filter bed setting was shown to be difficult to control in a laboratory scale, hence
even more difficult to implement in an on-site application. This section presents a more
suitable way to pack the ashes by synthesizing cubes, with a cement binder. In situ,
loose wall-like structures can be made out of these cubes, thus allowing AMD to pass
through the spaces and be treated. Results of their AMD treatment in a batch setting
are also presented.

5 x 5 cm cubes of TFA and volcanic ash (VA) were synthesized, using Ordinary
Portland cement (C) as a binder, and water (W). Two different C/TFA/VA ratios were
used for moulds A and B (Table 2 and Figure 4-5), whilst moulds C, D and E contained
a cube made only of cement, one of C and TFA and one of C and VA (Table 3 and
Figure 4-6).

31
Mould name Weight of materials (g) Ratios of materials
C TFA VA W C FA VA W
A 16 24 24 32 1 1.5 1.5 2
B 8 24 24 32 1 3 3 4
* NB: three cubes of each A and B were produced using this amount of materials

Table 2 :Moulds of mixtures of C, FA and VA *

Mould name Weight of materials (g) Ratio of materials

C W FA VA C W FA VA
C 25 15 - - 1 0.6 0 0
D 2.25 9 13.5 - 1 4 6 0
E 6.25 9 - 13.5 1 1.4 0 2.2
** NB: one cube of each C, D and E were produced using this amount of materials

Table 3: Moulds for comparison with C, or C with either FA or VA **

Figure 4-5: Moulds A and B

Figure 4-6: Moulds C, D and E

C, FA and VA absorb water differently and have quite different particle sizes, hence
reach similar consistency pastes at different material to water ratios. The criterion for
comparison of cubes C, D and E was the density of the paste created. For cubes A

32
and B we tried to see the difference in the heavy metal removal efficiency of various
TFA to VA ratio cubes. We tried to minimize the amount of cement used, while still
maintaining the integrity of the cubes when in contact with water and achieving efficient
AMD removal. Batch experiments were carried out by placing a cube of each ratio of
materials in 250 mL of water and leaving it gently stirring for 24 hrs.

ICP-OES showed similar results for Fe in samples A and B; with ~ 65% Fe was
removed. For Mn, A removed only 12%, whereas B 90% and for Zn A removed 74%
and B 95%. From C, D, and E, we tried to gain a rough estimate of the efficiency of
solely FA and V cubes and determine if cement itself has an active role in AMD
remediation. Cube C removed 77% of Fe and 100% of Mn and Zn, cube D removed
almost all the metals and cube E removed 75% Fe, 71% Mn and 100% Zn. From
these results, no conclusion can be drawn dealing with the efficiency of each material
in the cubes, but the overall efficiency of this type of material packing seems to be
good. TFA has very fine particles of carbon, which is a good adsorbent, hence adsorbs
a high volume of water, and less cement is needed to form a dense paste. VA has
larger particles and doesn’t adsorb water as fast or as well, hence more cement is
needed for the same cube volume to produce a dense paste. For C only samples, a
very small amount of water is needed.

5 Conclusion

This study concludes that EBFA, TFA and VA are good candidates for the remediation
of AMD in the Romanian mining area Almasu Mare. SEM-EDX did not show any
harmful leachates or compounds formed as a result of AMD treatment. EDX mapping
found silicates and metal (oxy)hydroxides in all three ashes (control sample and
treated sample). Silicate presence is confirmed in TFA by IR spectroscopy and XRD
(which shows crystalline quartz), in EBFA by IR spectroscopy, and in VA by IR
spectroscopy and XRD (which indicates the presence of clinoptilolite). Metal
(oxy)hydroxide presence in EBFA and TFA is confirmed by IR spectroscopy. EDX
analysis also shows some indication of Fe-oxides and Fe-sulfates; this is confirmed
by Raman spectroscopy in TFA. Inspection also indicates the presence of carbon in
EBFA and TFA. This is confirmed by Raman spectroscopy.

33
All the materials show efficient AMD treatment, at various liquid to solid ratios. For in
situ implementation of EBFA, an L/S = 250 should be used; this amount of EBFA fully
neutralizes the solution and removes almost all the Fe and Mn ions. TFA shows full
removal of all the heavy metal ions for L/S = 100, but the resulting pH is slightly high
(pH = 9.5). Care should be taken with these ashes due to the fact that their pH
neutralization mechanism involves Ca leachates, which increase the total dissolved
solids over maximum admitted limits. VA at L/S = 17 removed > 85% of all the heavy
metals and the resulting solution pH was 5.7.

The heavy metal removal mechanism for EBFA and TFA is adsorption. EDX analysis
shows the presence of all AMD specific elements on the surface of both EBFA and
TFA after treatment. Fe and Mn present in point analysis in control samples increase
in average weight percent in samples after treatment. ICP-OES results show efficient
remediation to occur in a short amount of time, which support this conclusion.
Adsorption isotherms derived from ICP results also support this conclusion; they show
both ashes to obey type II isotherm, which is specific to fine – non - porous solids. The
main metal removal mechanism in VA is ion - exchange. EDX mapping shows the
presence of some but not all the AMD specific elements, even though ICP-OES
confirms their removal, which indicates AMD compounds must have been sorbed into
the zeolite structure. SEM shows the zeolite formations unaffected by AMD, hence
allowing the possibility of ion – exchange. The adsorption isotherm obtained from ICP
data shows a type I isotherm, which is typical for microporous solids.

Using these materials for in situ AMD treatment is definitely a feasible option. A
combination of EBFA and VA is efficient for treatment at realistic contact times.
Various filter bed designs were tested, however the small particle size of TFA cause
clogging and are difficult to handle. Synthesizing mixed cubes of TFA, VA and cement
is a good alternative to pack the materials.

6 Further work

Clearly, proposing an on-site remediation design requires more work. A plan


outlining subsequent steps required to achieve this goal is presented below:

34
Further work outline

Completion of tests with the already On site geographical study


developed methodology
AMD creeks
BET on parameter AMD geography
Cubes Optimization of raw production rate of location
characterization cubes' parameters materials
collection

What are some easy ways to roughly estimate the


efficiency of the AMD treatment in situ, in real
time?
- Fe determination by inspection

What is the best way to pack our materials?


- powders in pouches of semi-permeable
membranes
- cubes

What is the best treatment design for our location?

A more extensive set of experiments should be run on the cubes. Ratios of the
materials should be varied in small steps, as should contact time with the AMD. The
software Minitab should be used to plan for the optimal number of experiments. SEM-
EDX and perhaps vibrational spectroscopy should be run on the cubes in order to see
what species are being formed on their surface and cross reference that with the data
that we have from the individual materials. A BET (Brauner-Emmet-Teller) test should
be performed on the cubes as well as the individual materials, in powder form, in order
to determine this packing method’s effect on the porosity of the material. This should
also give an indication of the available ion-exchange micropore sites in the zeolite.

Extensive field data on the geographical sites of the AMD affected rivers should be
collected; this should include depths, widths, loads and velocities and discharges of
watesr.

35
The amount of Fe in the AMD is significantly higher than any of the other heavy metals,
hence a rough determination of its concentration change should give us a good idea
about how the remediation materials are performing. A standard colorimetric
determination method based on the reaction of Fe2+ with o-phenanthroline can be
applied easily [51]. An album of standard solutions of known concentrations of Fe 2+
within the expected AMD range should be prepared and samples of AMD before and
after different stages of treatment should be cross-referenced against them.

Alternative ways to pack the ash should be explored. The powder materials could be
placed in semi-permeable membranes, or even dense fabrics from cheap natural
materials such as jute or cotton. If cement based cubes are adopted, we should find
ways to increase their porosity (for example by adding expanded polysterene beads
into the mixture and then melting them) in order to allow for a greater surface area of
contact between the AMD and the ash.

A potential design fitted to the Almasu Mare region is presented in Figure 6-1.
Dr. Luiza Campos and Dr. J.K. Kim in CEGE, UCL will take over the project. If the
application for funding is successful, we hope to have a small-scale fully functional
optimized AMD treatment model within the next academic year.

Figure 6-1: 2 In-situ AMD treatment, designed by Dr. J.K. Kim

36
Bibliography
1. Järup, L., Hazards of heavy metal contamination. Br. Med Bull 2003. 68(1): p.
167-182.
2. Gaikwad, R.W., Gupta, D.V, Acid mine drainage (AMD) management. Jr. of
Industrial Pollution Control, 2007. 23(2): p. 285-297.
3. Selim, H.M., Competitive Sorption and Transport of Heavy Metals in Soils and
Geological Media. p. 127.
4. Blowes, D.W., Ptacek, C.J., Jambor, J.L, Weisener, C.G. , The Geochemistry
of Acid Mine Drainage. Treatise on Geochemistry, 2003. 9: p. 19.
5. Nobutada, N., Safe and Reliable Drinking Water by the Ecological Filter of Slow
Sand Filtration. 2008(Ueda, Nagano June 16-17).
6. Sima M., D.B., Frei L., Senila M., Balteanu D., Zobrist J., Sulfide oxidation and
acid mine drainage formation within two active tailings impoundments in the
Golden Quadrangle of the Apuseni Mountains, Romania. J Hazard Mater,
2011. 189(3): p. 624-639.
7. Levei E., R.M., Miclean M., Borodi G., Senila M., Acid mine drainage prediction
for tailings in the Baia Mare and Southern Apuseni mining areas, Romania.
Carpathian J. of Earth and Environmental Sciences, 2013. 8(2013): p. 3.
8. Milu V., L.J., Peiffert C, Water contamination downstream from a copper mine
in the Apuseni Mountains. Environmental Geology, 2002. 42(7): p. 773-782.
9. Reynolds, T.D., Richards, P.A., Unit Operations and Processes in
Environmental Engineering. PWS Publishing Company, 1996(2): p. 128 - 140,
210-214.
10. Federation, W.E., Industrial wastewater management, treatment and disposal
3rd ed. McGraw-Hill, 2008: p. 186, 377.
11. Ersoz, M., Barrett, L., Best Practice Guide on Metals Removal from Drinking
Water by Treatment. IWA Publishing, 2012: p. 28.
12. Muhammad, N.P., J.; Smith, M.D.; Wheatley, A.D. , Removal of Heavy Metals
by Slow Sand Filtration. Proceedings of the 23rd WEDC International
Conference on Water Supply and Sanitation, Durban, South Africa, 1997: p.
167-170.
13. Emerick, J.C., Huskie, W.W., and Cooper, D.J., Treatment of discharge from a
high elevation metal mine in Colorado Rockies using an existing wetland.
Proceedings of a Conference on Mine Drainage and Surface Mine
Reclamation, 2001. I: Mine Water and Mine Waste(US Dep. Inter., Bur. Mines
Info Circ. No. IC-9183): p. 345-351.
14. Fu, F., Wang, Q. , Removal of heavy metal ions from wastewaters: A review. J.
of Environmental Management 2011. 92: p. 407-418.

37
15. Sud, D., Mahajan, G., Kaur, M.P., Agricultural waste material as potential
adsorbent for sequestering heavy metal ions from aqueous solutions – A
review. Bioresource Technology 2008. 99: p. 6017-6027.
16. Sajwan, K.S., Alva, A.K, Keefer, R.F., Chemistry of Trace Elements in Fly Ash.
Springer, 2003: p. 5-9.
17. Daci, M.N., Daci, N.M., Zeneli, L., Gashi, S., Hoxha, D., Coal ash as adsorbent
for heavy metal ions in standard solutions, industrial wastewater and streams.
Ecohydrology and Hydrobiology, 2011. 11(1-2): p. 129-132.
18. Alinor, I.J., Adsorption of heavy metal ions from aqueous solution by fly ash.
Fuel 2007. 86: p. 853-857.
19. Ahmaruzzman, M., Industrial wastes as low-cost potential adsorbents for the
treatment of wastewater laden with heavy metals. Advances in Colloid and
Interface Science 2011. 166: p. 36-59.
20. Panday, K.K., Prasad, G., Singh, V.N., Removal of Cr(V1) from aqueous
solutions by adsorption on fly ash-wollastonite. J. Chem. Technol. Biotechnol.,
1984. 34A(7): p. 367-374.
21. Vadapalli, V.R.K., Gitari, M.W., Petrik, L.F., Etchebers, Ellendt, A. , Integrated
acid mine drainage management using fly ash. J Environ Sci Health A Tox
Hazard Subst Environ Eng. , 2012. 47(1): p. 60-9.
22. Vadapalli, V.R.K., Klink, M.J., Etchebers, O., Petrik, L.F., Gitari, W., White,
R.A., Neutralisation of acid mine drainage using fly ash and strength
development of the resulting solid residues. S.Afr. J. Sci., 2008. 104: p. 317-
322.
23. Motsi, T., Rowson, T. N.A., Simmons, M.J.H., Adsorption of heavy metals from
acid mine drainage by natural zeolite. Int. J. Miner. Process, 2009. 92(1-2): p.
42-48.
24. Erdem, E., Karapinar, N., Donat, R., The removal of heavy metal cations by
natural zeolites. Journal of Colloid and Interface Science, 2004. 280(2): p. 309-
314.
25. Taffarel, S.R., Rubio, J., On the removal of Mn2+ ions by adsorption onto
natural and activated Chilean zeolites. Minerals Engineering, 2009. 22(4): p.
336-343.
26. Ostroski, I.C., Barros, M.A., Silva, E.A., Dantas, J.H., Arroyo, P.A., Lima, O.C.,
A comparative study for the ion exchange of Fe(III) and Zn(II) on zeolite NaY.
J. Hazard. Matter, 2009. 161(2-3): p. 1404-12.
27. Ouki, S.K., Kavannagh, M., Treatment of metal-contaminanted wastewaters by
the use of natural zeolites, Water Science and Technology, 1999. 39(10-11): p.
115-122.
28. Gaikwad, R.W., S. A. Misal, Dhirendra, Gupta D. V., Removal of metal from
acid mine drainage (AMD) by using natural zeolite of Nizarneshwar Hills of
Western India. Arabian Journal of Geosciences, 2011. 4(1-2): p. 85-89.
29. Levei, E., Roman, M., Miclean, M., Borodi, G., Senila, M., Acid mine drainage
prediction for tailings in the Baia Mare and Southern Apuseni mining areas,
38
Romania. Carpathian J. of Earth and Environmental Sciences, 2013. 8(3): p.
167-174.
30. Mahlangu, T.O., Mpenyana-Monyatsi, Momba, M.N.B, Mamba, B.B., A
simplified cost-effective biosand filter (BSFZ) for removal of chemical
contaminants from water. J. Chem. Eng. Mater. Sci. , 2011. 2(10): p. 156-167.
31. Taylor, H.F.W., The Chemistry of Cements. Academic Press, 1964. 1: p. 4.
32. Gitari, W.M., Petrik, L.F., Keys, D.L, Okujeni, C., Interaction of acid mine
drainage with Ordinary Portland Cement blended solid residues generated from
active treatment of acid mine drainage with coal fly ash. J Environ Sci Health A
Tox Hazard Subst Environ Eng., 2012. 46(2): p. 117-37.
33. Ciubota-Rosie, C., Gavrilescu, M., Macoveanu, M., Biomass – and important
renewable source of energy in Romania. Environmental Engineering and
Management Journal, 2008. 2008(7): p. 559-568.
34. Tetisan, A.I., Studii si cercetari privind utilizarea tufurilor zeolitice din zona
Barsana in tehnologii neconventionale de epurare a apelor uzate. Babes-Bolyai
University, Environment Science Faculty, 2010. PhD Thesis.
35. Bedelean, H., Maicaneanu, A., Burca, S., Stanca, M., Investigations on some
zeolitic volcanic tuffs from Cluj County (Romania)- used for zinc ions removal
from aqueous solution. Studia Universitatis Babes-Bolyai, Geologia, 2010.
55(1): p. 9-15.
36. Kutchko, B.G., Kim, A.G. , Fly ash characterization by SEM-EDS. Fuel 2006.
75: p. 2537-2544.
37. Minerals of Wales., Jarosite.
http://www.museumwales.ac.uk/800/?mineral=168.
38. Oliveira, C.R., Rubio, J., New basis for adsorption of ionic pollutants onto
modified zeolites. International Journal of Coal Geology 2007. 73: p. 359-370.
39. Socrates, G., Infrared and Raman Characteristic Group Frequencies: Tables
and Charts, 3rd Edition. 2004: Wiley.
40. Guedes, A., Valentim, B., Prieto, A.C., Sanz, A., Flores, D., Noronha, F.,
Characterization of fly ash from a power plant and surroundings by micro-
Raman spectroscopy. International Journal of Coal Geology 2008. 73: p. 359-
370.
41. Hodkiewicz, J., Characterizing Carbon Materials with Raman Spectroscopy.
Thermo Fischer Scientific. Application Note: 51901.
42. Project, T.R. Clinoptilolite- Na. http://rruff.info/Clinoptilolite-Na.
43. Hitachi High Tech Science Corporation, ICP analysis. http://www.hitachi-hitec-
science.com/en/products/icp/tec_descriptions/descriptions1_e.html.
44. University of Oxford, Surface Area Analysis.
http://saf.chem.ox.ac.uk/operating-principles-3.aspx.
45. Faust, S.D., Aly, O.M., Chemistry of Water Treatment. 1983: p. 52.

39
46. Srivastava, S., Goyal, P., Sorption Isotherms and Kinetics. Novel Biomaterials,
Environmental Science and Engineering, 2010: p. 87-91.
47. Tosun, I., Ammonium Removal from Aqueous Solutions by Clinoptilolite:
Determination of Isotherm and Thermodynamic Parameters and Comparison
of Kinetics by the Double Exponential Model and Conventional Kinetic Models.
Int. J. Environ. Res. Public Health, 2012.
48. Buasri, A., Chaiyut, N., Phattarasirichot, K., Yongbut, P., Nammueng, L., , Use
of Natural Clinoptilolite for the Removal of Lead(II) from Wastewater in Batch
Experiment. Chiang Mai J. Sci., 2008. 35(3): p. 447-456.
49. Korkmaz, M., Özmetin, C., Özmetin, E., Copper Sorption by Clinoptilolite:
Equilibrium and Thermodynamic Investigation. Sixteenth International Water
Technology Conference, IWTC 16 2012, 2012. Istanbul, Turkey.
50. Jacobs, P.A., Flanigen, E.M., Jansen,J.C. , van Bekkum, I., Introduction to
Zeolite Science and Practice. Elsevier Science Publishers B.V., 1991. 58: p.
503.
51. Spectrophotometric Determination of Iron http://http-
server.carleton.ca/~rburk/high_schools/speclab.pdf.
52. World Health Organization, Total dissolved solids in Drinking-water.
WHO/SDE/WSH/03.04/16.
http://www.who.int/water_sanitation_health/dwq/chemicals/tds.pdf

40
7 Appendices

7.1. Appendix 1. Sample collection of natural AMD, Almasu Mare, Romania

Almasu Mare is situated in the Apuseni mountain region in


Romania. It is part of the Golden Quadrilateral (Figure
beside3) and contains the largest area of gold deposits in
Europe, along with important quantities of Cu, Pb, Zn and
other minerals.

3
Tóth A., Quiquerez A., Márton I., (2006) Past and present mining in the Apuseni Mountains, Romania, Field Trip SEG Student Chapters Uni Geneva - ETH Zürich - Uni Budapest - Uni Cluj

41
The region of Almasu Mare was recommended by Dr. Denisa Jianu, University of Bucharest, (UB) Faculty of Geology and
Geophysics. This was an area where her group had previously done assessments of the water and soil quality, and findings of high
heavy metal content in the soils as well as acid mine drainage had already been confirmed4. Additional information on volcanic zeolitic
tuff geological formation and abundance in Romania were kindly provided by Dr. Barbara Soare from the same department. Dragos
Mitrica (UB Geology and Geophysics masters student) assisted with the volcanic tuff- Traistari quarry- and water sample collection.

Water samples were collected from three different creeks near Almasu Mare (Fig i), one creek directly collected AMD from Hanes
mine (Fig. ii), one from Rades mine (Fig. iii) and one was a stream which collected both of the previous creeks – the collection point
was about 500-600 m from the mines, at the entrance to a nearby village (Fig. iv)).

4
Iacob C., Panea I., Orza R., Furnica M., Jianu D., Mocanu V., (2011) Geophysical Investigations in the Polluted Mining Area of Zlatna, Romania - Final Report,
http://www.geo.edu.ro/gwb/images/Raport_final_resize.pdf
Iacob C., Panea I., Orza R., Furnica M., Jianu D., Mocanu V., (2011) Integrated geophysical and geochemical investigations for identifying potable water sources on Ampoi
Valley, Romania, Society of Exploration Geophysicists Annual Meeting, September 18 - 23, 2011, San Antonio, Texas

42
43
Fig. iv) Stream collection point at the entrance of the nearby village (left side) and village view (right side)

44
7.2. Appendix 2: Water quality data on the collected AMD, analyzed by Ecoind

Identification of samples: AMD: Hanes (1), Rades (2), Village river (3)

Measured Unit Sample Romanian


parameters maximum
admitted
values HG
1 2 3 352/2005
NTPA 001
-
pH 2.72 2.69 2.79 6.5-8.5
-1
Conductivity mS cm 3.84 5.60 2.99 -
sulfates mg/L 2921 5789 1938 600
Zn h/f mg/L 39/38 167/161 47/45 0.5
Fe h/f mg/L 326/152 1360/1064 127/72 5.0
Mn h/f mg/L 278/271 124/119 106/103 1.0
Cd h/f μg/L 54/48 209/189 65/62 200
Cu h/f μg/L 68/64 61/51 190/190 100
Ni h/f μg/L 201/186 215/180 101/101 500
Pb h/f μg/L 13/12 5/4 15/15 200
Al h/f mg/L 26/16 171/81 62/57 5.0
As h/f μg/L 16/14 <0.01 10/9 100
h/f – homogeneous probe/ filtered probe

45
7.3. Appendix 3 Particle size distribution of ash

Particle size distribution data for the three Particle size distribution
Six sieves of different sizes (named I – 70
types of ash
VI) were used (figure below). 10.0 g of
VA, 2.0 g of EBFA and 2.0 g TFA were Am. of 60
Ash Sieve ash in
passed through the sieves. The Sieve# 50
type size sieve
samples from each sieve were (mm) (g)

Percent of sample (%)


I 2.8 - 40
collected and weighed individually.
II 1.4 - EBFA
This produced the particle size 30
EBFA III 0.35 0.22 TFA
distribution for the ashes. IV 0.15 0.76
20 VA
V 0.07 0.93
VI 0.05 0.02
I 2.8 - 10
II 1.4 0.04
TFA III 0.35 1.3 0
0 1 2 3
IV 0.15 0.63
V 0.07 - -10
Particle diameter (mm)
VI 0.05 -
I 2.8 0.28
II 1.4 6.34
VA III 0.35 2.87
IV 0.15 0.02
V 0.07 0.02
VI 0.05 0.02

46
7.4. Appendix 4. TDS and pH results of selected experiments

Total dissolved solids (TDS)

TDS is the total amount of dissolved inorganic salts and organic materials present in a water sample (measured in mgL-1). It was
measured on a Sper Scientific Benchtop Conductivity/TDS/Salinity Meter -860032. According to the World Health Organization, the
drinking water limits are < 300 mgL-1 (excellent quality) to 1200 mgL-1 (poor) [52]. Both TFA and EBFA had higher than the maximum
admitted limit values of TDS for the L/S needed to fully neutralize the AMD.

A. EBFA

EBFA Sample mass, its reaction time, TDS and pH results

Ce (mg/L)
Sample Mass TDS pH Ca2+ Mg2+ Fe2+ Mn2+ Zn+2
(mg/L)
original 0 652 2.7 0.00 0.00 305.03 54.11 95.18
SAMD
1 hr 0.035 1168 4.1 * * * * *
1 hr 0.050 1442 4.4 197.55 1.49 63.37 39.80 55.72
1 hr 0.100 2350 7.0 313.30 2.24 0.09 19.89 0.74
1 hr 0.250 4770 10.2 * * * * *
24 hrs 0.050 1368 5.1 233.57 2.01 31.51 44.81 70.79
24 hrs 0.100 2240 7.4 297.63 2.04 0.04 17.10 0.58

47
Fig. 7A.ii. pH change with EBFA mass Fig. 7 A.i. TDS as a function of EBFA mass
12.0 6000
10.0 5000

TDS (mg/L)
8.0 4000
pH

6.0 3000
1 hr
4.0 2000
24 hrs
2.0 1000

0.0 0
0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.000 0.050 0.100 0.150 0.200 0.250 0.300
Mass (g) Mass (g)

Fig. 7 A.iii. Ca and Mg ion leachates with increase in EBFA


mass
350.00
300.00
250.00
Ce ( mg/L)

200.00 Ca 1 hr
150.00 Mg 1 hr
100.00
Ca 24 hrs
50.00
0.00 Mg 24 hrs
0.000 0.050 0.100 0.150
Mass (g)

48
The TDS showed an increase with the increase in mass of EBFA, but
Fig. 7 A.iv. Equilibrium concentration of Fe, Zn,
hardly any change with respect to the reaction time. This is confirmed by Mn ions for 1hr and
the Ce of Ca and Mg leachates in Fig. 7iii.; Mg concentrations are virtually 24 hrs, as a function of EBFA mass
constant at about 2 mgL-1, whilst the Ca concentration increases with time 350.00

for the first mass of EBFA, but slightly decreases for the second one, so 300.00

there isn’t any trend with respect to time. The pH results don’t show any 250.00 Fe 1 hr

change with time either; an optimal pH is obtained for 0.100 g EBFA (Fig. Fe 24 hrs
200.00
Mn 1 hr
7ii.). Most of the heavy metal uptake from the solution is shown to happen
150.00 Mn 24 hrs
within the first hour of reaction (Fig. 7iv.). The concentration of these still
100.00 Zn 1 hr
decreases for the 24 hrs experiments. Zn 24 hrs
50.00
Based on all of these results, we may conclude that EBFA remediates
0.00
AMD in a short amount of time, t ≤ 1 hr at L/S = 250, efficiently. The TDS 0 0.05 0.1 0.15
for this point is however about 1 gL-1 higher than the maximum admitted
limit for drinking water, hence an extra TDS removal step should be inserted in the in-situ remediation.

49
B. TFA
TFA Sample mass, its reaction time, TDS and pH results
Fig. 7 B.i. TDS as a function of TFA mass
Ce (mg/L)
Sample Mass TDS pH Ca2+ Mg2+ Fe2+ Mn2+ Zn+2 900
(mg/L) 850
original 0 652 2.7 0 0 305.03 54.11 95.18 800
s. AMD 750

TDS (mg/L)
1 hr 0.035 569 5.0 * * * * * 700
1 hr 0.050 600 5.0 * * * * * 650
1 hr 0.100 669 5.7 142.54 4.98 85.75 42.22 58.08 600
1 hr 0.250 837 7.0 242.51 5.14 0.07 24.78 0.48 550
500
24 hrs 0.100 659 7.0 170.32 18.51 1.59 42 37.45
450
24 hrs 0.250 859 9.3 279.87 22.58 0.8 0.48 0.54
400
0.000 0.050 0.100 0.150 0.200 0.250 0.300
Mass (g)
* No data provided
Fig. 7 B.iii. pH change with TFA mass
Fig. 7 B.ii. Ca and Mg ions leachate 10.0
300 9.0
250 8.0
200
Ce (mg/L)

pH
7.0
Ca 1 hr
150 6.0
Mg 1 hr
100 5.0
Ca 24 hrs
50 4.0
Mg 24 hrs 0.000 0.050 0.100 0.150 0.200 0.250 0.300
0
0.000 0.100 0.200 0.300 Mass (g)
Mass (g)

50
TFA shows a lower rate of Ca and Mg oxides
Fig. 7.B.iv. Equilibrium concentration of Fe, Zn, Mn ions for 1hr and
dissolution than the EBFA, as the results for TDS and 24 hrs, as a function of TFA mass
especially for pH change with time. This is confirmed by the 350

ICP results, which show a higher quantity of Ca2+ leached out 300

with time. The equilibrium concentration of Ca ions is high – 250

Ce (mg/L)
Fe 1 hr
almost 200 mgL-1 for 0.05 EBFA, comparing with zeolites 200 Mn 1 hr
which show 35 mgL-1 for 1.00 g of volcanic ash – while the 150 Zn 1 hr

Mg2+ concentration is low, having a maximum of about Fe 24 hrs


100
Mn 24 hrs
2 mgL-1. This supports the conclusion that EBFA’s
50 Zn 24 hrs
mechanism of pH correction is by dissolution of alkaline
0
materials, specifically CaO, not by adsorption of H+. 0 0.05 0.1 0.15 0.2 0.25 0.3
Mass (g)

There is a drastic decrease in the Ce of Fe2+ ions from time 1 hr to 24 hrs, however about 2/3 of iron is removed within the first
hour of the experiment for 0.1 g TFA. The changes with time in the Ce of Mn and Zn ions aren’t of comparable magnitude.

51
7.5. Appendix 5. XRD analysis

A. EBFA before experiment, showing Portlandite (P) peaks


relative intensity

degrees 2θ

52
relative intensity B. EBFA after experiment, showing a broad peak, characteristic of amorphous material

degrees 2θ

53
C. TFA before experiment, showing peaks of quartz (Q)
relative intensity

degrees 2θ

54
relative intensity D. TFA after experiment, showing peaks of low crystalline quartz

degrees 2θ

55
E. VA, showing peaks of clinoptilolite
relative intensity

degrees 2θ

56
7.6. Appendix 6. IR-ATR spectra

A. VA IR spectra of samples before and after AMD treatment


Relative intensity (a.u.)

wavenumbers (cm -1 )

57
Relative intensity (a.u.) B. TFA IR spectra of samples before and after AMD treatment

wavenumbers (cm -1 )

58
Relative intensity (a.u.) C. EBFA IR spectra of samples before and after AMD treatment

wavenumbers (cm -1 )

59
7.7. Appendix 7. Raman spectra

A. TFA, showing crystalline quartz and potential calcite peaks

TFA before experiment - white particle, laser wavelength 785 nm, 1 % power
1400

1200

1000
Raman intensity (a.u.)

800

600

400

200

0
0 500 1000 1500 2000 2500 3000 3500
Raman shift (cm -1)

60
B. VA, showing typical clinoptilolite

VA before experiment - laser wavelength 514 nm, 50% power


800

700

600

500
Raman intensity (a.u.)

400

300

200

100

0
0 500 1000 1500 2000 2500 3000 3500

-100
Raman shift (cm-1)

61
C. VA after experiment, showing potential Fe-oxides and Fe-sulfates

VA after experiment - brown particle, laser wavelength 514 nm, 100% power
6000

5000

4000
Raman intensity (a.u.)

3000

2000

1000

0
0 500 1000 1500 2000 2500 3000 3500
Raman shift (cm-1)

62
7.8. Appendix 8. Maximum adsorption capacity of ashes as a function of pH

EBFA maximum adsorption capacity vs pH TFA maximum adsorption capacity vs pH


180.00 300.00
160.00
250.00
140.00
120.00 200.00

qe (mg/g)
qe(mg/g)

100.00
Fe 150.00 Fe
80.00
60.00 Mn 100.00 Mn
40.00 Zn Zn
50.00
20.00
0.00 0.00
0 2 4 6 8 10 12 0.0 2.0 4.0 6.0 8.0 10.0
pH pH

VA maximum adsorption capacity vs pH


10.00
8.00
qe (mg/g)

6.00
Fe
4.00
Mn
2.00 Zn
0.00
3.0 4.0 5.0 6.0 7.0 8.0 9.0
pH

63
7.9. Appendix 9. Radii of hydration and hydration enthalpies

Cation rh (Å) Standard molar


hydration enthalpy5
(ΔhydH°) (kJ/mol)
H+ 2.82 -1103
Fe2+ 4.28 -1972
Fe3+ 4.57 -4462
Zn2+ 4.30 -2070
Mn2+ 4.38 -1874
Na+ 3.58 -416
Ca2+ 4.12 -1602
Mg2+ 4.28 -1949

5
Marcus, Y. (1978), The Thermodynamics of Solvation of Ions, Part 2. – The Enthalpy of Hydration at 298.15 K, J. Chem. Soc., Faraday Trans. 1, 1987, 83, 339-349

64
7.10. Appendix 10. Comparison of our VA sample to other microporous sorbents from literature

Adsorbent Metal K n Source


Our VA showed higher K values than samples of AC,
AC Fe 2+ 0.380 2.175 I. Uzun et
silica polyamine and another natural zeolite for iron ions, al6
Mn 2+ 0.057 1.339
slightly lower values for zinc ions and somewhere Silica Fe 2+ 0.604 2.32 H. Tutu et
polyamine al7
between the reference materials for manganese ions. Mn 2+ 1.257 1.35
composite
Zn 2+ 0.483 8.85
These results are still relative, as the experimental L/S
Natural zeolite Fe 3+ 2.87 6.12 T.Motsi et
as well as the original heavy metal solution different in al8
Mn 2+ 0.64 4.63
the quantity and association with other competitive Zn 2+ 1.09 5.64
metal ions.

6
Uzun, I., Guzel, F., (2000) Adsorption of Some Heavy Metal Ions from Aqueous Solution by Activated Carbon and Comparison of
Percent Adsorption Results of Activated Carbon with those of Some Other Adsorbents, Turk . J. Chem., 24 (2000), 291-297
7
Tutu, H., Bakatula, E., Dlamini, S., Rosenberg, E., Kailasam, V., Cukrowska, E.M., (2013), Kinetic, equilibrium and thermodynamic modelling of the sorption of metals from
aqueous solution by a silica polyamine composite, Water SA, vol. 39, no. 4 (2013)
8
Motsi, T., Rowson, N.A., Simmons, M.J.H., Adsorption of heavy metals from acid mine drainage by natural zeolite, Int. J. Miner. Process. 92 (2009) 42-48

65

Anda mungkin juga menyukai