Sie sind auf Seite 1von 20

Ecological Engineering 37 (2011) 70–89

Contents lists available at ScienceDirect

Ecological Engineering
journal homepage: www.elsevier.com/locate/ecoleng

Filter materials for phosphorus removal from wastewater in treatment


wetlands—A review
Christina Vohla a , Margit Kõiv a , H. John Bavor b , Florent Chazarenc c , Ülo Mander a,∗
a
Department of Geography, Institute of Ecology and Earth Sciences, University of Tartu, 46 Vanemuise St., Tartu 51014, Estonia
b
Water Research Laboratory, University of Western Sydney-Hawkesbury, Locked Bag 1797, Penrith South DC, NSW 1797, Australia
c
Ecole des Mines de Nantes, GEPEA UMR CNRS 6144, 4, rue Alfred Kastler, B.P. 20722, F-44307 Nantes Cedex 3, France

a r t i c l e i n f o a b s t r a c t

Article history: This paper aims to collect and analyse existing information on different filter media used for phosphorus
Received 18 November 2008 (P) removal from wastewater in constructed wetlands. The most commonly used materials are catego-
Received in revised form 4 August 2009 rized as natural materials (considered in 39 papers), industrial byproducts (25 papers) and man-made
Accepted 12 August 2009
products (10 papers). A majority of studies on sorbents have been carried out in lab-scale systems as batch
experiments, and only very few studies have highlighted results on full-scale systems. Among the great
variety of filter media studied, most of materials had a pH level >7 and high Ca (CaO) content. The highest
Keywords:
P-removal capacities were reported for various industrial byproducts (up to 420 g P kg−1 for some furnace
Ash
Bauxite
slags), followed by natural materials (maximum 40 g P kg−1 for heated opoka) and man-made filter media
Constructed wetlands (maximum 12 g P kg−1 for Filtralite). We found a significant positive Spearman Rank Order Correlation
Freundlich equation between the P retention and CaO and Ca content of filter materials (R2 = 0.51 and 0.43, respectively),
Langmuir equation whereas the relation of P retention to pH level was weak (R2 = 0.22) but significant. There is probably an
LECA optimal level of hydraulic loading rate at which the P removal is the highest. Additional important factors
Limestone determining the applicability of filter materials in treatment wetlands such as saturation time, availability
Peat at a local level, content of heavy metals, and the recyclability of saturated filter media as fertilizer should
Saturation
be taken into consideration.
Slags
© 2009 Elsevier B.V. All rights reserved.

1. Introduction In Europe, the supplementary treatment of wastewater from


houses and small rural communities specifically designed for P
Phosphorus (P) is an important nutrient that is critically removal is becoming important for the improvement of environ-
needed for the normal functioning of ecosystems. Nevertheless, mental quality in streams and lakes (Brix et al., 2001).
excess P and nitrogen (N) is the main cause of euthrophication Subsurface flow constructed wetlands (SSF CW) function very
(Vollenweider, 1968; Schindler, 1977; Tiessen, 1995). The main effectively in the treatment of the wastewater of small settlements
input of P into nature occurs through its use in agriculture and from and rural areas. SSF CWs are known to be efficient in the removal of
domestic wastewater (Soranno et al., 1996; Bennett et al., 2001). both biological oxygen demand (BOD5 ) and total suspended solids
Whereas P can be regulated by balanced fertilization in agricultural (TSS) from wastewater. However, nitrogen (N) and phosphorus (P)
use (Sharpley and Tunney, 2000), wastewater needs proper treat- removal is known to be somewhat problematic (Brix et al., 2001;
ment for the removal of P to levels that are acceptable for natural Vymazal et al., 1998).
systems (Kadlec and Knight, 1996). Subsurface flow and other treatment wetlands have been used
Phosphorus removal in conventional small-scale wastewater for the purification of industrial wastewater (agroindustry produc-
treatment systems is a critical issue that has not yet been suffi- tion, paper industry, etc.), mine water, agricultural runoff water
ciently solved. Moreover, continuing eutrophication, especially in and also landfill leachate purification, and most often for domestic
freshwater ecosystems, has led to increasing governmental reg- wastewater (Kadlec and Knight, 1996).
ulatory pressure for the lowering of P concentrations through Phosphorus removal in subsurface flow CWs is closely associ-
enhanced P removal from wastewater (Kadlec and Wallace, 2008). ated with the physical–chemical and hydrological properties of the
filter material, because P is mainly sorbed by or precipitated in filter
media (Faulkner and Richardson, 1989; Kadlec and Knight, 1996;
∗ Corresponding author. Sakadevan and Bavor, 1998; Vymazal et al., 2000). The removal of
E-mail address: ulo.mander@ut.ee (Ü. Mander). P in constructed treatment wetlands is a complicated and difficult

0925-8574/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.ecoleng.2009.08.003
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 71

to sustain process. The life expectancy of sub-surface flow CWs, These data were taken into account and analysed in combination
especially as concerns phosphorous removal, depends on the fil- with P retention values.
ter media that are used for the construction (Grüneberg and Kern, A valuable overview of sorbents that have been used for phos-
2001; Sakadevan and Bavor, 1998). A medium with a high phos- phorus removal was presented in Johansson Westholm (2006). In
phorus binding capacity is an important component of treatment our work we have considered many additional data for substrates
wetlands (Arias et al., 2001). and special sorbents. Analyses of the impact of pH and sorbents
The search for efficient and long-lasting filter media has been such as Ca and CaO are also included.
a key issue for more than a decade. However, the longevity of the Three basic databases are used for this overview: the ISI Web of
sorption capacity of phosphorous is still a main focus for research Science, Scopus, and ISI Proceedings. The main topics used for the
(Johansson Westholm, 2006). In addition to good hydraulic con- database search were as follows: “phosphorus removal”, “phospho-
ductivity, the chemical composition of the adsorption media is a rus retention”, “phosphorus sorption”, “phosphorus adsorption”,
critical factor in CW design. Because phosphorus is removed via and “phosphorus precipitation”. An analogous search was carried
sorption and precipitation processes, Ca, Fe and Al content is impor- out using the terms “phosphorous” and “phosphate(s)” instead of
tant in efficient P removal. Thus filter media should be selected very “phosphorus” in these word combinations. All of the term combi-
carefully. Even if a medium with high P binding capacity has been nations mentioned above were also analysed adding the following
selected, it may be saturated after a few years (Arias et al., 2001). terms: “constructed wetland(s)”, “soil filter(s)”, “sand filter(s)”, and
An obvious and sustainable solution would also be a separate filter “filter material(s)”. In addition, the following names indicating var-
unit containing replaceable material with a high P binding capacity ious filter materials were analysed in combination with the terms
(Brix et al., 2001). In such systems, appropriate pre-treatment will “phosphor(o)us” and “phosphate(s)”: “ash”, “bentonite”, “blast fur-
also allow for a longer lifetime of the filter media, by decreasing the nace slag”, “calcite”, “ceramic”, “clay”, “coal”, “Filtralite”, “gravel”,
risk of clogging and allowing one to use finer reactive filter media “LECA”, “limestone”, “LWA”, “oil shale ash”, “Opoka”, “peat”, “sand”,
with higher sorption capacity (Hedström, 2006). “shellsand”, “slag(s)”, and “wollastonite”.
The most commonly used materials can be broadly categorized
as natural materials (Table 1 ), industrial by-products (Table 2 )
1.1. Sorbents for phosphorus removal
and man-made products (Table 3). A total of 64 papers were used in
our analysis, and in 39 of the papers natural materials were studied,
A great variety of different types of materials for P retention in
whereas in 25 and 10 papers, industrial byproducts and man-made
CWs have been described. Traditionally, locally available sands and
materials were considered, respectively. We gathered data on Ca
soils have been used as filter media in sub-surface flow wetlands
and CaO content and the pH level of the filter media in relation
for retaining P, even without any data about P-removal efficiency
to P retention capacity. Data lacking one of the above-mentioned
(Johansson Westholm, 2006), for example sand filters or gravel
characteristics were excluded. Considering both the lack of data
(Rustige et al., 2003). Studies performed during the last decade have
and the wide range of experimental conditions for full-scale and
demonstrated the importance of finding locally available materials
bench-scale experiments, all of the data considering Ca and CaO
with high P retention capacity. Johansson Westholm (2006) also
content of filter materials came from batch/sorption tests alone.
emphasised the fact that regardless of the great amount of existing
Among the gathered data, very few data were provided about
scientific data, the potential and efficiencies of different materials
pH and CaO. That is why in the CaO analyses, only 37 data points
and studies are very difficult to directly be compared and general-
from 9 papers were used, compared to Ca analyses (78 data points
ized.
available from 22 papers).
There are many studies that use the special treatment of filter
All of the data were analysed for the normality of distribu-
media to enhance P retention capacity (Brogowski and Renman,
tion using the Lillefors test and the programme STATISTICA 7.0.
2004; Kwon et al., 2004; Li et al., 2006). Usually heating Ca-rich
Spearman Rank Order Correlations were provided to describe the
media at high temperatures is a common way to perform material
relationships between the P removal, pH, and Ca and/or CaO con-
improvement. During heating, CaO will probably form, which has
tent of the filter media. The level of significance of ˛ = 0.05 was
a more reactive Ca-phase than commonly existing CaCO3 .
accepted in all cases.
The objectives of this paper are to (1) summarize existing infor-
mation on different filter media used for phosphorus removal from
wastewater that could be applied to one database, with the advan-
3. Results and discussion
tages and disadvantages of each amendment and (2) to clarify
relationship between P retention capacity, pH level, Ca(CaO) con-
3.1. Filter materials used for phosphorus removal
tent, and the hydraulic parameters of the filter media.

3.1.1. Natural products


2. Methods 3.1.1.1. Apatite. Is a group of phosphate minerals, usually referring
to hydroxylapatite, fluorapatite, and chlorapatite, containing OH− ,
Data about various filter materials have been collected to give F− , or Cl− ions, respectively, in the crystal. Sedimentary apatites
an overview of advantages, but also to emphasise the critical condi- are charactarezied by high Ca and P content (31.4% Ca and 11.3%
tions and possible limiting factors for the usage of these materials. P in Bellier et al., 2006; 37.3% Ca and 16.8% P in Molle et al., 2005;
A preliminary screening of available data showed that majority of 37.6% Ca and 15.2% P in Joko, 1984). Presence of P-crystal in apatite
filter materials studied have high Ca and/or CaO content. Based on is believed to act like a catalyst which lowers the activation energy
these data, analysis of the influence of Ca(CaO) content and pH on barrier between the crystal and precipitated P. The distinction
the P retention capacity was performed. However, to give a bet- between adsorption and crystallization is not clear in P-sorption
ter overview of the different types of materials and their potential mechanisms on natural apatites, considering that their concomi-
as a filter media, articles where only P-removal effectiveness or tant action is expected to occur. P sorption from 0.4 up to 4.8 g
retention capacity were also presented, have been included in our P kg−1 apatite have been measured in batch experiments (Bellier
analysis. Some literature sources provided with information on the et al., 2006; Molle et al., 2005; Joko, 1984). Full-scale experiments
hydraulic loading rate (HLR) and hydrologic retention time (HRT). showed promising results (enough to reduce P-level to reach 2 mg
72
Table 1
Natural materials used as filter media for P removal.

Material Study type Description of the study Retention calculation P retention Ca (CaO) References pH

Alunite Batch Adsorption: 1 g of material Calculated Average of over 80% – Özacar (2006) 5
(calcinated) with 1 L of phosphate removal
solution (25–150 mg L−1 ),
contact time 29.1 min
Apatite Batch KH2 PO4 solution of Langmuir Langmuir b = 4.76 mg P 37.3 Molle et al. (2005) 7.0
(sedimentary) 0–500 mg P L−1 with g−1 , K = 0.15
contact time of 1–48 h
Column Synthetic solution 20 mg P Saturation value after
L−1 with hydraulic loading 550 d 13.9 mg P g−1 ;
rate (HLR) 1.35 L d−1 in 550 <2 mg P L−1 in the
days (hydraulic retention outlet
time (HRT) 5.6 h)
Apatite Batch/column 35 g of material in a 1 L Langmuir/Freundlich 0.28–1.09 g P kg−1 7–31.4% Bellier et al. (2006) 8
(igneous and glass flask filled with material in 24 h

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


sedimentary) 700 mL of solution isotherms
(5–150 mg P L−1 ) 24 and
96 h isotherms
Bauxite Batch 2.5–40 mg P L−1 , 20 g and Langmuir Adsorption 0.612 g – Drizo et al. (1999) 5.9
24 h and column study kg−1 . Column: 160 mg
(40d + 40d); HLR 3 L d−1 P kg−1 , longer study:
355 mg P kg−1
Bauxite Batch 1 g substrate and 100 dm−3 Calculated Maximal PO4 3− – Altundoğan and Tümen (2002)
solution, 10 mg P dm−3 and removal 67.3% at pH
shaken for 2 h 4.45
Bauxites (raw Batch 1 g substrate and 100 dm−3 Calculated Adsorption for raw and – Altundoğan and Tümen (2003) 4.5–5.2
and activated) solution, 10 mg P dm−3 and activated bauxite: 0.82
shaken for 2 h and 2.95 mg P g−1 ,
removal >95% for
activated bauxite
Dolomite Batch 0.2 g in 100 mL solution Calculated Sorption 21.7% Karaca et al. (2004) Highest adsorption at pH 11
10–60 mg PO4 -P L−1 , 7.34–52.02 mg P g−1
different adsorption times,
temperature, pH 1–11,
material from Turkey
Dolomite and Batch 10 g sorbate in 50 mL Langmuir 0.168 g P kg−1 28.7% Prochaska and Zouboulis (2006) 7.8
sand solution with 0–100 mg
L−1 , 24 h
0.121 g P kg−1 2.5% 6.9

Gravel Full-scale CW, Batch 3 gravel based CWs Langmuir and Freundlich P removal from −40% – Mann and Bavor (1993)
investigated, secondary to 40%. Adsorption
sewage effluent, 2 years. capacity ranged
Laboratory P adsorption 25.8–47.5 mg P g−1
study
Gravel Full-scale CW Gravel filled wetland, dairy Determined Total P (TP) removal 1.07% Tanner et al. (1999)
farm wastewater, mean 184–296 g m−2 ,
influent concentration substratum TP
15 mg P L−1 , 5 years; HLR accumulation
21.4–71.7 mm d−1 , HRT 115–128 g m−2
1.95–6.54 d
Gravel Full-scale CW VSSF planted gravel filter Calculated PO4 3− -P removal – Korkusuz et al. (2005)
in Turkey, HLR 3 m3 d −1 efficiency of the gravel
(0.1 m d−1 ) system 4.33%
Gravel Pilot-scale CW HSSF CW, planted, 2 years, Calculated Medium gravel 43.9%, 27.2% Akratos and Tsihrintzis (2007)
HRT 6–20 days; HLR with cobbles and
16–55 L d−1 cattail up to 67%
Gravels (South- Batch 3 g sorbate, 75 mL solution Calculated Removal 33–50%, 2–14% (3.6%) Vohla et al. (2005)
Estonian) with 5–1000 mg PO4 -P L−1 , sorption 3–3.6 g P kg−1
24 h (fine gravel: 88.6%)
Laterite Batch, pilot-scale Sorption test with leachate Calculated 99% removal of – Wood and McAtamney (1996)
(10 g granular laterite, phosphates in lab and
solution with 5–50 mg 96% in pilot CW
PO4 -P− L−1 ) and pilot CW
about 2 years
Limerock Meso-scale CW Experimental wetland Calculated TP removal for the – DeBusk et al. (2004)
system received effluent filter system was 0.32 g
from a treatment wetland P m−2 year−1 and 46%

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


for 19 months; HLR 11 cm
d−1
Limestone Batch 1 g substrate, P-solutions Calculated Sorption 0.25–0.3 mg P 21.2% Johansson (1999a) 8.9
(5–25 mg PO4 –P L−1 ), g−1
limestone from Sweden
Limestone Batch 2.5–40 mg P L−1 , 20 g and Langmuir Maximal adsorption – Drizo et al. (1999) 7.8
24 h, material from a capacity 0.682 g kg−1 .
quarry; HLR 3 L d−1
Limestone Full-scale CW SSF wetland cell treating Calculated P retention on average – Hill et al. (2000) 7.2
wastewater from dairy 4.3%. Mean reduction
farm. HRT 13.4 d, 1.5 years 14.5%
Maerl Pilot-scale CW, batch Pilot-scale wetlands in the Calculated TP removal 98%. Batch 32.04% Gray et al. (2000) pH out of tank 7.6
laboratory, 9 weeks, 21 L. removal 32–99%,
HLR 2.3 L d−1 ; HRT 5 d. capacity
Batch: 4 g, 100 mL of P 39.5–7490 mg kg−1
solution (0–5000 mg L−1 ).
72 h
Marl gravel Full-scale CW Filter treating swine Calculated Removal efficiency of 12% Szögi et al. (1997)
wastewater after anaerobic TP 37–52%
lagoon treatment
Opoka Batch 1 g substrate, P-solutions Calculated P-sorption capacity 20.0% Johansson (1999a) 8.3
(5–25 mg PO4 -P L−1 ) <20%, sorption
0–0.1 mg P g−1 .
Opoka (heated) Batch Artificial P solutions Calculated Maximum sorption 30.1 (42.1)% Brogowski and Renman (2004) 12.6
(K2 HPO4 ), distilled water capacity was 119.6 g
PO4 -P kg−1
Oyster shell Batch, column P solution 0–320 mg P L−1 . Langmuir Adsorption 37.80% Seo et al. (2005)
(burnt) Column, 100–g medium (1 833.3 mg kg−1 , column
month), 50 mg P L−1 , HRT (unsaturated media)
4 h, HLR 300 mL d−1 7925 mg kg−1
material from Korea
Peat Lab-scale biofilter Domestic wastewater (1 Calculated P removal in lab-scale – Talbot et al. (1996)
month) lab-scale filters filters 44%. Biofilter:
300, 600, 1200 L m−2 d−1 , 12% for TP
household biofilter
70–258 L m−2 d−1

73
74
Table 1 (Continued).

Material Study type Description of the study Retention calculation P retention Ca (CaO) References pH

Peat Small-scale CW in field Filters treated landfill Calculated Reduction of TP during 5.1% Kõiv et al. (2009a) 7.2
leachate from activated the first 6 months 77%
sludge treatment plant (12 from sludge water, 93%
months) and from biopond from biopond water
(6 months), volume
0.11 m3 , loaded 36–41 mm
d−1 and 82 mm d−1
Peat Pilot-scaleCW VF peat filters and HF Determined Removal in VF peat 5.1% Kõiv et al. (2009b) 7.2
filters treated municipal filters was 58% and
wastewater and landfill 63%, P binding capacity
leachate, VF 0.86 m3 , 0.081 g kg−1
76 mm d−1 and HF 1.24 m3 ,
53 mm d−1 (HLR 60 L d−1 ),
6 months
Polonite Column Columns: 20 ◦ C, Calculated Retained P 96.7% 24.5% Gustafsson et al. (2008) pH decreased from 11.7 to 9.5
water-saturated

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


conditions. PO4 -P and
NH4 -N (5.30 mg dm3 ), tap
water, HLR 530 L m−2 d−1 ,
68 weeks
Sand Column Columns with upflow Calculated P removal at low load 1.2% Farahbakhshazad and Morrison (2003)
hydraulic regime, load (0.5–2.0 mg L−1 )
0.5–2.0 mg L−1 , filled with increased from 40% to
sand; HRT 2 h; HLR 75% with 12
7.2–86.4 L d−1 recirculations
Sand Full-scale CW HSSF sand filter, 8 years; Determined P in soil after 8 years 4.15% Vohla et al. (2007) Outflow, pH 7.4
HLR 3.2–5.5 m3 d−1 0.117 g P kg−1 , 72%
Sand, dolomite Batch, column Sorption experiment: 1 g Langmuir, determined Sorption 0.417 g P kg−1 , 5.1–7.5% Pant et al. (2001) 7.9–9.1
sand and shale 20 mL (0–100 mg P L−1 ) columns 0.7–1.0 g P
24 h. Column: SSF study at kg−1 ; 95%. Sand is
loading of P 330 mg m−2 sensitive to P loading
d−1 . Loading: 105 mg P m−2
d−1 , HLR 60–120 L m−2 d−1 ,
33 months
Sorption 0.303 g P kg−1 , 26.4–26.6% 9.3–9.4
columns 0.2 g P kg−1 ;
44% removal
Sorption 0.192 g P kg−1 , 3–10.5% 8.1–8.9
columns 0.7–0.8 g P
kg−1 ; 37% removal
Sands (Danish) Batch, column Solution 320 mg P L−1 by Calculated, Langmuir, determined Calculated removal 0.02–7% Arias et al. (2001) 8.2–8.7
calculated retention and 0.272–3.941 g P kg−1 ;
0–10 mg P L−1 by Langmuir 0.02–0.129 g
Langmuir. Columns 12 P kg−1 ; determined
weeks; HLR 0.24–L d−1 ; 0.052–0.165 g P kg−1 .
HRT 12–14 h
Sands (South- Batch 3 g sorbate and 75 mL Calculated P removal between 8.6 0.33–11% Vohla et al. (2005) –
Estonian) solution (5–1000 mg PO4 and 27.2% or 2.45 g P
L−1 ), 24 h kg−1
Full-scale CW HSSF sand filter, 5 years; Purification efficiency 4.15% Outflow, pH 7.4
HLR 1.0–6.3 m3 d−1 78.4%. P in sand
52.8 mg kg−1
Shellsand Batch 3 g in 75 mL, 0–1000 mg P Calculated First batch 14–17 g P 28–30% Roseth (2000)
L−1 in 24 h, and also 3 g in kg−1 , second batch
5–1000 mg P L−1 in 48 h, 3–4 g P kg−1
Norwegian sand
Column Column, HLR 10 cm d−1 ; 32 Determined Adsorption capacity
d 3.5 g P kg−1
Shellsand Batch 5–1500 mg L−1 , 24 h Langmuir 0.8–8 g P kg−1 31.8% Søvik and Kløve (2005) 8.6
Meso-scale CW in field HSSF filter in greenhouse Determined 335 g P kg−1 , saturated
for household, 1 year, before 2 years, Ca-P
Norwegian sand, HLR (37–57%), Al-P (8–23%)
85 mm d−1
Shellsand Batch 3 g material for 24 h in Langmuir 9.6 g P kg−1 0.033% Àdàm et al. (2007) 8.8

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


90 mL solution (0–480 ppm
PO4 3− )
Column Vertical upflow columns Calculated 0.497 g P kg−1 , P
(330 d), HRT 4.1 d, HLR removal 97%
4.5 L d−1
Soils (sub-, top- Batch 3 g and 30 mL solution Langmuir Adsorption 0.006–0.3% Sakadevan and Bavor (1998) 4.2–7.2
and wetland (500–10,000 mg P L−1 ), 0.001–0.005 g P kg−1 .
soils) shaken for 48 h
Spodosol Batch Filter medium treated with Calculated P-removal 1 mg g−1 (1.5) Johansson (1999a) 6.3
artificial P solutions when P load 15 mg L−1 ,
(5–20 mg PO4 –P L−1 ). decline in removal
Sorption isotherm studies with higher load
Wollastonite Batch 5 g adsorbent and 75 mL Calculated PO4 -P reduction 32.9% Hedström (2006)
solution varying from 0.8 90–93%. P sorbed
to 1700 mg PO4 3− L−1 , for 0.1–12,000 mg kg−1
20 h
Wollastonite Column At 20 ◦ C, under Calculated 51.1% of P was 15.1% Gustafsson et al. (2008) 8.0
water-saturated removed
conditions. PO4 -P and
NH4 -N (5 and 30 mg dm−3 ),
tap water, HLR 610 L m−2
d−1 , 68 weeks
Wollastonite Column Vertical upflow columns Calculated >80% removal (up to – Brooks et al. (2000)
with HRT from 15 to 180 h, 96%), when the
secondary wastewater, 411 residence time was
d >40 h
Wollastonite Full-scale CW SSF wetland cell, Calculated Soluble P retention – Hill et al. (2000)
tailings wastewater from dairy 12.8%. Mean reduction
farm, HRT 13 d, 1.5 years was 27.5%.
Zeolite Batch 3 g and 30 mL solution Langmuir Adsorption 2.15 g P – Sakadevan and Bavor (1998)
500–10,000 mg P L−1 , 48 h kg−1

75
76
Table 2
Industrial by-products used as filter media for P removal.

Material Study type Description of the Retention calculation P retention Ca (CaO)% References Particle size (mm) pH
study

Coal ash Batch, column (2.5–40 mg P L−1 , 20 g Langmuir Sorption for fly ash – Drizo et al. (1999) 8.2
and 24 h) and column 0.862 g P kg−1 . Column
study, furnace bottom 300 mg P kg−1
ash (40 days); HLR 3 L
d−1
Coal ash Batch Bottom ashes: 5–40 mg Langmuir 0.081–29.5 g P kg−1 2.4–11.7 (5–16.5) Yan et al. (2007) 0.1–2 9.5–11.6
HPO4 2− L−1 , fly ashes
200–1500 mg L−1 , 72 h
Fly ash Batch 5 g adsorbate and Langmuir 3.08 g P kg−1 2.6 (0.41) Cheug and 7.7
(lagoon), 25 mL solution with Venkitachalam (2000)
precipitator fly different P
ash and sand concentration, shaken
24 h

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


13.77 g P kg−1 5.0 (4.3) 12.4
0.04 g P kg−1 0.2 (0.01) 6.4

Fly ash (acidic) Batch, column Batch equilibration Calculated PO4 3− immobilization – Grubb et al. (2000) 4.5
experiments using a 75–100%. Column:
low Ca 1% as CaO and removed 10 mg L−1
50–100 mg P L−1 , 24 h. over 85 pv
A column, 10 mg L−1
Initial pH 8.5
−1
Fly ash Batch Laboratory P Langmuir, Freundlich 160–420 mg P g – Mann and Bavor (1993) 0.08–4.75 slag 1
adsorption study
260 mg P g−1 9.5–19 slag 2
<0.5 Fly Ash

Fly ash (Red Batch 20 mL of 155 mg L−1 Langmuir 113.9–345.5 g P kg−1 32–33 (45.2–46) Li et al. (2006) 10.1–11.9
mud) KH2 PO4 , materials
heated at various T
8.9–78.4 g P kg−1 1.5–1.9 (2.1–2.7) 8.3–10.0
−1
Fly ash Batch 25–1000 mg P L , at Langmuir Retention: from 5.5 to 1.8–18 (0.07–5.6) Chen et al. (2007) 5
adjusted pH 5, 40 mL 42.6 g P kg−1
PO4 2− solution;
reaction time 24 h, fly
ash from coal power
plant, China
Fly ash Full-scale CW CWs for the treatment Calculated Majority of the TP was – He et al. (2007)
of eutrophic river adsorbed by fly ash
water, 3 stage system, stage, TP removal
one filled with fly ash, about 83%
HLR 60 m3 d−1 ; 14
weeks
Iron ore 67% (aerated) and 53% 0.04
(anaerobic), working
area 140 ␮g P g−1
Quartz sand 70% (aerated) and 43% 0.01
(anaerobic)
Ochre Batch P solutions (1500 and Langmuir 0.026 g P kg−1 . 90% of 7 Heal et al. (2003, 2005) 7.2
3000 mg P L−1 ) were all P forms were
added until P was no removed after 5 and
longer removed from 15 min shaking
the solution. Between
additions stand 48 h,
dried. high saturated
hydraulic conductivity
26–32 m d−1
Oil shale Batch 2.5–40 mg P L−1 , 20 g, Langmuir Adsorption 0.582 g P – Drizo et al. (1999)
(burnt) 24 h; HLR 3 L d−1 kg−1
Sediment of oil Batch Batch experiments, Calculated 25 g PO4 3− kg−1 or 8.2 g 20.9 (29.2) Vohla et al. (2005) <0.063 12.32
shale ash 5–1000 mg PO4 3− L−1 , P kg−1
24 h
Pilot-scale CW Experimental filter in First 4 months removal 20.9 (29.2) Vohla et al. (2005) <0.15 Outflow pH 7.6

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


saturated conditions, 71%, after average 52%
tertiary treatment for
P, HRT 1.5–2 d−1 , 1 year
Sediment of oil Batch Batch experiments, Calculated P-binding capacity up Kaasik et al. (2008) 1–2 and 0.01 (weathered) 12.32
shale ash 5–300 mg PO4 3− L−1 , to 65 mg P g−1 ,
48 h, 9 d (67–85%). P sorption
due to Ca-minerals
(ettringite, portlandite)
Slags
Basic oxygen Column Synthetic aqueous Calculated Average PO4 -P removal 25.45 Kim et al. (2006) Adjusted pH 6–9
furnace slag solutions of P and Ca 90.4%
ions 40–80 mg L−1
Basic oxygen Column Soil, the mixture of Calculated After 1 day 100% 22.97 Cha et al. (2007) 1–2 11
furnace slag Korean soil from river
and slag were studied,
unsaturated columns
for 4 day 18 mL h−1 and
3 d brake, saturated
8–9.8 mL cm−2 d−1
Natural soil Unsaturated columns: 0.57
after 1 day 86%, after
4th day 58.6, saturated
columns 97.8%
Blast furnace Batch 0–320 mg L−1 , raw Langmuir P-sorption capacity up 23.97 Korkusuz et al. (2007) <3 7.7
slag domestic wastewater, to 9150 mg P kg−1
HLR 100 mm d−1 ,
10 mg P L−1 , 8 g
Blast furnace Column Water-saturated, Calculated WCBFS removed >95% 21.60 Gustafsson et al. (2008) 0–4 Effluent pH decreased from 11.4 to 8.8
slag PO4 -P and NH4 -N (5 P first 300 pvs, after P
and 30 mg dm−3 ), tap retention 85.6%
water, HRT 340 L m−2
day−1 , 68 weeks. BFS
from steel industry

77
78
Table 2 (Continued)

Material Study type Description of the Retention calculation P retention Ca (CaO)% References Particle size (mm) pH
study

Blast furnace Small-scale CW Small-scale SSF CW, Calculated Up to 72% of P was 27.6 Grüneberg and Kern (2001) 0.3–1
slag dairy farm wastewater, retained, working area
planted, 9 L, load 200 ␮g P g−1
15 L m−2 d−1 ; hydraulic
conductivity: quartz
sand 1.4 × 10−3 m s−1 ,
slag 5 × 10−3 m s−1 ,
iron-ore 2.5 × 10−3 s−1 ;
4 months
Blast furnace Full-scale CW The CW system with Calculated Slag filters reduced TP – Cameron et al. (2003) 0.45–2.6 Up to 11.52
slag slag filters in Canada. up to 99%
Slag filters (trickling),
ca. 60 m3 day−1 , 7
months
Blast furnace Batch 180 ppm PO4 3− , 50 mL Freundlich 6.37 mg g−1 (500 ppm), 26.93 (37.7) Oguz (2004) 0.02–0.03
slag of solution, at 25–65◦C, during 60 min,
(Iskenderun) different pH values. Jar phosphate removal

C. Vohla et al. / Ecological Engineering 37 (2011) 70–89


tests: removal time of 99%
PO4 3− (150–500 ppm,
BFS 60 g L−1 ). Slag from
steel factory
Blast furnace Full-scale CW VSSF reed bed, Determined Average removal of TP 24.30 Korkusuz et al. (2005) 7.11–8.56
slag from granulated BFS, HRT 45%.
Turkey 3 m3 d−1 , HLR mm d−1 ,
11 months, slag from
iron and steel company
Electric arc Full-scale CW Detention time 3 day, Determined Slag maintained its 10.48 Shilton et al. (2005) N.A.
furnace loading 2000 m3 day−1 , maximum removal,
(melter) slag 1.5 years, slag from reaching TP retention
Steel Mill. of 1.23 kg T−1
Electric arc Column Columns, 20 mg P L−1 Determined Slag was nearly 100% – Drizo et al. (2006) 2.5–10 10.6–11
furnace slag (114 d) than 400 mg P efficient, removed
L−1 (21 d); HLR 2.2 mg P g−1
3.0–4.0 L d−1 ; HRTv
12–24 h
Fly ash Blast Batch Adsorbent (from Fumkin 32 mgPO4 g−1 2.9 Agyei et al. (2002) 0.026
furnace slag S-Africa) 0.5–5 g were
Portland used in 200 mL 100 mg
cement PO4 3− L−1 , pH 9 for 16 h
60 mg PO4 g−1 23.8 0.031
83 mg PO4 g−1 45.2

Soils, Sands, Batch KH2 PO4 at P Langmuir 0.29–0.97 g P kg−1 Xu et al. (2006) 4.4–4.9
Fly ash, Blast concentrations (pH 7.0)
furnace slag of 10–100 mg P L−1 for
sands and soils from
China, and of
100–1000 mg P L−1 for
fly ash and furnace slag
(24 h)
0.13–0.26 g P kg−1 3.3–5.6 6.7–8.1
8.89 g P kg−1 , 100% 12.3
removal
8.81 g P kg−1 , 100% 22.4 (3.1) N.A. 12.1
removal
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 79

Table 3
Man-made products used as filter media for P removal.

Material Study type Description of the Retention P retention Ca% References pH


study calculation

Filtra P Column At room Calculated Retained P 98.2%, 3.10 Gustafsson et al. (2008)
temperature under effluent pH fell
water-saturated from 12.9 to 11.6,
conditions. PO4 -P clogged after 971
and NH4 -N (5 and pv
30 mg dm3 ), tap
water, HLR
7–10 L m−2 day−1 ,
68 weeks
Filtralite PTM Small-, meso- and Small-scale Determined Small-scale: – Àdàm et al. (2006) >10
full-scale CW systems (953 g): extracted P
HLR 5.0 and 1.25 L 3887 mg P kg−1 .
d−1 , HRT 0.21 and Meso-scale:
0.83 d. Meso-scale 4500 mg P kg−1 .
filter (359 kg): HLR Full-scale: 52 mg P
∼71 L d−1 , HRT kg−1
3.6–4.3 d. Full-scale
system: VF and HF
filter (180 m3 ); HLR
1.6 m3 d−1 ; HRT 18
d
Filtralite PTM Full-scale CW The upflow filter, Calculated Reduction in the – Heistad et al. (2006) 12.5 (after 3 years 10)
6 m3 , HLR filter system is
450–864 L d−1 ; 3 99.4%
years
Filtralite PTM Batch 3 g, 24 h in 90 mL Calculated 2.5 g P kg−1 0.003 Àdàm et al. (2007) 10.7
solution with
phosphates
0–480 ppm
Column 2 vertical upflow Calculated and By mass balance
columns up to 330 extracted calculations
d, HRT 3.65 d; HLR 497 mg P kg−1 , by
4.9–5.5 L d−1 extraction up to
635 mg P kg−1 ,
removal 54%
Leca (Estonian) Batch 3 g and 75 mL Calculated Bounding capacity 0.1 Vohla et al. (2005)
solution with 0–10 mg P g−1 .
5–1000 mg PO4 Sorption of up to
L−1 , 24 h 7.98 mg P g−1
Leca (Estonian) Full-scale CW VSSF + HSSF filter Calculated TP removal 89% 2.1 Öövel et al. (2007) Outflow pH 7.4
bed, 64 PE, HRT
7.4 m3 d−1 , 2 years
LWA Batch, Full-scale HF CW, HRT 4 Calculated >95% 3.1 Zhu et al. (2003) 10
CW weeks, households,
4 years, Norsk Leca
LWA Batch Batch: 3 g in 75 mL, Calculated Sorption 12 g P Jenssen and Krogstad (2003) 12.3
(lab-made) 320–480 mg L−1 , kg−1 , higher
24 h. Laboratory dolomite content
made LWA and pH guarantees
sorption
LWA (USA) Batch Test Calculated 3465 mg P kg−1 46.5 Zhu et al. (1997) 10.1
(0–320 mg L−1 ), 8 g
substrate, 200 mL
solution, 24 h
Filtralite TM 209–2210 mg P 31.0 10.3
kg−1
Leca (Swedish) 46–565–mg P kg−1 8.5 10.5
Norlite Full-scale CW SSF wetland cell, Calculated Norlite removed 1.43 Hill et al. (2000) Outflow pH 7.7
wastewater from 34%
dairy farm

P L−1 ) but the long-term stability remains to be established (Bellier was 16.1–67.3%, and 48.3–97.9% in the case of activated bauxite.
et al., 2006). Batch tests with bauxite (Altundoğan and Tümen, 2002) showed
that orthophosphate removal was higher at pH 4.5 (67.3% of 10 mg
P L−1 initial solution).
3.1.1.2. Bauxite. Bauxite, a naturally occurring mixture of miner-
als rich in hydrated Al oxides and Fe oxides and low in alkali 3.1.1.3. Dolomite. Dolomite (CaMg(CO3 )2 ) is a common sedimen-
metals, alkaline earths and silicates, obtained from Alcan Chem- tary rock-forming mineral that can be found in massive beds several
icals Europe, Burntisland. Phosphate removal efficiency of more hundred feet thick. These are found throughout the world and are
than 95% for activated bauxites has been reported (Altundoğan quite common in sedimentary rock sequences. These rocks are
and Tümen, 2003). Adsorption of different P forms by raw bauxite appropriately enough referred to as dolomite or dolomitic lime-
80 C. Vohla et al. / Ecological Engineering 37 (2011) 70–89

stone. The P removal efficiency of dolomite was 44% of the influent (0.2%), sodium (0.17%), iron (2500 ppm), aluminium (500 ppm) and
P (Pant et al., 2001). Karaca et al. (2004) showed that the P adsorp- copper (15 ppm) (Inglethorpe, 1992). The removal of phosphorus
tion capacity for dolomite fluctuated from 9.7 to 52.9 g P kg−1 , with in the experiments conducted by Gray et al. (2000) was one of the
an increase in initial phosphate concentration from 10 to 60 mg L−1 . highest found in the literature. This study noted that maerl had
The amount of phosphate adsorbed also increased with the increase great potential as a constructed wetland substrate, due to its high
in pH from 1 to 11. phosphorus-adsorbing capacity (0.12–39 g P kg−1 , according to the
initial P concentration in solution: 5–5000 mg P L−1 ).
3.1.1.4. Gravels. In the gravel systems, the main P-removal mech-
anisms are adsorption and precipitation reactions with Ca, Al, and 3.1.1.8. Marble. Marble is a nonfoliated metamorphic rock result-
Fe but also, like in all media, biological assimilation and plant ing from the metamorphism of limestone, composed mostly of
uptake can play a remarkable role (Korkusuz et al., 2005). Tanner calcite (a crystalline form of calcium carbonate, CaCO3 ). Gervin
et al. (1999) showed that over 5 years, P accumulation decreased and Brix (2001) tested a planted vertical flow wetland system to
and substratum P-sorption capacity became saturated. Mann and reduce the phosphorus discharge from combined sewer overflows.
Bavor (1993) described testing the phosphorus removal efficiency The bed medium consists of a mixture of calcareous gravel and
gravel based constructed wetland system in a 2-year study in which crushed marble, which has a high binding capacity for phosphorus
removal ranged from −40% to 40%. Korkusuz et al. (2005) showed (up to 25 kg P m−3 ; Brix et al., 2001). During the initial two years of
that many substrates in subsurface constructed wetlands (e.g. pea operation. P removal was consistently high (94–99% of 1.3–1.8 mg
gravel, crushed stones, sand, etc.) usually do not contain high con- P L−1 inflow concentrations). When loaded with lake water, the
centrations of Ca, Al, Mg and Fe, and thus the removal of phosphate phosphorus removal was high during summer (71–97%) and lower
is, generally, low and varies widely among systems due to the during winter (53–75%), suggested to have been partly because of
different materials used. For domestic wastewater treatment in lower inlet concentrations.
Turkey, vertical subsurface flow constructed wetlands showed TP
removal efficiency for gravel wetland cells of only 4%. He et al. 3.1.1.9. Sands. In sand or gravel substrates, phosphorus is bound
(2007) reported TP removal during a 14-weeks experiment in a to the media mainly as a consequence of adsorption and precipita-
gravel wetland from 6.8% to 54%, on average 22.44%. In pilot-scale tion reactions with calcium (Ca), aluminium (Al) and iron (Fe). At
units containing medium gravel obtained from a quarry, Akratos pH levels greater than 6, the reactions are a combination of physical
and Tsihrintzis (2007) found greater removal efficiency for fine adsorption to iron and aluminium oxides and precipitation as spar-
gravel (89%), followed by medium gravel with cattail (67%) and ingly soluble calcium phosphates. At lower pH levels, precipitation
cobbles (57%). The removal efficiencies of the other two units as iron and aluminium phosphates (strengite, variscite) becomes
were significantly lower (medium gravel with reed 28.2%; medium increasingly important (Gerritse, 1993). The capacity of filter media
gravel alone 43.9%). PO4 3− and TP removal efficiency was predom- to remove P may therefore be dependent on the contents of these
inantly affected by porous media size and type. minerals in the substrate. This hypothesis is supported by the
observation that P-removal has been found to be particularly effi-
3.1.1.5. Laterite. Laterite is a hydrated mixture of Al, Fe and Ti cient in constructed reed beds containing ferruginous sand (Netter,
derived from the decomposition of aluminium-rich silicate, baux- 1992). However, P-removal efficiency is often high initially and
ite rocks and subsequent loss of alkalis, lime, magnesia and silica. then decreases after some time as the P-sorption capacity of the
Wood and McAtamney (1996) has found 99% removal of phos- sand is exhausted (Ciupa, 1996). Arias et al. (2001) found that the
phates from leachate in lab and 96% in pilot CW for laterite. P-removal capacity of some sands would be used up after only a few
Probably P ions are chemically adsorbed onto surfaces of hydrous months in full-scale systems, whereas that of others would persist
oxides of Fe and Al by ligand exchange. upnto several years. The most important characteristic of the sands
that determined their P-removal capacity was their Ca content. In
3.1.1.6. Limestone. Limestone is a very common sedimentary rock situations where the wastewater to be treated is more acid, the
of biochemical origin. It is composed mostly of the mineral calcite contents of Fe and Al may be more important, as the precipitation
(CaCO3 ). Sometimes it is almost pure calcite, but most limestones reactions with these ions are favoured at lower pH levels (Stumm
are filled with lots of other minerals and sand, and they are termed and Morgan, 1981). Pant et al. (2001) showed that some local sands
“dirty limestones”. Calcite is derived mostly from the remains of from Canada with elevated contents of Fe, Al and P have high P-
organisms such as clams, brachiopods, bryozoa, crinoids and corals. sorption capacity. Sand filters are also known as efficient units for
Some limestones may have been derived from non-biogenic calcite complex wastewater purification (BOD, COD, NH4 -N, and also in
formation. Although some limestones can be nearly pure calcite, some cases PO4 -P and faecal coliforms). Long-term purification,
there is often a large amount of sand or silt that is included in the however, has only been demonstrated in a few studies (Mander
shelly debris. DeBusk et al. (2004) showed that in mesocosm-scale et al., 2003; Vohla et al., 2007).
experimental wetland system, the TP removal for the limerock filter
system was 46%. Strang and Wareham (2006) described a removal 3.1.1.10. Opoka. Siliceous sedimentary rock Opoka was deposited
of only 20% in limestone filters. Johansson (1999a) found a retention in the Mesozoic era. In contrast to limestone that is crystalline,
capacity from 0.3 up to 20 g P kg−1 for limestone. Opoka is amorphous in its structure, making it highly porous. This
material, from south-eastern Poland, is composed of 50% CaCO3 ,
3.1.1.7. Maerl. Maerl are the dead deposits of calcareous red algae 40% SiO2 and 10% Al-, Fe- and other oxides. Opoka samples in
(Corallinaceae) found growing in shallow waters around the coast the study performed by Johansson and Gustafsson (2000) did not
of north-western Europe and the western Mediterranean. Upon retain phosphorus efficiently, although these extracts were super-
death, the algae accumulate to form large shoals, providing an saturated with respect to hydroxyapatite, and supersaturation was
important habitat for benthic communities (Hall-Spencer, 1998). probably too low to initiate Ca-P precipitation. Renman et al. (2004)
The use of maerl exploits its high mineral content; in addition describes the high efficiency of heated Opoka.
to high percentages of calcium carbonate (80%) and magnesium
carbonate (10%), it contains small amounts of 25 different trace 3.1.1.11. Peat. Compared to mineral soils, peat has a high organic
elements such as sulphur (0.6%), phosphorus (0.35%), potassium matter content and large surface area (more then 200 m2 /g), and
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 81

is highly porous (80–90%; Brown et al., 2000). The reduction of P Loamy soils were probably from the E horizon and extended into
in mineralized peat may occur through the sorption, sedimenta- the argillic (top of the Bt) horizon. The E horizon refers to depleted
tion and combination of complex compounds (Mann, 1990; Kadlec clay and Fe oxides; the parental material is rich in calcium, except
and Knight, 1996; Richardson et al., 1996). Alternatively, some for the top meter, due to leaching. Problems with soil SSF wetlands
P may be bound to the biofilm (Mann, 1990; Richardson et al., that might appear: declining hydraulic conductivity leads to lower
1996), but this can be quite irreversible such as the phosphorus retention time; during subfreezing temperatures unpredictable; to
can easily be transformed from organic to inorganic form. Impor- avoid clogging, SS should be kept out of the media (wetland owner,
tantly, the phosphorus forms chemical complexes with organic farmer), and it is not usable in vertical flow beds. The spodic B
and inorganic ligands, which can be adsorbed into the soil or horizon of podzolised forest soil described by Johansson (1999a)
will be removed by precipitation (Mæhlum, 1998). Interestingly, is rich in Al and F and is effective in removing As, but also P
Kõiv et al. (2009b) results show that the phosphorus removal (85–92%), whereas P was mainly sorbed to the Al- and Fe- com-
by HF peat filters (well-mineralized peat) did not occur if the P pounds. Sakadevan and Bavor (1998) demonstrated that soils from
concentration was lower than ∼1.5 mg L−1 . The nature of this phe- experimental and former wetlands removed P 100% if the loading
nomenon remains unclear, but this may suggest that the phosphate remained under the 100 mg P L−1 . Phosphorus sorption by the soils
molecules need to be bound to complexes (through formation of was strongly correlated with oxalate extractable Al and Fe. Organic
larger molecular aggregates) to be adsorbed to peat surfaces. In matter content rather had a negative effect, probably due to the
aerobic wetland conditions, phosphorus appears in dissolved com- competition with P ions for the adsorption sites.
plexes together with Ca and Mg ions in alkaline conditions, and
with Fe or Al ions in soil with acidic to neutral pH (Nieminen 3.1.1.15. Wollastonite. This calcium metasilicate is a mineral
and Jarva, 1996; Mæhlum, 1999), which can then be chemically mined for several purposes (primarily in ceramics, friction prod-
adsorbed by peat particles. Talbot et al. (1996) showed in a lab- ucts, paint filler and metalmaking) in upstate New York. Calcium
scale experiment that the efficiency of P removal from initial 14 mg metasilicate deposits exist in China, India, Finland, California and
P L−1 was 44%. Average effluent parameters over a 2-year study for Mexico as well (Hare, 1993). It has been used as a substrate for
the septic tank and peat filter system demonstrated no removal of constructed wetland ecosystems for the removal of soluble phos-
PO4 -P and 12% for total P. Rizzuti et al. (2002) found that peats with phorus from secondary wastewater. Brooks et al. (2000) studied
higher N and P retention capacities also have humic acid contents vertical upflow columns with hydraulic residence times varying
between 5 and 7%. Bulc et al. (1997) demonstrated that phospho- from 0.6 to 7.5 d. Removal higher than 80% (up to 96%) was achieved
rus removal efficiency from leachate over a 5-year study was on when the residence time was >1.7 d. A relationship was established
average 63%. between residence time and soluble phosphorus removal. Hill et
al. (2000) tested wollastonite tailings, which were a by-product of
3.1.1.12. Shale. Shale, an argillaceous rock, is derived from the mining activities that produced wollastonite and garnet (a ferrous
lower limestone group of the carboniferous system, which is highly metasilicate). The tailings were approximately 15% wollastonite
fissile and readily splits into very thin laminae. It is a readily avail- and 70–80% garnet, so they provide both calcium and iron as poten-
able material in Scotland (Drizo et al., 1997; Drizo, 1998). Drizo tial adsorption sites for P. Lab experiments showed high (90%) P
et al. (1999) found that X-ray fluorescence measurements showed removal. A retention time at least 62 h is necessary for optimal
that substantial precipitation of P had occurred on the shale sur- removal (Brooks et al., 2000). Mean reduction in P concentrations
faces. Shale shows promise as a substrate for constructed wetland was 28%. Hedström (2006) studied wollastonite from China. Phos-
systems, using low loading rates. Results from batch and column phate removal was minor when the initial P concentration was low.
tests suggest it may have a potential lifetime of 20 years in such Thus this material may not be suitable for wastewater treatment,
systems; however, data from pilot-scale and full-scale CWS would due to its low removal capacity at low concentrations. For PO4 -P
be needed to confirm that the lifetime is of this order of magnitude. concentration as low as 0.8 mg L−1 , the sorption of phosphate was
negligible. The results also indicated that possible precipitates are
3.1.1.13. Shellsand. Shellsand is a natural carbonatic material hydroxyapatite and calcium phosphate.
mainly sourced from shells, snails and coral algae with a grain con-
sistently similar to sand or gravel. The quality varies and can be 3.1.1.16. Zeolite. Zeolite is a hydrated aluminium-silicate mineral
separated according to its origin and weathering (Erstad, 1982). in which the aluminium and silicon polyhedra are linked by the
One million tonnes are annually harvested on the Norwegian coast- sharing of oxygen atoms. Drizo et al. (1999) described that zeo-
line (Roseth, 2000). Shellsand has shown a maximum P-sorption lite achieved an adsorption maximum of 0.462 kg−1 . Chen et al.
capacity of 9.6 g P kg−1 . The find for shellsand is consistent with (2006) has found the maximum P retention for different zeolites
earlier findings by Roseth (2000) and Søvik and Kløve (2005): in at different pH ranges, defined by Langmuir adsorption, to be only
vertical upflow columns the mass balance calculations showed that 0.01–0.05 g P kg−1 .
the column sorbed an average of 0.5 g P kg−1 of material. Àdàm et al.
(2006) found the overall P-removal rate by shellsand of 92% (Àdàm 3.1.2. By-products that are used for phosphorus removal
et al., 2006). Roseth (2000) demonstrated calculated phosphorus 3.1.2.1. BauxsolTM . BauxsolTM , is a neutralised bauxite residues, i.e.
adsorption for shellsand 14–17 g P kg−1 and 3.5 g P kg−1 for the seawater-neutralized red mud, and consists in a complex mixture
column. Søvik and Kløve (2005) found retention capacity in accor- composed primarily of Fe and Al oxides from the Queensland alu-
dance with the Langmuir equation for shellsand up to 8 g P kg−1 . mina refinery in Australia. According to batch tests (Akhurst et
al., 2006), adsorption of phosphates increased with decreasing pH,
3.1.1.14. Soils. P retention by mineral soil, humus and peat with maximum adsorption capacities at pH 5.2 (the lowest pH
is related to the content of iron and aluminium compounds investigated), increased BauxsolTM to initial phosphate concentra-
(Hartikainen, 1982; Heikkinen et al., 1995; Nieminen and Jarva, tion ratios and increased time.
1996; Giesler et al., 2005). These Al and Fe compounds are abundant
in upland soils with podzolised B horizons, but are scarce in peat 3.1.2.2. Burnt oil shale (BOS). Burnt oil shale is widely available in
soils (Nieminen and Jarva, 1996). For instance, fine loamy soil beds central Scotland as a waste product that results from heating oil
removed the highest average amount of P, i.e. 53% (Hill et al., 2000.) shale to produce mineral oil. Data found about burned oil shale
82 C. Vohla et al. / Ecological Engineering 37 (2011) 70–89

by Drizo et al. (1999) showed the adsorption maximum for that the ash removal system, and the hydration processes continue in
material to be 0.582 g P kg−1 , on the basis of the Langmuir equation. open plateaus, resulting in different secondary Ca-minerals. Batch
experiments by Kaasik et al. (2008) indicated the good (up to 65 mg
3.1.2.3. Coal fly ashes. Coals are composed of aluminium silicate P g−1 ) P-binding capacity of the hydrated oil shale ash sediment,
clays, carbonates, sulphides, chlorides and quartz that are oxidized with a removal effectiveness of 67–85%. This is the highest reten-
at high temperatures (>1500 ◦ C), which melts almost all of the tion capacity, and is found by calculating the sorbed P amount from
inorganic components with the exception of quartz. Coal fly ash the difference between initial and final P concentrations per gram
is an inorganic waste product from coal combustion, consisting of material. The high phosphorus sorption potential of hydrated oil
mainly of spherical glassy particles of silica (SiO2 ), alumina (Al2 O3 ) shale ash is considered to be due to the high content of reactive
and iron oxides. Fly ashes are widely used on agricultural land to Ca-minerals, of which ettringite Ca6 Al2 (SO4 )3 (OH)12 × 26H2 O and
improve the physical and chemical properties of soil (Kukier et portlandite Ca(OH)2 are the most important (Kaasik et al., 2008).
al., 1994). In addition, fly ashes are used in effluent treatment to Treatment with a P-containing solution causes partial-to-complete
remove various pollutants (COD, suspended solids, organic com- dissolution of ettringite and portlandite, and precipitation of Ca-
ponents, chrome dye, heavy metals, phenolic compounds, fluoride carbonate and Ca-phosphate phases, which was confirmed by X-ray
and also P (Eye and Basu, 1970; Gray et al., 1988; Banerjee et al., diffraction (XRD) and scanning electron microscope (SEM)-EDS
1997; Gupta et al., 1990; Piekos and Paslawska, 1999; Kao et al., studies. Vohla et al. (2005) found retention capacity of 8.2 g P
2000; Diamadopoulos et al., 1993; Tsitouridou and Georgiou, 1988; kg−1 , which is a lower capacity value, probably due to a different
Gray and Schwab, 1993). methodology and calculation techniques. Hydrated oil shale ash
Chen et al. (2007) found that high Ca, medium Ca and low Ca fly has also been shown to have quite an effective removal capacity in
ashes from coal-using power plants in China showed retention at field studies, showing removal efficiency in hydraulically saturated
different pH ranges to be up to 42.6 g P kg−1 . experimental filter treating domestic wastewater up to 71% and in
Yan et al. (2007) described retention capacity as defined by small-scale vertical flow filters (Kõiv et al., 2009a) treating leachate
Langmuir as being up to 29.5 g P kg−1 , while the initial pH in slurry up to 68%.
was alkaline (9.7–11.6).
He et al. (2007) studied constructed wetlands for the treatment 3.1.2.7. Slag. Slag is a porous non-metallic co-product produced
of low-concentration polluted eutrophic landscape river water in the iron and steel industry. In steelmaking industry, materials
with a three-stage system filled with fly ash. The TP adsorbed by including iron ore, scrap metal, and fluxing agents such as lime, are
the last 7 m of the fly ash stage was about 83%. heated beyond their melting points. Two main liquids are formed
the metal in fusion and a liquid with a lowest specific gravity forms
3.1.2.4. Ochre. The flooding of abandoned mines frequently results a layer on the surface of the melt that is called slag (Proctor et al.,
in the formation of acidic, Fe rich water due to the oxidation of 2000). Four main slag are identified including blast furnace slag (BF-
pyrite during mining operations. In treatment plants, atmospheric slag, is associated with iron production), basic oxygen furnace slag
oxidation and precipitation are harnessed, and thus large quanti- (BOF-slag, produced in a furnace supplied with oxygen to remove
ties of Fe(OH)3 and FeO(OH) precipitate accumulates, collectively carbon), electric furnace slag (EAF-slag in steel production, pro-
known as ‘ochre’. Heal et al. (2003, 2005) demonstrated maximal duced in an electric arc furnace, a furnace adapted for scrap metal
sorption capacity for ochre of 0.026 g P kg−1 . It was found that 75% recycling) and melter slag (specific production in New Zealand
and 90% of all P forms were removed after 5 and 15 min shaking, from sands rich in iron). A large range of sorption capacities were
and pH differing from 4 to 10 did not influence P removal. Ochre reported from batch experiments using slag ranging from 0.1 up
was also tested in the field, but due to the hydraulic conditions, the to 50 g P kg−1 slag (Mann and Bavor, 1993; Sakadevan and Bavor,
results remained moderate (27% of P was removed from a initial 1998; Grüneberg and Kern, 2001). Apparently, the large scatter of
solution of 20 mg P L−1 ). results of phosphorus sorption on slag relates to the fact that these
substances are not fully and unambiguously defined by their chem-
3.1.2.5. Red mud. Red mud is the by-product associated to alumina ical, mineralogical and phase compositions. The problem has now
production from bauxite using the Bayer Process. It causes serious been studied for nearly 20 years (Yamada et al., 1986), and it is
environmental problems due to its high alkalinity and large quan- still evident that more detailed research is needed to elucidate the
tities. Environmental and economical concerns have led to ongoing sorption of phosphates. As P is known to bind with Ca at high pH,
research to find effective ways to utilize the abundant and widely the formation of calcium phosphates was believed to be a major
available red mud (Shannon and Verghese, 1976) to remove phos- mechanism of P removal by slag systems (Baker et al., 1998). Sim-
phate. Li et al. (2006) found a very high maximal sorption capacity ilarly, calcium-containing alkaline slag has proven to have a high
of 345.5 g P kg−1 of extra heated red mud in batch-scale experi- P-sorption capacity as shown earlier in batch, column experiments,
ments. Treated red mud also had alkaline pH and high Ca content. but also in some field investigations (Yamada et al., 1986; Mann and
For raw red mud, the maximum sorption capacity using Langmuir Bavor, 1993; Johansson, 1999b; Sakadevan and Bavor, 1998).
was 113.9 g P kg−1 . To determine if P-loaded slags might be efficient as agricultural
fertilizers, Johansson and Hylander (1998) showed, in pot experi-
3.1.2.6. Sediment of oil shale ash. Kerogenous oil shale used in Esto- ments with barley, that P retained by slag was available for uptake
nian thermal power plants is a solid fuel of low energetic value, by plants. However, further investigations are required to under-
which after combustion leaves large amounts of ash (45–48% of stand the uptake mechanisms. Cameron et al. (2003) reported that
dry mass of shale). Estonian oil shale is highly calcareous (average during research conducted on slag column filters for the removal
calcite and dolomite content is 40–60% of the mineral matter), and of PO4 from wastewater, the pH of the effluent of the slag columns
the ash remaining after combustion is due to the thermal decom- reached a maximum of 11.5. It is therefore recommended that
position (peak temperatures 1500 ◦ C) of carbonate minerals and the pH of the pilot-scale slag filters be measured and corrected
subsequent reactions with flue gases rich in free lime (CaO) and if required before discharge to surface waters. Drizo et al. (2006)
anhydrite (CaSO4). The ash is transported to waste heaps (plateaus) reported that electric arc furnace (EAF) steel slag was nearly 100%
through a pipe system in water slurry at an ash–water ratio of efficient due to specific P adsorption onto metal hydroxides and the
1:20. The lime and anhydrite already begin to react with water in precipitation of hydroxyapatite. The void hydraulic retention time
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 83

(HRTv) was a key factor for growing hydroxyapatite crystals and but is often criticized for the high energy demand that the pro-
had a significant effect on P-removal efficiency using EAF steel slag. duction processes consume, along with the high pH generated in
Shilton et al. (2006) presents a decade of experience of P removal effluents. An effective design could include a peat filter after the
by active slag filters at a full-scale treatment plant. The average filters with calcareous materials to reduce the high pH to meet the
total phosphorus (TP) removal efficiency of slag was 77% during permitted discharge values (pH 9.0) and even afford an additional
the 5-year period. P-removal effect (Kõiv et al., 2009b; Mayes et al., 2009). In Filtralite-
PTM systems, the initial pH can be up to 12.7 and a decrease to 11
3.1.3. Man-made products that are used for phosphorus removal is generally observed after several months. Other filter materials
3.1.3.1. Alunite. Alunite, KAl3 (SO4 )2 (OH)6 , is a mineral that in its (slags, fly ashes etc.) with alkaline components would also cause
original form is not soluble in water. It is formed by the hydrother- a high effluent pH. Possible solutions in addition to the polishing
mal alteration of tuff. It contains approximately 50% SiO2 . The filter are aeration and dilution (Saltnes and Føllesdal, 2005).
alunite used in the study revealed by Özacar (2006) was obtained
from Turkey. Alunite samples were calcined in a muffle furnace at 3.1.3.4. Lightweight aggregates (LWA) or light-expanded clay aggre-
1000 ◦ C for 30 min. Using a two-stage adsorber lab scale, an average gates (LECA). Lightweight aggregates or light-expanded clay aggre-
of over 80% removal was achieved from a contact time of 29.1 min gates have shown both good water permeability and phosphorus
at pH 5. sorption capability (Mæhlum et al., 1995; Johansson, 1997a; Zhu
et al., 1997; Drizo et al., 1999; Harris and Mæhlum, 2003; Jenssen
3.1.3.2. Filtra P. Filtra P is produced after heating a mixture of and Krogstad, 2003; Öövel et al., 2007). LWA is manufactured by
limestone, gypsum and Fe oxides. The product, which is strongly running palletized clay aggregates through a rotary kiln at 1200 ◦ C;
alkaline because of its content of Ca(OH)2 (about 20%), is gran- the chemical compositions differ according to whether their parent
ulated to a particle size of 2–13 mm for use as a filter material material is clay or shale. These filter substrates must be available
in wastewater treatment. Samples of Filtra P were obtained from in large quantities at low cost, but with long-lasting phospho-
Nordkalk Oy Abp (Pargas, Finland). Gustafsson et al. (2008) studied rus sorption capacity (Johansson, 1997b). Kvarnstrom et al. (2004)
Filtra P in subsurface wetlands cell treating dairy wastewater for 68 demonstrated that all inorganic P that was accumulated in the LWA
weeks. They found 98.2% removal of P. The pH of the effluent water was easily soluble, mobile, and available to plants. Mæhlum et al.
remained alkaline, dropping from 12.9 to 11.6. Nevertheless, the (1995) described the P removal in the landfill leachate system with
material clogged after 971 pore volumes. two parallel horizontal subsurface flow LECA filters during mon-
itoring, on average 88% a year. Johansson (1997a,b) conducted a
3.1.3.3. Filtralite PTM . Filtralite PTM , a new generation of column study with Leca and found that sorption of P by Leca was
Norwegian-produced LWA, has been especially developed for low (8–13%). Jenssen et al. (1991) reported 10 times higher sorp-
P sorption (Zhu et al., 2003; Àdàm et al., 2005). This new LWA tion in batch experiments, but the reason was probably that lime
is made of natural mineral clay illite with natural additives, and was added to the Norwegian Leca to avoid aggregation. Zhu et al.
is manufactured by running palletized clay aggregates through a (1997) tested 3 types of LWA and founded the US LWA to be most
rotary kiln at 1200 ◦ C. Filtralite PTM has high pH (10) and high Ca effective in comparison with Norwegian and Swedish LWAs. The
and Mg content (Jenssen and Krogstad, 2003). After saturation, US LWA also had the highest Ca content. Vohla et al. (2005) tested
this material can be used as an alternative fertilizer in agriculture. an Estonian LWA (Ca 0.1%), but found retention capacity of up to
Àdàm et al. (2005) showed in the study with Filtralite PTM at lab 3.2 g P kg−1 . Öövel et al. (2007) analysed the subsurface flow CW
scale that the removal rate was dependent on the inlet P concen- treating schoolhouse wastewater with an Estonian LWA for about
tration, but was not correlated with the hydraulic loading rate. It 2 years. The purification effect for total P was on average 89%.
was also noted that resting periods regenerated the P-sorption
capacity. 3.1.3.5. Norlite. Norlite is a lightweight coarse aggregate of fired
The retention capacity results from the batch study were quite shale. It is a construction material that is classified as a lightweight
consistent with those reported by Zhu (1998) (0.4–3.5 g P kg−1 ), aggregate. Norlite may be compared with LECA. It consists of 4.7%
although Zhu used the very first generation of Filtralite without iron oxide, 3.6% magnesium oxide, 3.2% alkalis, 2.0% calcium oxide,
dolomite addition. The founded P retention capacity was, how- 20.2 aluminium and 64.2% silica (Hill et al., 2000). Hill et al. (2000)
ever, much lower than that shown in earlier studies revealed by has shown that Norlite removed 33.7% of TP during the 1.5-year
Jenssen et al., 2003 (8–12 g P kg−1 ) for Filtralite P. Àdàm et al. study in a subsurface wetland cell treating dairy wastewater. Based
(2006) described the P-sorption capacity found in a batch study of on these data, Norlite is recommended for use in agricultural
Filtralite-PTM 8 g P kg−1 . Results from the small and meso-scale sys- systems. The author recommended this material, as it had good
tem showed that it was difficult to extrapolate the results obtained hydraulic conductivity and low risk of clogging when SS in water
from model studies to full-scale constructed wetland systems. was higher.
Jenssen and Krogstad (2003) analysed the horizontal subsurface
flow LWA bed and found still effluent P concentrations of less than 3.1.3.6. Oyster shell. Oyster shell, in Korea, a large quantity of oys-
0.5 mg L−1 after 6.5 years. The system was loaded with wastewater ter shell is generated as a by-product of the oyster culture on
having and average concentration of 11 mg L−1 . Heistad et al. (2006) the south coast. The amount of oyster shell was about 0.25 mil-
studied a wastewater treatment system for use in single houses lion tonnes/year as of late 2003, of which only 10% was treated
(aerobic biofilter and an upflow saturated filter) for 3 years. The for oyster spat attachment (2000 tonnes/year) and fertilizer (2000
system performed excellent total phosphorus removal through- tonnes/year). Kwon et al. (2004) found that raw material removed
out the experiment. The average reduction in the filter system was no P in plastic jars, stirred with a deionized P solution. Heated oys-
99.4%. Over 3 years, the pH value from the upflow filter, which is ter shell, however, removed up to 68%, and pyrolized oyster shell
problematic using such alkaline materials, dropped from 12.5 to up to 98% of P.
10.
Due to high P-removal capacity Filtralite-PTM has been reported 3.1.3.7. Polonite. Polonite is manufactured from siliceous sedimen-
to be an effective filter media used in CW systems (Jenssen and tary Opoka rock in Poland that is heated to 900 ◦ C. The material
Krogstad, 2003; Àdàm et al., 2005; 2006, 2007; Heistad et al., 2006), consists mainly of reactive lime and wollastonite phases, which
84 C. Vohla et al. / Ecological Engineering 37 (2011) 70–89

Fig. 1. Relationship of P retention capacity (g P kg−1 ), Ca content (%) and pH level


of various filter materials. Fig. 2. Relationship of P retention capacity (g P kg−1 ), CaO content (%), and pH level
of various filter materials.

contribute to the strongly alkaline conditions. The Ca content in


Polonite averages 24.5%. Gustafsson et al. (2008) conducted a col- load, P retention is also higher (Àdàm et al., 2005; Akhurst et
umn study with Polonite and found the removed amount of P to be al., 2006). In addition, the greater amount of sorbent can also
96.7%. During the 68-week study period, the pH remained alkaline. influence the higher sorption, especially when one uses natural
Effluent pH decreased from 11.7 to 9.5. materials like sands and gravels. That is due to the way the P is
bound. In natural materials the P is usually bound by adsorption
3.2. Phosphorus retention capacity related to pH level and Ca or sorption, and thus the greater the specific area of the material
and/or CaO content of filter media in which the P can be bound, the higher the retention.
2. Different chemical composition of material. The high Ca content
As calcium ions can form stable and insoluble products with does not always guarantee a rapid reaction between Ca and P.
phosphate, calcium-based materials are considered to be one of For instance, the reaction is many times easier if the Ca in the
the potential sorbents for phosphorus removal. Ca-bound P is also material is in the form of CaO instead of CaCO3 , which has much
more available for plants than Al- or Fe-bound P (Tisdale et al., 1993; lower dissolving ability. For instance, Kaasik et al. (2008) showed
Krogstad et al., 2005). Willadsen et al. (1990) and Johansson and that hydrated sediment of oil shale ash consists of both CaCO3
Hylander (1998) showed that substrates containing CaCO3 sorbed and Ca6 Al2 (SO4 )3 (OH)12 ·26H2 O (ettringite) and Ca(OH)2 (port-
less P than substrates with more aggressive CaO or Ca(OH)2 con- landite). After the sorption test, the calcite content remained the
tent. same, but the ettringite and portlandite content declined drasti-
The retention capacity of the analysed materials was from 0.001 cally or was even lost in the material. Brogowski and Renman
(soil, Sakadevan and Bavor, 1998) to 345.5 g P kg−1 (heated red (2004) demonstrated that after the heating of Opoka, which
mud, Li et al., 2006). Most of the inserted data remained between naturally contain Ca as CaCO3 , the sorption capacity increased
P retention capacities of 0–25 g P kg−1 . The Ca content ranged significantly, from 0.1 g P kg−1 (Johansson, 1999a) to 39.0 g P
from 0.03 (Filtralite-PTM , Àdàm et al., 2007) to 46.5% (USA LWA- kg−1 . After heating Opoka, the CaCO3 in the material was trans-
UTELITETM , Zhu et al., 1997) (average content 9.3%). The pH of the formed to CaO. The higher is the temperature (up to 1000 ◦ C), the
selected filter materials ranged from 3 (bentonite, Xu et al., 2006) higher is the CaO content and P retention capacity. Kwon et al.
to 12.6 (heated Opoka, Brogowski and Renman, 2004) (average pH (2004) showed that heating (pyrolizing) oyster shell above the
was 8). As shown by Fig. 1, pH did not influence retention capac- 650 ◦ C, the CaCO3 content fell and P-removal efficiency increased
ity. Based on the three-dimensional figures, we can see a slightly to 98% in the material. Untreated material showed no P removal.
increasing trend for P retention capacity with increasing Ca con-
tent over 15% (Fig. 1). We can also see that most of the materials In some cases the Ca content was probably too low, and the Ca
with high pH also have higher Ca content (>15%). However, the high species were not reactive compared to other elements such as Al
Ca content does not guarantee higher P retention capacity. Several and Fe. In addition, adjusted acidic pH could influence the sorption
reasons for that phenomenon can be mentioned: capacity, favouring P binding reactions between Fe and Al more
than with Ca. Thus the removal mechanism was related more to
1. Different study conditions: P load, substrate amount, reaction the reaction with Al and Fe (Grubb et al., 2000).
time, the method of calculating P retention (Langmuir or using Most of the P retention capacities used for the analyses
water data). Arias et al. (2001) showed that if one uses Langmuir, against Ca/CaO content remained between 0 and 50 g P kg−1
the retention capacity was lower than if retention was calculated (Figs. 1 and 2). The CaO content ranged from 0.01% (sand, Cheug
from the difference of P concentration before and after the test. and Venkitachalam, 2000) to 46.0% (acid-treated red mud, Li et al.,
Many authors have claimed that if one uses a higher initial P 2006) (average content 13%). The pH range remained the same as
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 85

Table 4 maximum P-sorption capacity of 9.6 g P kg−1 . The overall removal


Determination coefficients (R2 ) of Spearman Rank Correlation between the P reten-
capacity of P during the column study was 92%. That is similar to
tion capacity, pH level and Ca and CaO content of all of the filter material studied.
the P-removal efficiency found in the Kodijärve sand (Ca content
P retention Ca (%) CaO (%) 41.5 g kg−1 ) filter in the first years of operation.
P retention On the other hand, the studies with limestone, which consisted
Ca (%) 0.43 mainly of calcium carbonate (CaCO3 ) in the form of calcite, have
CaO (%) 0.51 –a not been as effective as expected. Strang and Wareham (2006)
pH 0.22 0.34 0.42
demonstrated only about 20% of the total reduction achieved in
a
The correlation between Ca and CaO was not counted. the limestone wetland system. Hence the chemical form of Ca in
the filter media is also important characteristic when choosing a
in previous Ca analyses, and it did not influence the P retention potential filter media for P removal. Opoka (bedrock material from
capacity. Fig. 2 shows a slightly increasing trend for the P reten- south-eastern Poland) also mainly contains CaCO3 (50%), but low
tion capacity with increasing CaO content over 15%. In these terms retention of P was detected in the material, even in batch tests
it is similar to Ca content analyses (Fig. 1). Likewise, most of the (20%). Low P-sorption capacity was probably due to the type of Ca
materials with high pH have higher CaO content (>15%) (Fig. 2). in the material. CaCO3 is less active than CaO (Johansson, 1999a).
The determination coefficient (R2 ) between the Ca content and Heating the Opoka (CaO is augmented in that process) markedly
P retention capacity was only 0.43. However, the correlation was increased P sorption (average removal 96–99%). Maximum sorp-
stronger between the CaO and P retention (R2 = 0.51) and between tion capacity for heated Opoka was detected to be 119.6 g PO4 -P
the CaO and pH (R2 = 0.42) (Table 4). kg−1 (Brogowski and Renman, 2004).
The highest P retention capacities for various filter materials Wollastonite tailings are also an example of Ca rich sorbent,
were detected in studies in which batch analyses were used. It is which consist of approximately 15% wollastonite and 70–80% gar-
evident that the highest numbers are described when industrial net, and provide both calcium and iron as potential adsorption
by-products such as blast furnace slag studied by Mann and Bavor sites for P. Lab experiments showed high (90%) P removal. In the
(1993) and heated red mud from the study by Li et al. (2006) are SSF wetland cells, the soluble P retention in wollastonite over 1.5
used (Table 5). P retention capacities for natural and man-made years averaged only 12.8 ± 9.5%. The mean reduction in P concen-
products are from ten to a hundred times smaller. We also included tration was 27.5 ± 30.8%. Calculations for the lifetime were not
heated Opoka as a natural product, regardless of the fact that Opoka useful because of oversized filter (Hill et al., 2000). Hence the siz-
was extra treated. ing and optimization of hydraulic conditions in the filter cell plays
The difference between the maximum retention capacity of an important role in removal effectiveness. For instance, Drizo et
industrial by-products and man-made products might arise from al. (2006) reported hydraulic retention time as a key factor for the
the different calculation of P-sorption capacity. For the above- growing of hydroxyapatite crystals, and thus had a significant effect
mentioned by-products, the Langmuir equation was used to on P removal by electric arc furnace steel slag.
determine the maximum retention capacity, but in the case of The mixing of materials which can improve over all performance
man-made products, the retention capacity was calculated from can also be effective, for instance, Gervin and Brix (2001) studied
the difference between the initial and final P concentration of the a vertical flow system to reduce P in combined sewer overflows
solution per gram. using a substrate of gravel mixed with crushed marble. Gravel has
The smallest P retention capacity was detected when using soils, good hydraulic conductivity, and marble has been shown to have
sands and fly ashes with low Ca content (0.001 and 0.02 g P kg−1 ). good P removal. During the initial two years of operation, phos-
The smallest values of P retention capacity for man-made prod- phorus removal for combined sewer overflows was consistently
ucts have been 10–100 times higher (0.42 and 0.57 g P kg−1 ) than high (94–99% of inflow concentrations). In addition to the mix-
natural and by-products (0.002 and 0.003 g P kg−1 ) (Table 5). One ing of materials, the addition of a reactive source has also been
reason for this may be the higher Ca, but also Mg, Fe and Al content used when processing special man-made products for P retention.
in the material. Lime addition during the process of the production of Leca can, for
Substrates such as pea gravel, crushed stones, sand, etc. in the instance, increase P removal about 10 times, which is the range
subsurface constructed wetlands do not usually contain high con- between the sorption capacities of Swedish and Norwegian Leca
centrations of Ca, Mg, Al and Fe, the removal of phosphate is (Johansson, 1997a,b; Jenssen et al., 1991).
generally low, and varies extensively among systems due to the In addition to chemical composition, which determines the
different materials used. Korkusuz et al. (2005) reported in the sorption capacity of the filter media, phosphorus availability in sat-
results of the 1-year study that average removal efficiencies for the urated media is also important. The material that is enriched with P
gravel SSF vertical flow wetland cells were only recorded as PO4 3− - can be used as a fertilizer, provided that P is available to the plants,
P (1%) and total phosphorus (4%). Studies with shellsand (Àdàm and the contents of toxic compounds and pathogens do not restrict
et al., 2007; Søvik and Kløve, 2005; Roseth, 2000), which con- such use. For example, slag originating from the steel industry can
tains 10 times more Ca (327.6 mg kg−1 ) than Filtralite P, reported a effectively sorb phosphorus, and in saturated slag the phospho-

Table 5
Maximum phosphorus retention capacities (two for each group) of natural materials, industrial by-products and man-made products.

Maximum P retention, g P kg−1 Ca% (CaO%) Material; reference

40 30.1 (42.1) Heated Opoka; Brogowski and Renman (2004)


Natural materials
5.2 21.7 Dolomite; Karaca et al. (2004)

420 – Blast furnace slag (calculated after Langmuir); Mann and Bavor (1993)
Industrial by-products
65 20.9 (29.2) Hydrated sediment of oil shale ash; Kaasik et al. (2008)

12 – Lab-made LWA; Jenssen and Krogstad (2003)


Man-made products
3.5 46.5 USA LWA; Zhu et al. (1997)

Only materials with reliable hydraulic conductivity and useful for constructed wetlands are shown.
86 C. Vohla et al. / Ecological Engineering 37 (2011) 70–89

rus is also available for plants. In LECA material without added


Ca, phosphorus is mainly bound with Al- and Fe-compounds. Al-
compounds and phosphorus formulate compounds are known to
be difficult for plants to obtain. Phosphates bound to Ca in hydroxyl-
apatite form (Hap, Ca10 (PO4 )6 (OH)2 ) are also not easily released
from the material in a short time, because of their low solubility
(Johansson, 1997a,b; Kvarnstrom et al., 2004). Ca-bound P has also
been found to take the form of tricalcium phosphate (Ca3 (PO4 )2 )
and dibasic calcium phosphate (CaHPO4 ), which are more solu-
ble forms (Kim et al., 2006). For instance, bound P in Filtralite P,
Filtra P and Polonite is found to be at least partly plant available
(Kvarnstrom et al., 2004; Hylander et al., 2006; Gustafsson et al.,
Fig. 3. Relationship of P retention capacity (g P kg−1 ) and hydraulic retention time
2008). Producing Filtralite and LWA products, however, is a process
(HLR; mm d−1 ) of various filter materials. Quadrates indicate full-scale and pilot-
with a huge energy demand, which is why the benefit of using LWA scale constructed wetlands, triangles: column experiments.
as substrate is doubtful (Drizo et al., 1999). Phosphorus-sorbing
materials with additional beneficial characteristics may be used.
For example, lime can be used to counteract acidification so as to tion occurs at medium HRT level (Song et al., 2002). In particular,
improve soil structure (Rex, 2000). Most of the filter materials that too long HRT induces chemical clogging of hydrated oil shale ash
have been used are characterised by high pH values (9–12), which and reduces its phosphates removal capacity (Liira et al., 2009).
first create an unfavourable environment for bacteria (Renman et Fig. 3 only illustrates a general relationship between the hydraulic
al., 2004), but secondly, effluent with pH >9 is also not allowed to be parameters and P retention. For each individual filter material the
discharged into water bodies. In that case, an extra filter decreasing optimal range can be different.
pH or extra methods such as dilution or aeration are needed.
3.5. Lifetime of substrates in treatment wetlands
3.3. Influence of material media size and composition on
P-removal processes As the materials by themselves and the methods for testing P-
removal capacity vary greatly, it is quite difficult to make conclusive
Large differences between P-sorption capacities were recorded lifetime calculations. Many authors have stated that when testing
according to the nature of the treated water (e.g. Table 2). Several materials, those that show very high phosphorus removal in the
columns investigations were stopped due to clogging. Clogging was laboratory do not give similarly good results in field systems (Arias
a big issue when using reactive media in filtration systems. This was et al., 2001; Drizo et al., 2002; Àdàm et al., 2006). There are a great
mainly associated to the use of fine media size, to biofilm devel- variety of reported results described for the same type of mate-
opment caused by large amount of TSS and COD in the supplied rial. Most of the studies that demonstrate the phosphorus retention
effluent, and in a minor extent, to CO2 capture and precipita- capacity of various materials have been conducted in the labora-
tion under CaCO3 inside media, and to the flow rate. Since the tory. Only a few materials out of the extensive group that has been
main mechanism of P sorption on reactive material is related to tested so far have also been studied in full-scale systems.
CaO dissolution followed by P precipitation and crystallisation it It has been stated only the comparison between various mate-
is likely that the smaller the size of the material, the greater the rials and relative potential as a filter media can be made on the
surface available for CaO dissolution. This could lead to increasing basis of batch experiments (Drizo et al., 1999; Arias et al., 2001,
pH and probably to increase CaCO3 precipitation and crystallisa- 2003). Since hydraulic conditions do play an important role in
tion and finally cementation of the system. To overcome clogging removal processes and their effectiveness, at least column studies
several solutions have been proposed including the use of pre- are needed to predict possible removal capacity in the real system.
filter to remove soluble CO2 and organic carbon (Chazarenc et al., Àdàm et al. (2006) showed that even using smaller imitating filters,
2007), and the use of larger sized media (more than 5 mm medium it is difficult to extrapolate the results to full-scale systems.
diameter, e.g. Shilton et al., 2006). One other major issue was the If one uses maximal P retention capacity calculated from the
high pH recorded at the outlet especially at the beginning of the Langmuir equation or from final concentrations, it is easy to make
experiments. In many studies pH was greater than 9 and more- overestimations. P retention capacity is influenced by study dura-
over, in nearly one third of the studies, a pH decrease at outlet tion, water/material ratio and temperature, the range of initial P
was observed along with a decrease in P-removal efficiency. Thus concentrations, the chemicals used, rotation speed and many other
reducing outlet pH represents a major challenge but doing so could differences in methods described in greater detail in Drizo et al.
affect P-removal efficiency so care must be taken when solving this (2002). For instance, the regeneration effect of removal capacity in
problem. filter media during the resting period (Faulkner and Richardson,
1989; Cooper and Findlater, 1990) can yield higher overall P reten-
3.4. Influence of hydraulic loading rate and retention time on tion capacity. If we use the resting periods in one study and not
P-removal processes in another, the longevity prediction based on these results can
differ considerably (Àdàm et al., 2006). Brooks et al. (2000) have
Hydraulic loading rate and hydrological retention time and their stated that seeding effects could be a possible explanation for the
relation to P retention were correctly registered only in a few lit- increment in the P-sorption capacities of the material after the rest-
erature sources (Tables 1–3). Due to the large variability of data, ing periods. The seeds of calcium phosphates increased adsorption
we could not find any significant correlation between the HRT and processes by enlarging the specific surface area of the material
P retention. However, analysis made on full-scale and pilot-scale or increased nucleation and precipitation of amorphous calcium
treatment filter systems and column experiments shows that there phosphates over the nucleated.
is probably an optimal level for HLR at which the P removal is the Several by-products are tested as reactive media (materials rich
highest (Fig. 3). This is similar to the findings from other studies in Ca, Mg, Al and Fe) for P removal in real CW systems. Pilot and
which suggest that for selected filter media the optimal P reten- full-scale systems with the most often studied by-product apart
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 87

from ashes – blast furnace slag – have shown removal rates of 98% tems confirm, that the P retention capacity of most filter materials
(Grüneberg and Kern, 2001) and 99% (Cameron et al., 2003). Shilton significantly decreases after a 5-year period of application.
et al. (2006) studied the full-scale active slag filters over 5 years, and Some new and possibly promising materials such as Filtralite,
reported an average TP removal efficiency of 77%. Opoka, and dehydrated oil shale ash with extremely high P-
At the beginning of the use of constructed wetlands for the sorption rates should be evaluated for efficiency of P removal over
removal of contaminants from wastewater, it was most common longer time. Also, well monitored long-term experiments in full-
for natural sands or gravels to be used as a filter media in beds. size systems are needed.
As is known, in the sand or gravel substrate, phosphorus is mainly The recyclability of saturated materials as fertilizer plays an
bound to the media mainly as a consequence of adsorption and important role in the further application of the technology of
precipitation reactions with calcium (Ca), aluminium (Al) and iron treatment wetlands. Therefore adequate experimental studies are
(Fe). The most commonly chosen filter media did not contain large needed to ensure the applicability of recycled filter media. Success-
amounts of these minerals, and thus the filter bed did not removed ful use of filter media also depends on their availability at the local
P for more than a few years. For instance, gravel-based systems in level and the possibility of harmful side-effects such as contami-
Richmond, Australia showed a decline in removal efficiency after nation of the environment due to the high concentration of heavy
only 1–2 years of operation (Mann and Bavor, 1993). Tanner et al. metals.
(1999) showed the saturation in the gravel-filled wetland system The follow-up research on filter materials for phosphorus reten-
that purified dairy farm wastewater during a 5-year period. Vohla tion should be focussed on their hydraulic parameters combined
et al. (2005, 2007) described the saturation of an HSSF sand fil- with the analysis how to avoid the possible clogging. Likewise,
ter after 5 years of operation. Kadlec and Knight (1996) showed more attention should be given to investigate obvious constraints
that initial P-removal rates from wetland systems in the U.S.A. are such as poor hydraulic conductivity, low surface area, or high con-
often in excess of 90%, but decline sharply after only 4–5 years tent of undesirable contaminants such as heavy metals.
of cumulative P addition. Peat has also shown good results (up to
58–93%) in P removal in small-scale vertical flow filters, but satu- Acknowledgements
ration appeared after 6 months of operation (Kõiv et al., 2009a, b).
Meso-scale CW with shellsand saturated before 2 years had passed This study has been supported by the Ministry of Education and
(Søvik and Kløve, 2005). Korkusuz et al. (2005) reported average P Science of Estonia (grants SF0182534s03 and SF0180127s08) and
removal of only 4.33% in the gravel bed during a 1-year monitoring Estonian Science Foundation grants 6083 and 7527.
period.
Of man-made products, LECA systems are known to have an
extremely high P-sorption capacity (4 kg m−3 ), and have indicated References
promising P removal (90%) after 2–3 years of operation (Zhu et
Àdàm, K., Søvik, A.K., Krogstad, T., 2006. Sorption of phosphorous to Filtralite-P
al., 2003; Heistad et al., 2006; Öövel et al., 2007). The results from (TM)—the effect of different scales. Water Res. 40, 1143–1154.
the literature indicate that the life span of CW for P removal dif- Àdàm, K., Krogstad, T., Vrale, L., Søvik, A.K., Jenssen, P.D., 2007. Phosphorus reten-
fers. However, even if a medium with high P binding capacity has tion in the filter materials shellsand and Filtralite P (R)—batch and column
experiment with synthetic P solution and secondary wastewater. Ecol. Eng. 29,
been selected, it will become saturated after a few years (Arias 200–208.
et al., 2001). Some new man-made products for P removal have Àdàm, K., Krogstad, T., Suliman, F.R.D., Jenssen, P.D., 2005. Phosphorous sorption
not initially shown such promising results. For example, Filtra P by Filtralite-PTM —small-scale box experiment. J. Environ. Sci. Health A40 (6/7),
1239–1250.
(P removal 98%) clogged after 971 pore volumes probably due to Agyei, N.M., Strydom, C.A., Potgieter, J.H., 2002. The removal of phosphate ions from
structural degradation. aqueous solution by fly ash, slag, ordinary Portland cement and related blends.
In cases in which CW systems are used, where all contaminants Cement Concrete Res. 32, 1889–1897.
Akhurst, D.J., Jones, G.B., Clark, M., McConchie, D., 2006. Phosphate removal from
(SS, BOD, TN, TP etc.) are removed in one filter bed, the clogging of
aqueous solutions using neutralised bauxite refinery residues (BauxsolTM ). Env-
substratum active particle surfaces and pore spaces, with organic iron. Chem. 3, 65–74.
matter (litter, biofilm etc.), can lead to decreasing P removal. In Akratos, C.S., Tsihrintzis, V.A., 2007. Effect of temperature, HRT, vegetation and
porous media on removal efficiency of pilot-scale horizontal subsurface flow
addition, clogging also decreases the hydraulic retention time in
constructed wetlands. Ecol. Eng. 29, 173–191.
the filter bed. An obvious and sustainable solution would also be Altundoğan, H.S., Tümen, F., 2002. Removal of phosphorus from aqueous solutions
a separate filter unit containing replaceable material with a high P by using bauxite. 1. Effect of pH on the adsorption of various phosphates. J. Chem.
binding capacity (Brix et al., 2001). In that case, appropriate pre- Technol. Biotechnol. 77, 77–85.
Altundoğan, H.S., Tümen, F., 2003. Removal of phosphates from aqueous solutions by
treatment would also allow for a longer lifetime of the filter media, using bauxite. II. The activation study. J. Chem. Technol. Biotechnol. 78, 824–833.
by decreasing the risk of clogging or even decreasing the pH and Arias, C.A., Del Bubba, M., Brix, H., 2001. Phosphorus removal by sands for use as
allowing the use of finer reactive filter media with higher sorption media in subsurface flow constructed reed beds. Water Res. 35 (5), 1159–1168.
Arias, C.A., Brix, H., Johansen, N.H., 2003. Phosphorus removal from municipal
capacity (Hedström, 2006). wastewater in an experimental two-stage vertical flow constructed wetland
system equipped with a calcite filter. Water Sci. Technol. 48 (5), 51–58.
Baker, M.J., Blowes, D.W., Ptacek, C.J., 1998. Laboratory development of perme-
able reactive mixtures for the removal of phosphorus from onsite wastewater
4. Conclusions disposal systems. Environ. Sci. Technol. 32 (15), 2308–2316.
Banerjee, K., Cheremisinoff, P.N., Cheng, S.L., 1997. Adsorption kinetics of oxylene
A great variety (more than 30 main categories and over 80 by flyash. Water Res. 31, 249–261.
Bellier, N., Chazarenc, F., Comeau, Y., 2006. Phosphorus removal from wastewater
subtypes) of both natural and man-made materials plus indus-
by mineral apatite. Water Res. 40, 2965–2971.
trial by-products have been applied as filter media for phosphorus Bennett, E.M., Carpenter, R., Caraco, N.F., 2001. Human impact on erodable phos-
retention in constructed wetlands. phorus and eutrophication: a global perspective. BioScience 51, 227–234.
Brix, H., Arias, C.A., Del Bubba, M., 2001. Media selection for sustainable phospho-
A majority of these materials have a pH value >7.0 and high Ca
rus removal in subsurface flow constructed wetlands. Water Sci. Technol. 44
and/or CaO content. Thus the main process of P retention used in (11–12), 47–54.
constructed wetlands is precipitation. Brogowski, Z., Renman, G., 2004. Characterization of opoka as a basis for its use in
Although very few investigations have been performed in the wastewater treatment. Polish J. Environ. Stud. 13, 15–20.
Brooks, A.S., Rozenwald, M.N., Geohring, L.D., Lion, L.W., Steenhuis, T.S., 2000. Phos-
long-term saturation time of materials, most of calculations based phorus removal by wollastonite: a constructed wetland substrate. Ecol. Eng. 15,
on the batch experiments suggest, and some data from full size sys- 121–132.
88 C. Vohla et al. / Ecological Engineering 37 (2011) 70–89

Brown, P.A., Gill, S.A., Allen, S.J., 2000. Metal removal from wastewater using peat. P.D. (Eds.), Constructed Wetlands for Wastewater Treatment in Cold Climates,
Water Res. 34 (16), 3907–3916. Advances in Ecological Sciences, vol. 11. WIT Press, Southampton, Boston, pp.
Bulc, T., Vrhovšek, D., Kukanja, V., 1997. The use of constructed wetlands for landfill 273–297.
leachate treatment. Water Sci. Technol. 35, 301–306. Hartikainen, H., 1982. Relationship between phosphorus intensity and capacity
Cameron, K., Madramootoo, C., Crolla, A., Kinsley, C., 2003. Pollutant removal from parameters in Finnish mineral soils. II. Sorption–desorption isotherms and their
municipal sewage lagoon effluents with a free-surface wetland. Water Res. 37, relation to soil characteristics. J. Agric. Sci. Finland 54, 251–262.
2803–2812. He, S.-B., Yan, L., Kong, H.-N., Liu, Z.-M., Wu, D.-Y., Hu, Z.-B., 2007. Treatment efficien-
Cha, W., Kim, J., Choi, H., 2007. Evaluation of steel slag for organic and inorganic cies of constructed wetlands for eutrophic landscape river water. Pedosphere
removals in soil aquifer treatment. Water Res. 40, 1034–1042. 17, 522–528.
Chazarenc, F., Brisson, J., Comeau, Y., 2007. Slag columns for upgrading phosphorus Heal, K.V., Younger, P.L., Smith, K.A., Glendinning, S., Quinn, P., Dobbie, K.E., 2003.
removal from constructed wetland effluents. Water Sci. Technol. 56, 109–115. Novel use of ochre from mine water treatment plants to reduce point and diffuse
Chen, J., Kong, H., Wu, W., Chen, X., Zhang, D., Sun, Z., 2007. Phosphate immobi- phosphorus pollution. Land Contam. Reclam. 11, 145–152.
lization from aqueous solution by fly ashes in relation to their composition. J. Heal, K.V., Dobbie, K.E., Bozika, E., McHaffie, H., Simpson, A.E., Smith, K.A., 2005.
Hazard. Mater. B139, 293–300. Enhancing phosphorus removal in constructed wetlands with ochre from mine
Chen, J., Kong, H., Wu, D., Hu, Z., Wang, Z., Wang, Y., 2006. Removal of phosphate drainage treatment. Water Sci. Technol. 51, 275–282.
from aqueous solution by zeolite synthesized from fly ash. J. Colloid Interf. Sci. Hedström, A., 2006. Wollastonite as a reactive filter medium for sorption of wastew-
300, 491–497. ater ammonium and phosphorus. Environ. Technol. 27, 801–809.
Cheug, K.C., Venkitachalam, T.H., 2000. Improving phosphate removal of sand infil- Heikkinen, K., Ihme, R., Osma, A.-M., Hartikainen, H., 1995. Phosphate removal by
tration system using alkaline fly ash. Chemosphere 41, 243–249. peat from peat mining drainage water during overland flow wetland treatment.
Ciupa, R., 1996. The experience in the operation of constructed wetlands in North- J. Environ. Qual. 24, 597–602.
Eastern Poland. In: Proceedings of Fifth International Conference Wetland Heistad, A., Paruch, A.M., Vråle, L., Ádám, K., Jenssen, P.D., 2006. A high-performance
Systems for Water Pollution Control, IWA and Universität für Bodenkultur, compact filter system treating domestic wastewater. Ecol. Eng. 28, 374–379.
Vienna (Chapter IX/6). Hill, C.M., Duxbury, J., Geohring, L., Peck, T., 2000. Designing constructed wetlands
Cooper, P.F., Findlater, B.C., 1990. Constructed Wetlands in Water Pollution Control. to remove phosphorus from barnyard runoff: a comparison of four alternative
Pergamon Press, Oxford, UK. substrates. J. Environ. Sci. Health A35, 1357–1375.
DeBusk, T.A., Grace, K.A., Dierberg, F.E., Jackson, S.D., Chimney, M.J., Gu, B., 2004. Hylander, L.D., Kietlinska, A., Renman, G., Simán, G., 2006. Phosphorus retention in
An investigation of the limits of phosphorus removal in wetlands: a meso- filter materials for wastewater treatment and its subsequent suitability for plant
cosm study of a shallow periphyton-dominated treatment system. Ecol. Eng. 23, production. Bioresour. Technol. 97, 914–921.
1–14. Inglethorpe, S.D.J. 1992. Measurement of selected physical properties of three sam-
Diamadopoulos, E., Ioannidis, S., Sakellaropoulos, G.P., 1993. As (V) removal from ples submitted by SAC consultants. British Geological Survey Technical Report
aqueous solutions by fly ash. Water Res. 27, 1773–1777. MPSR/92/17.
Drizo, A., 1998. Phosphate and ammonium removal from wastewater, using con- Jenssen, P.D., Krogstad, T., 2003. Design of constructed wetlands using phospho-
structed wetland systems. Ph.D. Thesis. University of Edinburgh, Scotland. rus sorbing lightweight aggregate (LWA). In: Mander, Ü., Jenssen, P.D. (Eds.),
Drizo, A., Frost, C.A., Smith, K.A., Grace, J., 1997. Phosphate and ammonium removal Constructed Wetlands for Wastewater Treatment in Cold Climates, Advances in
by constructed wetlands with horizontal subsurface flow, using shale as a sub- Ecological Sciences, vol. 11. WIT Press, Southampton, Boston, pp. 259–272.
strate. Water Sci. Technol. 35, 95–102. Jenssen, P.D., Krogstad., T., Briseid, T., Norgaard, E., 1991. Testing of reactive filter
Drizo, A., Frost, C.A., Grace, J., Smith, K.A., 1999. Physico-chemical screening of media (LECA) for use in agricultural drainage systems. In: International Seminar
phosphate-removing substrates for use in constructed wetland systems. Water of the Technical Section of CIGR on Environmental Challenges and Solutions in
Res. 33 (17), 3595–3602. Agricultural Engineering, Ås-NLH.
Drizo, A., Comeau, Y., Forget, C., Chapuis, R.P., 2002. Phosphorus saturation poten- Johansson, L., 1997a. Phosphorus sorption to filter substrates-potential benefits for
tial: a parameter for estimating the longevity of constructed wetland systems. on-site wastewater treatment. Doctoral Thesis. Royal Institute of Technology
Environ. Sci. Technol. 36, 4642–4648. (KTH), Sweden.
Drizo, A., Forget, C., Chapuis, R.P., Comeau, Y., 2006. Phosphorus removal by electric Johansson, L., 1997b. The use of LECA (Light Expanded Clay Aggregates) for the
arc furnace steel slag and serpentinite. Water Res. 40, 1547–1554. removal of phosphorus from wastewater. Water Sci. Technol. 35, 87–93.
Erstad (Jørgensen), K.J., 1982. Sammenlikning av norske kalkingsmidler I sammen- Johansson, L., 1999a. Industrial by-products and natural substrata as phosphorus
heng med finfordelingsgradens betydning for virkninga på jordreaksjon og sorbents. Environ. Technol. 20, 309–316.
plantevekst. Hovedoppgave ved NHL/Institutt for Jordkultur. Johansson, L., 1999b. Blast furnace slag as phosphorus sorbents column studies. Sci.
Eye, J.D., Basu, T.K., 1970. The use of fly ash in wastewater treatment and sludge Total Environ. 229, 89–97.
conditioning. J. Water Pollut. Control Fed. 42, 125–135. Johansson, L., Gustafsson, J.P., 2000. Phosphorus removal from wastewater by filter
Farahbakhshazad, N., Morrison, G.M., 2003. Phosphorus removal in a vertical upflow media: retention and estimated plant availability of sorbed phosphorus. Water
constructed wetland system. Wat. Sci. Technol. 48, 43–50. Res. 34, 259–265.
Faulkner, S.P., Richardson, C.J., 1989. Physical and chemical characteristics of fresh- Johansson, L., Hylander, L.D., 1998. Phosphorus removal from wastewater by filter
water wetland soils. In: Hammer, D. (Ed.), Constructed Wetlands for Waste media: retention and estimated plant availability of sorbed phosphorus. J. Polish
Water Treatment. Municipal, Industrial and Agricultural. Lewis Publishers, Acad. Sci. 458, 397–409.
Chelsea, MI, pp. 121–129. Johansson Westholm, L., 2006. Substrates for phosphorus removal—potential ben-
Gerritse, R.G., 1993. Mobility of phosphate from waste water in calcareous sands of efits for on-site wastewater treatment? Water Res. 40, 23–36.
Rottnest Island (W.A.). Aust. J. Soil Res. 31, 235–244. Joko, I., 1984. Phosphorus removal from wastewater by the crystallization method.
Gervin, L., Brix, H., 2001. Reduction of nutrients from combined sewer overflows and Water Sci. Technol. 17, 121–132.
lake water in a vertical-flow constructed wetland system. Water Sci. Technol. Kaasik, A., Vohla, C., Mõtlep, R., Mander, Ü., Kirsimäe, K., 2008. Hydrated calcareous
44 (11–12), 171–176. oil-shale ash as potential filter media for phosphorus removal in constructed
Giesler, R., Andersson, T., Lövgren, L., Persson, P., 2005. Phosphate sorption in wetlands. Water Res. 42, 1315–1323.
aluminium- and iron-rich humus soils. Soil Sci. Soc. Am. J. 69, 77–86. Kadlec, R.H., Knight, R.L., 1996. Treatment Wetlands. Lewis Publishers, CRC Press,
Gray, C.A., Schwab, A.P., 1993. Phosphorus-fixing ability of high pH, high calcium, Boca Raton, 893 pp.
coal-combustion, waste materials. Water Air Soil Pollut. 69, 309–320. Kadlec, R.H., Wallace, S.D., 2008. Treatment Wetlands, 2nd ed. CRC Press, Boca Raton,
Gray, M.N., Rock, C.A., Pepin, R.G., 1988. Pretreating landfill leachate with biomass 952 pp.
boiler ash. J. Environ. Eng. 114, 465–470. Kao, P.-C., Tzeng, J.-H., Huang, T.-L., 2000. Removal of chlorophenols from aqueous
Gray, S., Kinross, J., Read, P., Marland, A., 2000. The nutrient assimilative capacity of solution by fly ash. J. Hazard. Mater. 76, 237–249.
maerl as a substrate in constructed wetland systems for waste treatment. Water Karaca, S., Gürses, A., Ejder, M., Açıkyıldız, M., 2004. Kinetic modeling of liquid-phase
Res. 34, 2183–2190. adsorption of phosphate on dolomite. J. Colloid Interface Sci. 277, 257–263.
Grubb, D.G., Guimaraes, M.S., Valencia, R., 2000. Phosphate immobilization using an Kim, E.H., Yim, S.B., Jung, H.C., Lee, E.J., 2006. Hydroxyapatite crystallization from
acidic type F fly ash. J. Hazard. Mater. 76, 217–236. a highly concentrated phosphate solution using powdered converter slag as a
Grüneberg, B., Kern, J., 2001. Phosphorus retention capacity of iron-ore and blast seed material. J. Hazard. Mater. 136 (3), 690–697.
furnace slag in subsurface flow constructed wetlands. Water Sci. Technol. 44, Kõiv, M., Kriipsalu, M., Vohla, C., Mander, Ü., 2009a. The secondary treatment of
69–75. landfill leachate in vertical flow filters using hydrated oil shale ash and peat.
Gupta, G.S., Prasad, G., Singh, V.N., 1990. Removal of chrome dye from aqueous Fresen. Environ. Bull. 18, 189–195.
solutions by mixed adsorbents fly ash and coal. Water Res. 24, 45–50. Kõiv, M., Vohla, C., Mõtlep, R., Liira, M., Kirsimäe, K., Mander, Ü., 2009b. The per-
Gustafsson, J.P., Renman, A., Renman, G., Poll, K., 2008. Phosphate removal by formance of peat-filled subsurface flow filters treating landfill leachate and
mineral-based sorbents used in filters for small-scale wastewater treatment. municipal wastewater. Ecol. Eng. 35, 204–212.
Water Res. 42, 189–197. Korkusuz, E.A., Beklioglu, M., Demirer, G.N., 2005. Comparison of the treatment per-
Hall-Spencer, J., 1998. Conservation issues relating to maerl beds as habitats for formances of blast furnace slag-based and gravel-based vertical flow wetlands
molluscs. J. Conchol. Special Publ. 2, 271–286. operated identically for domestic wastewater treatment in Turkey. Ecol. Eng.
Hare, C.H., 1993. The evolution of calcium metasilicate in paint and coatings. Mod. 24, 187–200.
Paint Coating. 83 (12), 56–63. Korkusuz, E.A., Beklioglu, M., Demirer, G.N., 2007. Use of blast furnace granulated
Harris, T.Z., Mæhlum, T., 2003. Nitrogen removal in light-weight aggregate pre- slag as a substrate in vertical flow reed beds: field application. Bioresour. Tech-
treatment filter columns and mesocosm wetlands. In: Mander, Ü., Jenssen, nol. 98, 2089–2101.
C. Vohla et al. / Ecological Engineering 37 (2011) 70–89 89

Krogstad, T., Sogn, T.A., Asdal, A., Saebo, A., 2005. Influence of chemically and bio- Saltnes, T., Føllesdal, M., 2005. Final Report: Material Development. 02056 Wastew-
logically stabilized sewage sludge on plant-available phosphorous in soil. Ecol. ater treatment in filter beds, maxit Group AB.
Eng. 25, 51–60. Schindler, D.W., 1977. Evolution of phosphorus limitation in lakes. Science 195,
Kukier, U., Sumner, M.E., Miller, W.P., 1994. Boron release from fly ash and its uptake 260–262.
by corn. J. Environ. Qual. 23, 596–603. Seo, D.C., Cho, J.S., Lee, H.J., Heo, J.S., 2005. Phosphorus retention capacity of filter
Kvarnstrom, M.E., Morel, C.A.L., Krogstad, T., 2004. Plant-availability of phospho- media for estimating the longevity of constructed wetland. Water Res. 39 (11),
rus in filter substrates derived from small-scale wastewater treatment systems. 2445–2457.
Ecol. Eng. 22, 1–15. Shannon, E.E., Verghese, K.I., 1976. Utilisation of aluminised red mud solids for
Kwon, H.-B., Lee, C.-W., Jun, B.-S., Yun, J-d., Weon, S.-Y., Koopman, B., 2004. Recy- phosphorus removal. JWPCF 48, 1948–1954.
cling waste oyster shells for eutrophication control. Resour. Conserv. Recycl. 41, Sharpley, A.N., Tunney, H., 2000. Phosphorus research strategies to meet agricultural
75–82. and environmental challenges of the 21st century. J. Environ. Qual. 29, 176–
Li, Y., Liu, C., Luan, Z., Peng, X., Zhu, C., Chen, Z., Zhang, Z., Fan, J., Jia, Z., 2006. Phosphate 181.
removal from aqueous solutions using raw and activated red mud and fly ash. Shilton, A.N., Elmetri, I., Drizo, A., Pratt, S., Haverkamp, R.G., Bilby, S.C., 2006. Phos-
J. Hazard. Mater. B137, 374–383. phorus removal by an ‘active’ slag filter-a decade of full scale experience. Water
Liira, M., Kõiv, M., Mander, Ü., Mõtlep, R., Vohla, C., Kirsimäe, K., 2009. Active fil- Res. 40, 113–118.
tration of phosphorus on Ca-rich hydrated oil-shale ash: does longer retention Shilton, A., Pratt, S., Drizo, A., Mahmood, B., Banker, S., Billings, L., Glenny, S., Luo,
time improve the process? Environ. Sci. Technol. 43, 3809–3814. D., 2005. ‘Active’ filters for upgrading phosphorus removal from pond systems.
Mæhlum, T., 1998. Cold-climate constructed wetlands: aerobic pre-treatment and Water Sci. Technol. 51 (12), 111–116.
horizontal subsurface flow systems for domestic sewage and landfill leachate Song, Y., Hahn, H.H., Hoffmann, E., 2002. Effects of solution conditions on the precip-
purification. Ph.D. Thesis. Agricultural University of Norway, Ås. itation of phosphate for recovery: a thermodynamic evaluation. Chemosphere
Mæhlum, T., 1999. Wetlands for treatment of landfill leachate in cold climates. In: 48, 1029–1034.
Mulamoottil, G., McBean, E.A., Rovers, F. (Eds.), Constructed Wetlands for the Soranno, P.A., Hubler, L., Carpenter, S.R., Lathrop, R.C., 1996. Phsophorus loads to
Treatment of Landfill Leachates. Lewis Publishers, Florida, Boca Raton, US, pp. surface waters: a simple model to account for spatial pattern of land use. Ecol.
33–46. Appl. 6, 865–878.
Mæhlum, T., Jenssen, P.D., Warner, W.S., 1995. Cold-climate constructed wetlands. Søvik, A.K., Kløve, B., 2005. Phosphorus retention processes in shell sand filter sys-
Water Sci. Technol. 32, 95–101. tems treating municipal wastewater. Ecol. Eng. 25 (2), 168–182.
Mander, Ü., Teiter, S., Kuusemets, V., Lõhmus, K., Öövel, M., Nurk, Augustin, J., Strang, T.J., Wareham, D.G., 2006. Phosphorus removal in a waste-stabilization pond
2003. Nitrogen and phosphorus budgets in a subsurface flow wastewater treat- containing limestone rock filters. J. Environ. Eng. Sci. 5, 447–457.
ment wetland. In: Brebbia, C.A. (Ed.), Water Resources Management. IWIT Press, Stumm, W., Morgan, J.J., 1981. Aquatic Chemistry—An Introduction Emphasizing
Southampton, Boston. Chemical Equilibrist in Natural Waters. Wiley, New York.
Mann, R., 1990. Phosphorus removal by constructed wetlands: substratum adsorp- Szögi, A.A., Humenik, F.J., Rice, J.M., Hunt, P.G., 1997. Swine wastewater treatment
tion. In: Cooper, P.F., Findlater, B.C. (Eds.), Constructed Wetlands in Water by media filtration. J. Environ. Sci. Health B32, 831–843.
Pollution Control. Pergamon Press, Oxford, UK. Talbot, P., Bélanger, G., Pelletier, M., Laliberté, G., Arcand, Y., 1996. Development of a
Mann, R.A., Bavor, H.J., 1993. Phosphorus removal in constructed wetlands using biofilter using an organic medium for on-site wastewater treatment. Water Sci.
gravel and industrial waste substrata. Water Sci. Technol. 27, 107–113. Technol. 34, 435–441.
Mayes, W.M., Batty, L.C., Younger, P.L., Jarvis, A.P., Kõiv, M., Vohla, C., Mander, U., Tanner, C.C., Sukias, J.P.S., Upsdell, M.P., 1999. Substratum phosphorus accumulation
2009. Wetland treatment at extremes of pH: a review. Sci. Total Environ. 407, during maturation of gravel-bed constructed wetlands. Water Sci. Technol. 40,
3944–3957. 147–154.
Molle, P., Lienard, A., Grasmick, A., Iwema, A., Kabbabi, A., 2005. Apatite as an inter- Tiessen, H., 1995. Phosphorus in the Global Environment: Transfers, Cycles, and
esting seed to remove phosphorus from wastewater in constructed wetlands. Management. John Wiley and Sons, New York.
Water Sci. Technol. 51 (9), 193–203. Tisdale, S.L., Nelson, W.L., Beaton, J.D., Havlin, J.L., 1993. Soil Fertility and Fertilizers,
Netter, R., 1992. Flow characteristics of planted soil filters. Water Sci. Technol. 29, 5th ed., Macmillan Coll. Div., p. 634.
29–36. Tsitouridou, R., Georgiou, J., 1988. Contribution to the study of phosphate sorption
Nieminen, M., Jarva, M., 1996. Phosphorus adsorption by peat from drained mires by three Greek fly ashes. Toxicol. Environ. Chem. 17, 129–138.
in Southern Finland. Scand. J. Forest Res. 11, 321–326. Vohla, C., Alas, R., Nurk, K., Baatz, S., Mander, Ü., 2007. Phosphorus retention capacity
Oguz, E., 2004. Removal of phosphate from aqueous solution with blast furnace slag. in a horizontal subsurface flow constructed wetland. Sci. Total Environ. 380,
J. Hazard. Mater. B114, 131–137. 66–74.
Öövel, M., Tooming, A., Mauring, T., Mander, Ü., 2007. Schoolhouse wastewater Vohla, C., Põldvere, E., Noorvee, A., Kuusemets, V., Mander, Ü., 2005. Alternative
purification in a LWA-filled hybrid constructed wetland in Estonia. Ecol. Eng. filter media for phosphorus removal in a horizontal subsurface flow constructed
29, 17–26. wetland. J. Environ. Sci. Health A 40, 1251–1264.
Özacar, M., 2006. Contact time optimization of two-stage batch adsorber design Vollenweider, R.A., 1968. The scientific basis of lake and stream eutrophication
using second-order kinetic model for the adsorption of phosphate onto alunite. with particular reference to phosphorus and nitrogen as eutrophication factors.
J. Hazard. Mater. B 137, 218–225. Technical Report OECD, DAS/C81/68, Paris, France.
Pant, H.K., Reddy, K.R., Lemon, E., 2001. Phosphorus retention capacity of root bed Vymazal, J., Brix, H., Cooper, P.F., Green, M.B., Haberl, R., 1998. Constructed Wet-
media of subsurface flow constructed wetlands. Ecol. Eng. 17, 345–355. lands for Wastewater Treatment in Europe. Backhuys Publishers, Leiden, The
Piekos, R., Paslawska, S., 1999. Fluoride uptake characteristics of fly ash. Fluoride 32, Netherlands.
14–19. Vymazal, J., Brix, H., Cooper, P.F., Haberl, R., Grüneberg, B., Kern, J., 2000.
Prochaska, C.A., Zouboulis, A.I., 2006. Removal of phosphates by pilot vertical-flow Phosphorus retention capacity of iron-ore and blast furnace slag in sub-
constructed wetlands using a mixture of sand and dolomite as substrate. Ecol. surface flow constructed wetlands. In: 7th International Conference of
Eng. 26, 293–303. Wetlands Systems for Water Pollution Control, vol. 1, University of
Proctor, D.M., Fehling, K.A., Shay, E.C., Wittenbornand, J.L., Green, J.J., Avent, C., Florida, Grosvenor Resort, Lake Buena Vista, Florida, November 11–16, p.
Bigham, R.D., Connolly, M., Lee, B., Shepker, T.O., Zak, M.A., 2000. Physical and 113–20.
chemical characteristics of blast furnace, basic oxygen furnace, and electric arc Willadsen, C.T., Riger-Kusk, O., Qvist, B., 1990. Removal of nutrient salts from two
furnace steel industry slags. Environ. Sci. Technol. 34, 1576–1582. Danish root zone systems. In: Cooper, B.C. (Ed.), Use of Constructed Wetlands in
Renman, G., Kietlinska, A., Cabamas, V., 2004. Treatment of phosphorus and bacteria Water Pollution Control. Pergamon, Oxford, pp. 115–126.
by filter media in onsite wastewater disposal systems. In: 2nd International Wood, R.B., McAtamney, C.F., 1996. Constructed wetlands for waste water treat-
Symposium, Ecosan Closing the Loop, Lübeck, Germany, pp. 573–576. ment: the use of laterite in the bed medium in phosphorus and heavy metal
Rex, M., 2000. Blast furnace and steel slags as liming materials for sustainable removal. Hydrobiologia 340, 323–331.
agricultural production. In: Proceedings of the 2nd European Slag Conference, Xu, D.F., Xu, J.M., Wu, J.J., Muhammad, A., 2006. Studies on the phosphorus sorption
Düsseldorf, EUROSLAG Publication, No. 1, pp. 137–149. capacity of substrates used in constructed wetland systems. Chemosphere 63,
Richardson, C.J., Qian, S.S., Craft, C.B., 1996. Predictive models for phosphorus reten- 344–352.
tion in wetlands. In: Vymazal, J. (Ed.), Proceedings of Conference on Nutrient Yamada, H., Kayama, M., Saitu, K., Hara, M., 1986. A fundamental research on phos-
Cycling and Retention in Wetlands and Their Use for Wastewater Treatment, phate removal by using slag. Water Res. 20, 547–557.
Prague, Institute of Botany, Trebon, Czech Republic, pp. 125–150. Yan, J., Kirk, D.W., Jia, C.Q., Liu, X., 2007. Sorption of aqueous phosphorus onto
Rizzuti, A.M., Cohen, A.D., Hunt, P.G., Ellison, A.Q., 2002. Retention of nitrogen and bituminous and lignitous coal ashes. J. Hazard. Mater. 148, 395–401.
phosphorous from liquid swine and poultry manures using highly characterized Zhu, T., 1998. Fate of phosphorous in a light-weight aggregate (LWA) wastewater
peats. J. Environ. Sci. Health B37, 587–611. treatment wetland. Ph.D. Thesis. Agricultural University of Norway, Ås.
Roseth, R., 2000. Shell sand a new filter medium for constructed wetlands and Zhu, T., Jenssen, P.D., Mæhlum, T., Krogstad, T., 1997. Phosphorus sorption and chem-
wastewater treatment. J. Environ. Sci. Health A35, 1335–1355. ical characteristics of lightweight aggregates (LWA)—potential filter media in
Rustige, H., Tomac, I., Höner, G., 2003. Investigations on phosphorus retention in treatment wetlands. Water Sci. Technol. 35, 103–108.
subsurface flow constructed wetlands. Water Sci. Technol. 48, 67–74. Zhu, T., Mæhlum, T., Jenssen, P.D., Krogstad, T., 2003. Phosphorus sorption charac-
Sakadevan, K., Bavor, H.J., 1998. Phosphate adsorption characteristics of soils, slags teristics of light-weight aggregate. Water Sci. Technol. 48, 93–100.
and zeolite to be used as substrates in constructed wetland systems. Water Res.
32, 393–399.

Das könnte Ihnen auch gefallen