Sie sind auf Seite 1von 10

Available online at www.sciencedirect.

com

Bioresource Technology 99 (2008) 5881–5890

Waste treatment and biogas quality in small-scale agricultural digesters


a,*
Stephanie Lansing , Raúl Botero Botero b, Jay F. Martin a

a
Department of Food, Agricultural and Biological Engineering, The Ohio State University, 590 Woody Hayes Drive,
Columbus, OH 43210-1057, United States
b
EARTH University, Apartado Postal 4442 – 1000, San Jose, Costa Rica

Received 22 June 2006; received in revised form 7 September 2007; accepted 25 September 2007
Available online 26 November 2007

Abstract

Seven low-cost digesters in Costa Rica were studied to determine the potential of these systems to treat animal wastewater and pro-
duce renewable energy. The effluent water has a significantly lower oxygen demand (COD decreased from 2968 mg/L to 472 mg/L) and
higher dissolved nutrient concentration (NH4-N increased by 78.3% to 82.2 mg/L) than the influent water, which increases the usefulness
of the effluent as an organic fertilizer and decreases its organic loading on surface waters. On average, methane constituted 66% of the
produced biogas, which is consistent with industrial digesters. Through principle component analysis, COD, turbidity, NH4-N, TKN,
and pH were determined to be the most useful parameters to characterize wastewater. The results suggest that the systems have the abil-
ity to withstand fluctuations in the influent water quality. This study revealed that small-scale agricultural digesters can produce methane
at concentrations useful for cooking, while improving the quality of the livestock wastewater.
Ó 2007 Elsevier Ltd. All rights reserved.

Keywords: Anaerobic digestion; Manure; Nitrogen; Renewable energy; Water quality

1. Introduction by the digester systems in this study was used to meet the
farmers’ energy needs for cooking.
In developing countries, water pollution and access to Biogas is derived through anaerobic digestion of bio-
energy resources present challenges to human health, envi- mass, such as animal wastes, municipal wastewater, and
ronmental health, and economic development. Small-scale, landfill waste. Anaerobic digestion is a microbially medi-
economically feasible technologies that combine wastewa- ated biochemical degradation of complex organic material
ter treatment and energy production can simultaneously into simple organics and dissolved nutrients. Digesters are
protect the surrounding water resources and enhance physical structures that facilitate anaerobic digestion by
energy availability. In developing countries, animal rearing providing an anaerobic environment for the organisms
creates a resource that can be used in anaerobic digesters to responsible for digestion. Processing livestock manure
produce methane gas for cooking. In this study, seven agri- through anaerobic digesters captures methane, which can
cultural digesters in Costa Rica that used excreta from be used as an energy source while reducing emissions of
bovines and swine were studied to determine the extent this greenhouse gas.
of wastewater treatment and the quality of the produced Currently, digesters are concentrated in developing coun-
gas. The variability between these digesters was studied tries, with over 5 million household digesters constructed in
to determine if management styles and wastewater strength China and India alone (Huttunen and Lampinen, 2005).
had an effect on wastewater treatment and the concentra- Digesters built around the world vary in their design com-
tion of methane in the produced biogas. Biogas produced plexity, construction materials, and costs. In developed
countries, digesters often are concrete stirred tank reactors
*
Corresponding author. Tel.: +1 918 749 1282; fax: +1 614 292 9448. (CSTRs), in which a portion of the produced biogas is uti-
E-mail address: lansing.10@osu.edu (S. Lansing). lized to heat the digester (Berglund and Börjesson, 2006).

0960-8524/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2007.09.090
5882 S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890

In developing countries, many of the digesters do not have as detailed in Table 1 (Botero and Preston, 1987; Erickson
mixing components, do not require continuous monitoring, et al., 2004; Erickson and Fung, 1998). Previous studies
and are adaptable to any tropical climate (Chará et al., 1999; have shown that small-scale digesters reduce the amount
Ong et al., 2000). The Taiwanese-model digesters are simple, of organic matter and solids in the animal wastewater by
flow-through reactors consisting of a tubular polyethylene 55–90% (Chen and Shyu, 1996; Pedraza et al., 2001; Botero
bag, PVC piping, and plastic hosing to transport the biogas and Hernández, 2005). In addition, the digester effluent has
from the digester (Fig. 1) (Chará et al., 1999; Botero and little odor and an increased concentration of dissolved
Preston, 1987; An and Preston, 1999). The construction, nutrients, which provides farmers with an improved
materials, and labor costs for construction of a Taiwanese- organic fertilizer (Parsons, 1984; Botero and Preston,
model digester can vary from $34 USD in Vietnam (An 1987: Chará et al., 1999; Sophea and Preston, 2001; Sophin
et al., 1997) to $150 USD in Costa Rica. and Preston, 2001; Thy and Preston, 2003).
Biogas from Taiwanese-model digesters has been suc- The majority of digester studies have been conducted on
cessfully collected and used for cooking, eliminating the technologically advanced, lab-scale digesters (Griffin et al.,
need to buy propane or use firewood. By using biogas in 1998; Lettinga et al., 1999; Sterling et al., 2001; Collins
place of burning biomass, indoor air quality is dramatically et al., 2003; Erickson et al., 2004; McMahon et al., 2004).
improved and deforestation is reduced (Chará et al., 1999). There has been a lack of experimental studies on Taiwan-
The biogas from digesters is composed of 50–70% methane, ese-model digesters because of their development as a prac-
30–40% carbon dioxide, and trace amounts of other gases, tical, easy to implement technology for utilization in
tropical countries (Botero and Hernández, 2005). Studies
have shown that the Taiwanese model-digesters can pro-
duce biogas with methane concentrations above 60%, with
temperatures at or below the mesophilic range (20–40 °C)
(An and Preston, 1999; Botero and Preston, 1987). Most
previous studies have concentrated on the general success
of a single digester (Gowda, 1995; An et al., 1997; Moog
et al., 1997; Chará et al., 1999; Sophea and Preston,
2001; Sophin and Preston, 2001; Esquivel et al., 2002; Bote-
ro and Hernández, 2005). Rigorous water quality analyses
of the influent and effluent wastewater have rarely been
Fig. 1. Taiwanese-model digesters are flow-through reactors consisting of used on small-scale systems due their concentration in
a tubular polyethylene bag, PVC piping, and plastic hosing. The biogas is
transported from the digester to the kitchen/cooking area, where it is
remote areas of developing countries. This study strives
utilized for up to 12 h of cooking per day. The length of the digester can to statistically access the ability of multiple small-scale
vary from 10 to 20 m, with a 5 m circumference. The wastewater to biogas digesters to treat animal wastewater and obtain biogas with
ratio is approximately 1:3 (Botero and Preston, 1987). high methane levels.
In this study, twelve influent and effluent wastewater
Table 1
parameters were analyzed to determine statistically signifi-
Relative composition of biogas in a Taiwanese-model anaerobic digester cant trends. The study was conducted on seven digester
(Botero and Preston, 1987) sites to assess variability within these systems, which were
Gas component Percentage in digester (%) identical in construction materials, but differed in digester
Methane (CH4) 50–70
length, wastewater management styles, wastewater source,
Carbon dioxide (CO2) 30–40 and hydrologic loading (Table 2). The objectives of this
Hydrogen (H2) 1.0 study were to determine the following: (1) significant waste-
Nitrogen (N2) 0.5 water characteristics in the treatment process that should
Carbon monoxide (CO) 0.1 be monitored in the future; (2) the variability of water qual-
Oxygen (O2) 0.1
Hydrogen sulfide (H2S) 0.1
ity parameters and methane concentration between differ-
ent digesters.

Table 2
The characteristics for each of the seven digesters studied
Farm Elevation (m) Waste source Washes per day Washing time (min) Digester volume (m3) Digester retention time (days)
1 50 65 pigs 2 45 70 11.1
2 50 30 cows 2 20 60 19.3
3 50 40 pigs 2 30 40 16.2
4 70 12 pigs 1 20 40 44.4
5 250 7 pigs 2 10 40 51.3
6 350 5 pigs 3 5 25 91.4
7 250 3 pigs 2 15 25 55.6
S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890 5883

2. Methods ciple components. Each principal component accounts for


as much of the data variability as possible, with the first
2.1. Study site component retaining the characteristics that contribute
the most to the variability (Jolliffe, 2002).
The seven digesters studied were located in the Limon Multivariate analysis of variances (MANOVA) was con-
Province of Costa Rica (10°N, 83°W). Four digesters were ducted to determine if there were significance differences in
located on small-production farms and three digesters were the overall dataset. Individual analysis of variance
located at EARTH University, an international undergrad- (ANOVA) and subsequent Tukey–Kramer multiple com-
uate university for the study of sustainable agriculture. The parison analyses determined which wastewater variables
digesters analyzed in this study were all Taiwanese-model were significantly different between the seven farm digest-
plastic digesters (Fig. 1). The digester varied in elevation ers. A multiple linear regression (MLR) was preformed
from 350 m (Farm 6) to 50 m (Farms 1, 2 and 3) (Table to compare the CH4 concentration to the wastewater vari-
2). All digesters used animal wastewater, with the majority ables that were determined to be significant from the PCA
using swine manure (Table 2). Manure was washed directly analysis. All statistical analyses were conducted using
from the stalls into the digester. The frequency and amount SASÓ (SAS Institute, Inc) or PC-ORDÓ software (McCune
of wastewater used in each digester varied between the and Medford, 1999).
seven farms, with an average stall-washing time of Ten of the twelve environmental variables collected dur-
15 min, 2 times a day. The retention time in each digester ing the study period were used in the statistical analyses
varied from 11 to 91 days depending on the amount of (temperature, pH, conductivity, COD, BOD5, turbidity,
water used for stall washing and the digester volume (Table TSS, NH4-N, PO4-P, and TKN). The high concentrations
2). of solids in digester influent samples interfered with the
DO and NOx-N analyses (APHA, 1989), and thus, these
2.2. Analytical methods results were not used in the statistical analyses. The follow-
ing environmental variables were log-transformed in order
Wastewater influent and effluent samples were collected to meet the assumptions of the PCA and MANOVA: con-
from each digester from July, 2005 to October, 2005. Tem- ductivity, turbidity, COD, BOD5, TSS, PO4-P, and TKN.
perature, pH, dissolved oxygen (DO), and conductivity Before transformation, these variables had high skewness
measurements were taken on-site using a hand-held 556 (2.9–0.58), high kurtosis (11.1–0.69), were not normal, or
MPS YSI probe. Four site visits were made to each farm.
TM
did not have homogeneity of variances. After transforma-
Three influent and three effluent wastewater samples were tion, the skewness (0.79–0.05), kurtosis (2.39–0.12), nor-
collected from each digester during each site visit. In total, mality, and the homogeneity of variances improved, with
at least twelve influent and twelve effluent samples were COD, BOD5, PO4-P, and TKN having homogeneity of
taken from each digester. All wastewater analyses were variance at the 0.05 significance level.
conducted at EARTH University’s Soil and Water Labora-
tory. Each sample was analyzed according to standard
methods (APHA, 1989) for the following: biochemical oxy- 3. Results
gen demand (BOD5), chemical oxygen demand (COD), tur-
bidity, total suspended solids (TSS), ammonium-nitrogen 3.1. Water quality improvements
(NH4-N), nitrate-nitrite (NOx-N), and orthophosphate
(PO4-P). In addition, samples collected during the final During the digestion process, all of the organic matter
two site visits were analyzed for total kjeldahl nitrogen and solid variables showed significant decreases (ANOVA
(TKN) (APHA, 1989). p-values <0.001) (Tables 3 and 4). The average COD of
An IR-30M hydrocarbon meter (Environmental Sensors the influent wastewater decreased 84.1% from 2970 mg/L
Co.) was used to detect on-site methane (CH4) concentra- to 472 mg/L, and the BOD5 decreased 79.4% to 96.2 mg/
tions, and a Z-900 hydrogen sulfide (H2S) meter (Environ- L. The average turbidity decreased 90.5% from 1820
mental Sensors Co.) was used to detect on-site H2S NTU to 172 NTU, and the TSS concentration decreased
concentrations. 85.6% L to 319 mg/L (Table 3). The average TKN concen-
tration decreased 45.7% from 306 mg/L to 166 mg/L (p-
2.3. Statistical analysis value < 0.001).
Dissolved nutrient (PO4-P, NH4-N) concentrations and
A principle component analysis (PCA) was conducted to conductivity increased as the wastewater moved through
determine: (1) wastewater variables that were altered dur- the digester (Table 3), with the NH4-N concentration
ing the digestion process, (2) examine redundancy of vari- increasing 78.3% to 82.2 mg/L (p-value < 0.001). The DO
ables, and (3) identify the most important variables to concentration slightly increased in the digester, but both
include in future analyses. PCA is a mathematical proce- the average influent (0.21 mg/L) and effluent (0.54 mg/L)
dure that reduces data dimensionality by transforming DO concentrations were anaerobic. The average pH
correlated variables into uncorrelated variables called prin- decreased from 7.34 to 6.64 (p-value < 0.001).
5884 S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890

Table 3
Average influent and effluent data ± SE (n) for Taiwanese-model digesters (Farms 1–7)
Average influent Average effluent Percent decrease/increase
Temperature (°C) 26.2 ± 0.2 (74) 26.1 ± 0.1 (80) 0.4% decrease
Conductivity (mS/cm) 1.59 ± 0.1 (74) 1.73 ± 0.1 (80) 8.8% increase
pH 7.34 ± 0.1 (74) 6.64 ± 0.04 (80) 9.5% decrease
DO (mg/L) 0.21 ± 0.1 (73) 0.54 ± 0.1 (80) 157% increase
BOD (mg/L) 467 ± 40 (74) 96.2 ± 11 (80) 79.4% decrease
COD (mg/L) 2970 ± 260 (73) 472 ± 40 (79) 84.1% decrease
Turbidity (NTU) 1820 ± 200 (73) 172 ± 15 (80) 90.5% decrease
TSS (mg/L) 2210 ± 223 (72) 319 ± 56 (80) 85.6% decrease
NOx-N (mg/L) 0.92 ± 0.2 (71) 0.18 ± 0.03 (80) 80.4% decrease
PO4-P (mg/L) 13.3 ± 1.7 (74) 15.4 ± 1.4 (80) 15.8% increase
NH4-N (mg/L) 46.1 ± 5.1 (72) 82.2 ± 5.0 (79) 78.3% increase
TKN (mg/L) 306 ± 28 (35) 166 ± 13 (37) 45.7% decrease

Table 4
Individual ANOVA and MANOVA results for the differences between all influent and effluent values and the differences between individual farm influent
and effluent values are given below
Influent/effluent By farm: influent By farm: effluent
p-value (degree of freedom) F-stat p-value (degree of freedom) F-stat p-value (degree of freedom) F-stat
Temperature (°C) 0.672 (1,152) 0.18 <0.001 (6,28) 13.1 <0.001 (6,30) 5.85
Log(Conductivity) 0.003 (1,152) 8.92 0.004 (6,28) 4.12 <0.001 (6,30) 199
pH <0.001 (1,152) 77.1 <0.001 (6,28) 24.0 <0.001 (6,30) 19.9
Log(BOD5) <0.001 (1,152) 129 <0.001 (6,28) 5.91 0.001 (6,30) 5.63
Log(COD) <0.001 (1,150) 197 <0.001 (6,28) 10.1 0.017 (6,30) 3.12
Log(Turbidity) <0.001 (1,151) 272 0.005 (6,28) 3.95 0.005 (6,30) 4.01
Log(TSS) <0.001 (1,150) 156 0.010 (6,28) 3.52 <0.001 (6,30) 6.48
Log(PO4-P) 0.070 (1,152) 3.34 <0.001 (6,28) 10.7 <0.001 (6,30) 26.7
NH4-N (mg/L) <0.001 (1,149) 25.7 0.007 (6,28) 3.74 <0.001 (6,30) 72.7
Log(TKN) <0.001 (1,70) 25.5 0.222 (6,28) 1.48 0.023 (6,30) 2.92
MANOVA <0.001 (10,61) 54.2 <0.001 (60,104) 5.63 <0.001 (60,115) 9.61
Wilks’ k = 0.10 Wilks’ k < 0.001 Wilks’ k < 0.001
The p-value (degrees of freedom) and F-statistics are given for each ANOVA, and the p-value (degrees of freedom), approximate F, and Wilks’ k are given
for each MANOVA.

All the collected influent and effluent data from the


digesters at Farms 1 through 7 were used in the PCA and
MANOVA analyses. Axis 1 and axis 2 of the PCA analysis
had eigenvalues (3.85 and 2.59, respectively) above the bro-
ken stick values (2.93 and 1.93, respectively). Axis 1
explained 39% of the variance, and axis 2 explained 25%.
Together the two axes cumulatively explained 64% of the
variance (Fig. 2).
The PCA analysis shows that log(BOD5), log(COD),
log(TSS), and log(turbidity) are highly correlated in
multivariate space and have a high positive influence on
the location of axis 1. NH4-N, log(conductivity) and
log(PO4-P) are highly correlated and have a high positive
influence on the location of axis 2 and a negative influence
on the location of axis 1. The wastewater variables that
measure organic matter and solids (BOD5, COD, TSS,
and turbidity) were located opposite of the variables that
measure dissolved species (NH4-N, conductivity and PO4-
P) in multivariate space. Log(TKN) influences the location
of both axis 1 and axis 2, but is not clustered in multivar-
iate space with either the organic matter/solids or the dis- Fig. 2. Axis one and axis two of the digester principle component analysis
(PCA) for influent and effluent data from Farms 1–7. Axis one has an
solved species. eigenvalue of 3.85, and the percentage of variance explained is 39%. Axis
All of the environmental variables included in the PCA two has an eigenvalue of 2.59, and the percentage of variance explained is
influenced the location of the two axes, and thus, all the 25%.
S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890 5885

variables were used in the MANOVA analyses. The values were used in the PCA analysis comparing individual
MANOVA shows significant differences between farm farm digester influent and effluent water (Fig. 3). Axis 1
digester influent and effluent data (p-value < 0.001; Wilks and axis 2 had eigenvalues (5.66 and 2.18, respectively)
k = 0.10) (Table 4). The individual ANOVAs show signifi- above the broken stick values (2.93 and 1.93, respectively).
cant differences between influent and effluent values of all Axis 1 explained 57% of the variance, and axis 2 explained
the wastewater parameters at the 0.05 significance level, an additional 21% of the variance. Together axis 1 and 2
except temperature and log(PO4-P) (Table 4). The individ- explained 78% of the variance.
ual ANOVAs with the largest F-statistics, and therefore the The farm digester effluent values have a positive influ-
largest contrast between influent and effluent concentra- ence on the location of axis 1, and the influent values have
tions are log(turbidity) (F-stat = 272) and log(COD) (F- a negative influence on axis 1. The effluent values are more
stat = 197). tightly correlated in multivariate space along axis 1 than
the influent values. Farm 2, which is the only farm digester
3.2. Water quality comparisons between individual farms that utilized cow manure as a wastewater source for the
entire study, had the highest positive influence on axis 2.
The data from each individual farm was compared to The influent and effluent data from the digester on Farm
examine differences between influent and effluent data 3 negatively influenced the location of axis 2. Farm 3 had
across the seven farms. For each wastewater variable, the a low retention time (16 days), a high number of pigs
effluent and influent data from each individual farm were (40), and solid separation prior to the digester entrance
averaged (Table 5), and these average influent and effluent (Fig. 3 and Table 2).

Table 5
Average influent and effluent data ± SE (n) for Taiwanese-model digesters at Farms 1–7
Farm 1 Farm 2 Farm 3 Farm 4 Farm 5 Farm 6 Farm 7
Influent temp. 27.6 ± 0.3 (12) *H 28.4 ± 0.3 (12) *H 26.3 ± 0.3 (11) 23.8 ± 0.2 (11) 25.5 ± 0.2 (12) 25.7 ± 0.4 (10) 25.6 ± 0.4 (6)
(°C)
Effluent temp. 27.2 ± 0.2 (12) *H 27.8 ± 0.3 (12) *H 25.3 ± 0.1 (12) 25.2 ± 0.1 (12) 25.2 ± 0.2 (12) 25.8 ± 0.4 (8) 26.1 ± 0.2 (12)
(°C)
Influent cond 1.0 ± 0.2 (12) 1.3 ± 0.2 (12) 3.0 ± 0.2 (11) *H 2.4 ± 0.4 (11) 0.8 ± 0.1 (12) 1.0 ± 0.2 (10) 1.9 ± 0.7 (6)
(mS/cm)
Effluent cond 1.5 ± 0.02 (12) 0.9 ± 0.01 (12) *L 2.9 ± 0.1 (12) *H 1.6 ± 0.1 (12) 1.9 ± 0.1 (12) 2.1 ± 0.2 (8) 1.34 ± 0.1 (12)
(mS/cm)
Influent pH 8.4 ± 0.1 (12) *H 7.2 ± 0.1 (12) 6.8 ± 0.04 (11) 7.3 ± 0.2 (11) 7.1 ± 0.1 (12) 7.3 ± 0.1 (10) 7.2 ± 0.1 (6)
Effluent pH 7.1 ± 0.03 (12) *H 6.2 ± 0.1 (12) *L 6.8 ± 0.04 (12) 6.5 ± 0.03 (12) 6.7 ± 0.1 (12) 6.9 ± 0.1 (8) 6.4 ± 0.03 (12)
Influent BOD5 156 ± 25 (12) *L 373 ± 72 (12) 551 ± 74 (11) 887 ± 138 (11) 471 ± 91 (12) 275 ± 45 (10) 668 ± 62 (6)
(mg/L)
Effluent BOD5 115 ± 14 (12) 116 ± 12 (12) 87.3 ± 10 (12) 55.7 ± 8 (12) 126 ± 68 (12) 60.4 ± 10 (8) 102 ± 24 (12)
(mg/L)
Influent COD 957 ± 160 (12) *L 3220 ± 630 (12) 3580 ± 650 (10) 4220 ± 410 (11) 3600 ± 810 (12) 1310 ± 160 (10) 4680 ± 1050 (6)
(mg/L)
Effluent COD 418 ± 29 (12) 714 ± 83 (12) 294 ± 28 (12) 550 ± 121 (12) 194 ± 37 (12) 302 ± 69 (8) 802 ± 159 (11)
(mg/L)
Influent turb. 923 ± 201 (12) 2100 ± 730 (12) 1620 ± 450 (11) 2040 ± 310 (11) 2470 ± 560 (12) 561 ± 147 (9) 3590 ± 530 (6)
(NTU)
Effluent turb. 92.6 ± 5 (12) 279 ± 37 (12) 107 ± 15 (12) 195 ± 33 (12) 59.2 ± 10 (12) 153 ± 17 (8) 313 ± 57 (12)
(NTU)
Influent TSS 769 ± 135 (12) 2260 ± 660 (12) 2710 ± 610 (11) 3030 ± 410 (9) 3450 ± 720 (12) 1070 ± 230 (10) 2260 ± 390 (6)
(mg/L)
Effluent TSS 35.8 ± 10 (12) *L 241 ± 58 (12) 83.9 ± 16 (12) 964 ± 251 (12) 230 ± 76 (12) 411 ± 204 (8) 296 ± 92 (12)
(mg/L)
Influent PO4-P 8.61 ± 1.7 (12) 2.72 ± 2.2 (12) *L 32.8 ± 4.4 (11) 6.97 ± 1.1 (11) 17.7 ± 4.6 (12) 3.78 ± 1.0 (10) 27.0 ± 6.0 (6)
(mg/L)
Effluent PO4-P 17.5 ± 1.7 (12) 1.51 ± 1.5 (12) *L 26.4 ± 2.8 (12) 10.5 ± 1.1 (12) 8.11 ± 1.1 (12) *L 30.3 ± 5.2 (8) 18.6 ± 3.3 (12)
(mg/L)
Influent NH4- 60.8 ± 14 (12) 20.5 ± 2.6 (12) 106 ± 11 (11) *H 45.6 ± 13 (9) 14.2 ± 2.5 (12) 37.1 ± 12 (10) 37.8 ± 13 (6)
N (mg/L)
Effluent NH4- 84.5 ± 3.0 (12) 18.5 ± 1.8 (12) *L 129 ± 7.9 (12) 82.9 ± 3.9 (12) 96.7 ± 6.9 (12) 149 ± 8.6 (7) *H 42.0 ± 4.0 (12) *L
N (mg/L)
Influent TKN 286 ± 46 (11) 249 ± 21 (12) 490 ± 37 (2) 288 ± 4.9 (3) 306 ± 54 (3) 160 (1) 551 ± 230 (3)
(mg/L)
Effluent TKN 199 ± 26 (11) 108 ± 10 (12) 200 ± 5.5 (2) 219 ± 6.1 (3) 152 ± 29 (2) 228 (1) 177 ± 46 (6)
(mg/L)
*H and *L indicate the values are significantly higher or lower, respectively, than at least five of the other six farms based on Tukey–Kramer analyses.
5886 S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890

Table 6
Average percentage of methane (CH4) and hydrogen sulfide (H2S) in the
produced biogas are given for the digesters at Farms 1–7
Percent methane (CH4) Hydrogen sulfide (H2S) in ppm [%]
Farm 1 68.8% ± 11 (2) 84.2 ± 0.1 (2) [0.0084%]
Farm 2 61.7% ± 24 (2) 4.8 ± 3.4 (2) [0.0048%]
Farm 3 72.5% ± 0.4 (2) 84.4 ± 0.4 (2) [0.0084%]
Farm 4 68.4% ± 3 (2) 84.3 ± 0.1 (2) [0.0084%]
Farm 5 67.9% ± 0.5 (2) 0.37 ± 0.3 (2) [0.00004%]
Farm 6 63.9% ± 0.7 (2) 83.6 ± 0.6 (2) [0.0084%]
Farm 7 61.4% ± 0.7 (3) 42.5 ± 42 (3) [0.0043%]
The CH4 data are given in percentage ± SE (n), and the H2S data are given
in ppm ± SE (n), with the percent H2S in brackets.

Statistical analyses did not reveal any significant rela-


tionship between influent water quality and the percentage
of CH4 in the biogas (MLR p-value = 0.308; VIFs < 2.0).
Based on the PCA results and the variance inflation factors
(VIF), the variables included in the MLR were log(turbid-
ity), NH4-N, temperature, and pH. The MLR analysis
Fig. 3. Axis one and axis two of the digester principle component analysis between CH4 and influent log(TKN) and log(NH4-N)
(PCA) for influent and effluent data from Farms 1–7. Axis one has an was also not significant (p-value = 0.283; VIFs = 1.2).
eigenvalue of 5.66, and the percentage of variance explained is 57%. Axis The MLR was conducted with average influent wastewater
two has an eigenvalue of 2.18, and the percentage of variance explained is
variable values and CH4 concentrations from the digesters
21%.
at Farms 1–7.
Biogas production data is not presented here due to
Statistical analyses (MANOVA and ANOVA) revealed problems getting reliable data with standard gas meters.
that there were significant differences between the influent The small-scale digesters in this study did not have suffi-
and effluent data from the seven farm digesters. MANOVA cient pressure to allow for the biogas to pass through the
analyses showed significant differences between both the meters. Without meter interference, the farmers were able
influent data (p-value < 0.001; Wilks k < 0.001) and the to cook using the produced biogas for an average of six
effluent data (p-value < 0.001; Wilks k < 0.001) when com- hours a day.
pared across individual farm digesters (Table 4). The
individual ANOVAs show significant differences at the
0.05 significance level between the influent data from each 4. Discussion
farm digester for all environmental parameters, except
log(TKN), and significant differences between farm effluent The digesters in this study consistently reduced organic
data for all environmental parameters. The individual matter and solids by 79% to 91%, with COD decreasing
ANOVAs with the largest F-statistics were pH (24.0) for from 2968 mg/L to 472 mg/L. TKN decreased by 46% to
influent analysis and log(conductivity) (198) for effluent 166 mg/L. Mineralization occurred during the digestion
analysis. Subsequent Tukey–Kramer multiple comparison process, increasing the average NH4-N concentration by
analyses on the individual ANOVAs are detailed in Table 5. 78.3% to 82.2 mg/L. The effluent water had a significantly
lower oxygen demand and higher dissolved nutrient con-
3.3. Biogas quality – methane (CH4) concentration centration compared to the influent water, which increases
its usefulness as an organic fertilizer and decreases its
The average methane concentration from the seven organic loading.
digesters over the study period was 66.3% (663,000 ppm). Enough biogas was produced during the digestion pro-
Hydrogen sulfide levels at all seven farms were below cess to allow for an average of 6 h of cooking per day.
0.01% (100 ppm) (Table 6). The highest CH4 concentration The concentration of CH4 in the biogas averaged 66%.
was at Farm 3 (72.5%), which used swine manure and had The concentration of methane in the biogas from the
the highest TSS and COD loading rates (167 mg/L-day and plug-flow, Taiwanese-model digesters analyzed in this
221 mg/L-day, respectively) (Tables 2, 5 and 6). The lowest study were at levels comparable to the high-tech, com-
CH4 concentrations were found at Farm 2 (61.7%), which pletely stirred digester favored in developed countries
used dairy manure, but had the second highest TSS and (Chen and Shyu, 1996; Kayhanian and Rich, 1995; Par-
COD loading rates (117 mg/L-day and 167 mg/L-day, sons, 1984). This study revealed that small-scale agricul-
respectively), and Farm 7 (61.4%) which used swine man- tural digesters can produce methane at concentrations
ure, and had the second lowest TSS loading rate useful for cooking, while improving the quality of the live-
(40.6 mg/L-day). stock wastewater. Further studies need to be conducted on
S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890 5887

biogas production to fully understand the Taiwanese- digesters are not regularly emptied, as practiced in CSTRs.
model digestion systems. The digesters in this study were 1–8 years old and none had
failed previously due to solid accumulation. Although it is
4.1. Reductions in organic matter and solids difficult to understand mass balances in flow-through
digestion systems, the 2007 Protocol for Quantifying and
The digesters analyzed in this study had high, statisti- Reporting the Performance of Anaerobic Digestion Systems
cally significant reductions of organic matter and solids for Livestock Manures suggests monitoring fixed solids
(Tables 3 and 4). Seventy-nine to 91% of the influent con- (FS) and total phosphorus (TP) to determine the amount
centrations of BOD5, COD, TSS, and turbidity were of solid accumulation in digesters, since FS and TP are
removed, which is greater than or approximately equal to not transformed in the digestion process (USDA, 2007).
values reported for other Taiwanese-model digesters In future Taiwanese-model digesters studies, a better
(Chará et al., 1999; Esquivel et al., 2002; Pedraza et al., understanding of solid accumulation, and thus nitrogen
2001). Based on the PCA analysis, turbidity and COD have cycling, could be gained using FS and TP data.
the greatest influence on the data in multivariate space and In most previous Taiwanese-model digester studies,
have the least amount of variability among the collected TKN was not directly measured because of high cost and
effluent samples when compared to BOD5 and TSS error rates (Peters et al., 2003), but reporting TKN values
(Fig. 2 and Table 3). There is a large equipment cost asso- could provide insights that are not revealed by testing
ciated with conducting COD and turbidity tests, but the COD alone. Farms 1 and 3 had the highest TKN loadings
simplicity of these test yields a lower likelihood of human (25.3 and 30.2 mg/L-day, respectively), the lowest COD/
error (APHA, 1989; Crites and Tchobanoglous, 1998). This TKN ratios (3.3 and 7.3, respectively) and the highest
study showed that these four variables (COD, BOD5, TSS, CH4 concentration in the biogas (68.8% and 72.5%,
turbidity) are highly correlated and duplication may not be respectively).
necessary, unless environmental regulations specifically The higher TKN loadings at Farm 1 and 3 could have
require them. It has been recommended in the 2007 Proto- also led to higher ammonification rates, which would
col for Quantifying and Reporting the Performance of increase the pH levels of wastewater (Sterling et al., 2001)
Anaerobic Digestion Systems for Livestock Manures that and contribute to the increased methane quality at these
COD and total solids analyses be used in future digesters digesters (Table 5). Farm 1 had an effluent pH of 7.2, which
studies (USDA, 2007). was significantly higher than the pH levels at the other
farms. Farm 3 also had a high effluent pH (6.8). The opti-
4.2. Nitrogen cycling in taiwanese-model digesters mal pH is 6.4–7.6, with maximum methane quality occur-
ring above pH 7 (An and Preston, 1999; Parsons, 1984).
TKN is a measure of both total organic nitrogen and The pH dropped an average of 0.7 units during the diges-
NH4-N in wastewater. The TKN decreased by an average tion process due to the production of fatty acids during
of 45% in the digesters studied. The total nitrogen levels digestion (Botero and Hernández, 2005), but the higher
were likely reduced through microbial uptake and some NH4-N mineralization at Farms 1 and 3 could have kept
settling of solids within the digester. The PCA analysis the pH near optimal levels and led to the greater percentage
showed that TKN was positively correlated with organic of CH4 in the biogas.
matter and negatively correlated with NH4-N in multivar- Moderate concentrations of NH4-N can increase the pH
iate space (Fig. 2). As the organic matter decreased in the and increase CH4 production, but when NH4-N concentra-
digestion process, the TKN also decreased, but at a slower tions increase above a certain threshold, the activity of
rate because the organic nitrogen was converted to NH4-N methanogens can be inhibited. Methane production inhibi-
through mineralization (Crites and Tchobanoglous, 1998). tion has occurred with concentrations of NH4-N above
Denitrification also likely occurred, as the average NOx- 4920 mg/L, with 100% inhibition with NH4-N concentra-
N concentration decreased by 80.4% (from 0.92 to 0.18 mg/ tions above 8000 mg/L (Sterling et al., 2001; Sung and
L), but the loss of NOx-N is not reflected in the TKN Liu, 2003). The CH4 inhibition observed in these previous
analyses (Table 3). The Anammox process is an unlikely studies occurred at NH4-N concentrations considerably
explanation for the total nitrogen removal due to the low higher than the influent and effluent values found in this
NOx-N concentration. In addition, previous studies have study. Farm 6 had the highest concentrations of NH4-N
shown that the microorganisms responsible for the Anam- in the effluent (149 mg/L) and had a moderate percentage
mox process are less competitive than denitrifying organ- of CH4 in the produced biogas (63.9%). Farm 2 had the
isms in anaerobic digesters (Dong and Tollner, 2003). lowest (18.5 mg/L) NH4-N effluent concentration and a
Nitrification, converting the NH4 to NOx, is unlikely to low percentage of CH4 in the biogas (61.7%). In this study,
have occurred during the digestion process due to the low NH4-N concentration did not have a significant affect on
DO levels in the digester (>0.6 mg/L). methane quality (p-value = 0.104). Further studies need
Settling of solids might have also contributed to the to be conducted in order to determine if there is a relation-
decrease in TKN levels. Taiwanese-model digesters are ship between TKN, NH4-N, pH, CH4, and biogas
flow-through digesters, and therefore the contents of the production.
5888 S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890

4.3. Increases in dissolved nutrients during digestion capital costs to purchase the generator might negate the
economic value of these simple systems and complicate
NH4-N, conductivity, and PO4-P were highly correlated their minimalist management. Further studies need to be
in the PCA (Fig. 2). All farm digesters in this study had an conducted to determine if enough biogas is produced to
increase in NH4-N during digestion (Table 5). Conductivity provide an economic benefit.
increased in some digesters but decreased in others (Table
5). Conductivity is useful as a general parameter of the
strength of dissolved species in wastewater, but the domi- 5. Conclusions
nant species contributing to the conductivity are not iden-
tified. PO4-P influent concentrations were not statistically The digesters in this study were effective at generating a
different from effluent concentrations. Nitrogen is taken high quality biogas to meet the farmers’ cooking needs.
up by the microorganisms more readily than phosphorus, Organic matter and solids concentrations were consistently
and thus, the amount of nitrogen in the wastewater source reduced in the effluent waters and NH4-N concentrations
is more important than phosphorus to gauge microbial were increased. The methane production in these digesters
community stability (Tchobanoglous et al., 2003). NOx-N creates a number of indirect environmental and societal
concentrations in the digester studied were always less than benefits, including (1) a reduction in deforestation associ-
1.0 mg/L, which was expected due to the low DO levels ated with firewood collection, (2) less hours devoted to fire-
(>0.6 mg/L). It is recommended that NH4-N be monitored wood collection, (3) eliminating the need to purchase
in future digester studies, while PO4-P, conductivity, and propane for cooking, (4) less organic matter in effluent
NOx-N analyses were found to be less important in charac- waters, (5) an organic liquid fertilizer, and (6) a reduction
terizing the digester effluent. in greenhouse gas emissions to the atmosphere.
The NH4-N concentration in the effluent determines the Through statistical and analytical analyses, the most
usefulness of the digester effluent as an organic fertilizer, important wastewater parameters that should be used for
and affects its potential impact on the quality of receiving characterization of the influent and effluent from Taiwan-
waters. Through this study, it was determined that the ese-model digester were determined. COD, turbidity,
effluent from Farm 6 would be more suitable to use as an NH4-N, pH, and TKN should be utilized in future studies
organic fertilizer due to the high NH4-N content, while for proper characterization of the digester wastewater. In
the effluent from Farm 2 (the bovine farm) would not be addition, the 2007 Protocol for Quantifying and Reporting
as useful as fertilizer, but would have less effect on aquatic the Performance of Anaerobic Digestion Systems for Live-
life if discharged into nearby waters. stock Manures recommends analyzing TP, FS, and total
solids (USDA, 2007). There were significant variations
among the wastewater influent and effluent values of the
4.4. Methane quality seven different digesters analyzed in the study, but these
variations did not significantly influence the methane con-
In this study, there were variations in the wastewater centrations, but might have influenced the quantity of bio-
influent concentrations (Table 5), the management styles, gas produced. The digester effluent concentrations
and the elevations of the digesters (Table 2), but all the exhibited less variation than the influent concentrations,
digesters were able to achieve CH4 concentrations greater which suggests that these systems have the ability to with-
than 60%, with an average concentration of 66%. While stand fluctuations in influent wastewater quality.
there might have been differences in the quantity of biogas The ability of the various small-scale digesters in this
produced, there was low variability in the CH4 concentra- study to treat a wide range of wastewater quality with vary-
tions (61.4–72.5%) in the biogas when compared to the ing management styles consistently is a testament to the
high variability of the wastewater influent concentrations design of these simple digesters. This research is very timely
(Tables 5 and 6). There was less variation in effluent water in that the majority of investigations on digesters have con-
quality when compared to influent water quality in the centrated on large-scale systems, when in fact, the over-
PCA, which suggests that these digesters had the ability whelming majority of digesters are small-scale systems
to withstand fluctuations in influent water quality (Fig. 3). located in the developing world. There is still much to be
In order to run a generator fueled with biogas, the CH4 learned from these simple systems.
concentration needs to be greater than 50% and the H2S Further studies need to be conducted on biogas produc-
concentration needs to be less than 1%. All of the digesters tion in Taiwanese-model digesters to access their ability to
in this study meet these minimum criteria to power a gen- generate electricity and compare their methane production
erator. The ability of a farm manager to use a digester in rate to the 0.3496 m3 per kg of COD destroyed seen in
combination with a generator depends on the digester size, large-scale digesters (USDA, 2007). Additionally, eco-
the daily production of biogas, and ability to store biogas nomic, environmental, and social analyses need to be con-
when the generator is not in use. The digesters in this study ducted to determine if combining these digesters with
may have the potential to be upgraded for electricity gener- electrical generators would be beneficial for the rural,
ation based on the percentage of methane produced, but small-scale farmers.
S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890 5889

Disclaimer Botero, R., Hernández, C., 2005. Manejo productivo de excretas en


sistemas ganaderos tropicales. In: Revista INFOHOLSTEIN, Aso-
ciación de Criadores de Ganado Holstein de Costa Rica, San Jose,
‘‘This report was prepared as an account of work spon- Costa Rica.
sored by an agency of the United States Government. Nei- Botero, R., Preston, T.R., 1987. Biodigestor de bajo costo para la
ther the United States Government nor any agency thereof, producción de combustible y fertilizante a partir de excretas. Manual
nor any of their employees, makes any warranty, express or para su instalación, operación y utilización, Centro para la Investiga-
implied, or assumes any legal liability or responsibility for ción en Sistemas Sostenibles de Producción Agropecuaria (CIPAV),
Cali, Colombia.
the accuracy, completeness, or usefulness of any informa- Chará, J., Pedraza, G., Conde, N., 1999. The productive water decon-
tion, apparatus, product, or process disclosed, or repre- tamination system: a tool for protecting water resources in the tropics.
sents that its use would not infringe privately owned Livestock Research for Rural Development 11 <www.cipav.org.
rights. Reference herein to any specific commercial prod- conlrrdnlrrd11n11ncha111.htm>.
uct, process, or service by trade name, trademark, manu- Chen, T.H., Shyu, W.H., 1996. Performance of four types of anaerobic
reactors in treating very dilute dairy wastewater. Biomass and
facturer, or otherwise does not necessarily constitute or Bioenergy 11, 431–440.
imply its endorsement, recommendation, or favoring by Collins, G., Woods, A., McHugh, S., Carton, M.W., O’Flaherty, V., 2003.
the United States Government or any agency thereof. Microbial community structure and methanogenic activity during
The views and opinions of authors expressed herein do start-up of psychrophilic anaerobic digesters treating synthetic indus-
not necessarily state or reflect those of the United States trial wastewaters. FEMS Microbiology Ecology 46, 159–170.
Crites, R., Tchobanoglous, G., 1998. Small and Decentralized Wastewater
Government or any agency thereof.’’ Management Systems. McGraw-Hill Companies, Inc., USA.
Dong, X., Tollner, E.W., 2003. Evaluation of Anammox and denitrifica-
tion during anaerobic digestion of poultry manure. Bioresource
Acknowledgements Technology 86, 139–145.
Erickson, L.E., Fayet, E., Kakumanu, B.K., Davis, L.C., 2004. Anaerobic
digestion. In: Carcass Disposal: A Comprehensive Review by the
This material is based upon work supported by the National Agricultural Security Center Consortium USDA APHIS
Department of Energy [National Nuclear Security Admin- Cooperative Agreement Project: Carcass Disposal Working Group.
istration] under award number (DE-FG02-04ER63834), National Agricultural Biosecurity Center, Kansas State University,
and the Ohio State University’s Targeted Investments in Manhattan, KS (Chapter 7).
Excellence ‘Carbon-Water-Climate’ Project. We would like Erickson, L.E., Fung, D.Y.C., 1998. Handbook on Anaerobic Fermen-
tation. Marcel Dekker, New York.
to thank the Department of Energy (DOE) and the Ohio Esquivel, R.R., Méndiz y Cazern, M.D., Preston, T.R., Pedraza, G., 2002.
State University for funding this research, and the labora- Apectos importantes al introducir biodigestores en explotaciones
tory and DOE staff at EARTH University for their assis- lecheras a pequeña escala. Livestock Research for Rural Development
tance in the research, including Carlos Hernández, Jane 14 <www.cipav.org.conlrrdnlrrd14n13nviey143.htm>.
Yeomens, Jorge Vinicio Murillo Rojas, Marianela Castro Gowda, M.C., 1995. Rural wastewater management in a southern Indian
village – a case study. Bioresource Technology 53, 157–164.
Valverde, Herbert Arrieta, Carlos Araya, Luis Emilio Pin- Griffin, M.E., McMahon, K.D., Mackie, R.I., Paskin, L., 1998. Metha-
eda, and Gerardo Cedeño. We also wish to thank the stu- nogenic population dynamics during start-up of anaerobic digesters
dent workers at the Ohio State University and EARTH treating municipal solid waste and biosolids. Biotechnology and
University, including Melanie Miller, Blanca Rivas, Marco Bioengineering 57, 342–355.
Güilcapi, and Joaquı́n A. Vı́quez. We would also like to Huttunen, S., Lampinen, A. (Eds.), 2005. Bioenergy Technology Evalu-
ation and Potential in Costa Rica. University of Jyväskylä Printing
thank our partners at Central State University for their House, Jyväskylä, Finland.
assistance, including Sritharan Subramania and Bryan Jolliffe, I.T., 2002. Principal Component Analysis. second ed.. Springer,
Smith. Additional thanks are also extended to Richard New York.
Fortner, David Hansen, and Pat Rigby for their adminis- Kayhanian, M., Rich, D., 1995. Pilot-scale high solids thermophilic
trative guidance. anaerobic digestion of municipal solid waste with an emphasis on
nutrient requirements. Biomass and Bioenergy 8, 433–444.
Lettinga, G., Rebac, S., Parshima, S., Nozhevnikova, A., van Lier, J.B.,
Stams, A.J.M., 1999. High-rate anaerobic treatment of wastewater at
References low temperatures. Applied and Environmental Microbiology 65, 1696–
1702.
An, B.X., Preston, T.R., 1999. Gas production from pig manure fed at McCune, Medford, 1999. Multivariate Analysis of Ecological Data,
different loading rates to polyethylene tubular biodigesters. Livestock Version 4.20. MJM Software, Gleneden Beach, Oregon, USA.
Research for Rural Development 11 <www.cipav.org.conlrrdnlrrd11n McMahon, K.D., Zheng, D., Stams, A.J.M., Mackie, R.I., Paskin, L.,
11nan111.htm>. 2004. Microbial population dynamics during start-up and overload
An, B.X., Preston, T.R., Dolberg, F., 1997. The introduction of low-cost conditions of anaerobic digesters treatment municipal solid waste and
polyethylene tube biodigesters on small-scale farms in Vietnam. sewage sludge. Biotechnology and Bioengineering 87, 823–834.
Livestock Research for Rural Development 9 <www.cipav.org. Moog, F.A., Avilla, H.F., Agpaoa, E.V., Valenzuela, F.G., Concepcion,
conlrrdnlrrd9n2nan92.htm>. F.C., 1997. Promotion and utilization of polyethylene biodigester in
APHA, 1989. Standard Methods for the Examination of Water and smallhold farming systems in the Philippines. Livestock Research for
Wastewater. American Health Association, Washington, DC. Rural Development 9 <www.cipav.org.conlrrdnlrrd9n2nmoog92.htm>.
Berglund, M., Börjesson, P., 2006. Assessment of energy performance in Ong, H.K., Greenfield, P.F., Pullammanappallil, P.C., 2000. An opera-
the life-cycle of biogas production. Biomass and Bioenergy 30, 254– tional strategy for improved biomethanation of cattle-manure slurry in
266. an unmixed, single-stage, digester. Bioresource Technology 73, 87–89.
5890 S. Lansing et al. / Bioresource Technology 99 (2008) 5881–5890

Parsons, R.A., 1984. On-farm Biogas Production. Northeast Regional Sterling Jr., M.C., Lacey, R.E., Engler, C.R., Ricke, S.C., 2001.
Agricultural Engineering Service, Ithaca, New York. Effects of ammonia nitrogen on H2 and CH4 production during
Pedraza, G.X., Chará, J., Conde, N., Giraldo, S., Giraldo, L., 2001. anaerobic digestion of dairy cattle manure. Bioresource Technol-
Evaluación de los biodigestores en geomembrana (PVC) y plástico de ogy 77, 9–18.
invernadero en clima medio para el tratamiento de aquas residuals de Sung, S., Liu, T., 2003. Ammonia inhibition on thermophilic anaerobic
origin porcino. Livestock Research for Rural Development 14 digestion. Chemosphere 53, 43–52.
<www.cipav.org.conlrrdnlrrd14n11nPedr141.htm>. Tchobanoglous, G., Burton, F.L., Stensel, H.D., 2003. Wastewater
Peters, J., Combs, S., Hoskins, B., Jarman, J., Kovar, J., Watson, M., Engineering: Treatment and Reuse. McGraw Hill, New York,
Wolf, A., Wolf, N., 2003. Recommended Methods of Manure NY.
Analysis. Cooperative Extension Publishing, Madison, WI. Thy, S., Preston, T.R., 2003. Effluent from biodigesters with different
Sophea, K., Preston, T.R., 2001. Comparison of biodigester effluent and retention times for primary production and feed of Tilapia (Oreochr-
urea as fertilizer for water spinach vegetable. Livestock Research for omis niloticus). Livestock Research for Rural Development 15
Rural Development 13 <www.cipav.org.conlrrdnlrrd13n 16nkean136. <www.cipav.org.conlrrdnlrrd15n19nsant159.htm>.
htm>. USDA Rural Development, Association of State Energy Research and
Sophin, P., Preston, T.R., 2001. Effect of processing pig manure in a Technology Transfer Institutions (ASERTTI), and USEPA AgSTAR
biodigester as an input for ponds growing fish in polyculture. Program. 2007. A Protocol for Quantifying and Reporting the
Livestock Research for Rural Development 13 <www.cipav. Performance of Anaerobic Digestion Systems for Livestock Manures.
org.conlrrdnlrrd13n16npich136.htm>. <http://www.epa.gov/agstar/pdf/protocol.pdf>.

Das könnte Ihnen auch gefallen