You are on page 1of 12

International Journal of Environmental Science and Technology (2019) 16:2721–2732

https://doi.org/10.1007/s13762-018-1997-z

ORIGINAL PAPER

Measurement, control, and modeling of ­H2S emissions from a sewage


treatment plant
M. Baawain1 · A. Al‑Mamun1 · H. Omidvarborna1 · I. N. Al‑Sulaimi2

Received: 26 May 2017 / Revised: 24 October 2017 / Accepted: 15 September 2018 / Published online: 22 September 2018
© Islamic Azad University (IAU) 2018

Abstract
Hydrogen sulfide is reported as one of the most common odor-producing compounds from sewage treatment plants. The
literature review revealed that very little information is available on the emission of ­H2S from different units of sewage
treatment plants. Besides, the removal efficiency as well as dispersion studies of H ­ 2S should be explored more. Therefore,
the purposes of this study were to measure the emissions of H ­ 2S, evaluate the efficiency of odor control unit, and explore
the dispersion potential of ­H2S from a sewage treatment plant in Muscat, Sultanate of Oman. Site survey analysis revealed
four sources (tanker discharge area, grit/sand and fog removal channels, pre-aeration basin, and screen building) of H ­ 2S in
the plant premises. Additionally, measurement analysis showed that the efficiency of odor control unit, which treated the
emissions from all sources, except the tanker discharge facility, was 90.2%. The Gaussian plume dispersion modeling studies
demonstrated that the allowable concentration of dispersed H ­ 2S (1 ppm) was observed > 30 m from the identified sources.
Although the combined outlet of the odor control unit was greater than 1 ppm, the plant is complying with the current envi-
ronmental specifications at the boundaries (≤ 1 ppm) set by the environmental authorities. Hence, a set of recommendations
is suggested to improve the efficiency of the system to resolve the odor emission from the plant.

Keywords  Air pollution modeling · Gaussian plume dispersion · Hydrogen sulfides · Monitoring and assessment · Sewage
treatment plants

Introduction are headaches, eye irritation, and respiratory-related prob-


lems. In addition, odor generates negative economic effects
Various odorous compounds are emitted from industrial in the society by reducing the property price and tourism
sources, including organic/inorganic pollutants and par- prospects due to aesthetic nuisance (Jehlickova et al. 2008;
ticulates. Odorous compounds arise from many sources, Sucker et al. 2009; Zarra et al. 2012). That is the reason
mainly in developing countries (Huang et al. 2016), such why odor-related complaints from the communities close to
as non-engineered landfills and improper-managed sewage such sources have been constantly increasing and becoming
treatment plants (STPs) (Iranpour et al. 2005; Zarra et al. a significant source of environmental annoyance during the
2008). In such sites, trace amount of odorous compounds, last decades (Van Harreveld 2001, Karageorgos et al. 2010;
such as hydrogen sulfide (­H2S), causes unpleasant nui- Latos et al. 2011).
sances for plant operators and people in the neighborhood. H2S, a colorless and flammable gas with a characteristic
The major health effects related to odorous compounds odor of rotten eggs, is the major reported compound close to
STPs, which can be detected at very low concentrations (Koe
Editorial responsibility: M. Abbaspour. 1985; Riva and Elettronica 1992; Simms et al. 2000; Kong
et al. 2015). ­H2S has a low odor threshold limit (Table 1),
* M. Baawain
msab@squ.edu.om and its smell can be detected even below 1 ppm. The odor
detection threshold limit of H­ 2S (approximately 0.0004 ppm)
1
Department of Civil and Architectural Engineering, College is the lowest among some other odorous substances as sum-
of Engineering, Sultan Qaboos University, P.O. Box 33, marized in Table 1 (Bhawan and Nagar 2008). ­H2S can be
Al‑Khodh, 123 Muscat, Oman
toxic and even fatal, if both humans and animals are exposed
2
Al‑Ansab Sewage Treatment Plant, Haya Water, in greater than 2000 ppm within seconds or minutes (rapidly
P.O. Box 1047, Al Khuwair, 133 Muscat, Sultanate of Oman

13
Vol.:(0123456789)

2722 International Journal of Environmental Science and Technology (2019) 16:2721–2732

Table 1  Details of offensive odorous compounds emitted from industries (Capelli et al. 2009; Li et al. 2015)
Compound/odorant Formula Major health effects Odor detec-
tion threshold
(ppm)

Inorganic compounds
Ammonia NH3 Nose and throat irritation, bronchiolar and alveolar edema, and airway destruction 17.0
Chlorine Cl2 Eye/skin/airway irritation, sore throat, and cough 8.0E − 02
Hydrogen sulfide H2S Nausea, headaches, delirium, disturbed equilibrium, tremors, convulsions, skin and 4.7E − 04
eye irritation
Ozone O3 Asthma, chest pain, coughing, shortness of breath, and throat irritation 0.5
Sulfur dioxide SO2 Irritates the nose, throat, and airways to cause coughing, wheezing, shortness of 2.7
breath, or a tight feeling around the chest
Acids
Acetic acid CH3COOH The effects of acetic acid vapor include irritation and damage of the eyes, nose, throat, 1.0
and lungs
Butyric acid CH3(CH2)2COOH 1.2E − 01
Propionic acid (CH2)2COOH 2.8E − 02
Amines
Methylamine CH3NH2 Commonly irritates the nose and throat causing coughing and wheezing. 4.7
Dimethylamine (CH3)2NH 3.4E − 01
Trimethylamine (CH3)2N 4.0E − 04
Ethylamine C2H3NH2 2.7E − 01

absorbed through the lungs) (EPA 2003; Pavilonis et al. liquid solvent (Burgess et al. 2001). There are two types of
2013). absorption system available in sewage treatment plants: the
H2S occurs naturally in crude petroleum and natural gas. chemical absorption system (e.g., wet/chemical scrubber)
However, ­H2S is also produced by bacterial breakdown of and biological absorption system (e.g., bio-filters and bio-
organic materials and human/animal wastes (e.g., sewage). trickling filters) (Lebrero et al. 2011). Some of the STPs in
In STPs, sulfur exists as either organic sulfur from feces crowded locations have utilized wet scrubbers for their off-
or inorganic sulfur from the sulfate ion (SO−2
4 ). In typical gas treatment and typically achieved 98% of odor removal
domestic sewage, microbial reduction of SO−24 is the domi- (Sanchez et al. 2006). Such methods have already diluted
nant mechanism of H ­ 2S formation (Bentzen et al. 1995). the odor to the threshold units. The chemical wet scrubber
Besides, ­H2S can be formed whenever sulfur-containing is one of the most common odor treatment technologies used
compounds are exposed to organic materials at high tem- in the STPs. Typically, such scrubbers are highly effective
peratures (Sulfide 1981). A prevailing mechanism of ­H2S in treating H
­ 2S odor by using a chemical solution capable of
formation is shown in reaction 1. absorbing large amounts of the target gas (Davidson 2003;
Lebrero et al. 2011). Since most STP odors are sulfide based,
Anaerobic microorganisms
(no dissolved oxygen)
in scrubber towers, sodium hypochlorite and sodium hydrox-
SO−24
+ Organic matter ⟶ H2 S + CO2 ide (bleach and caustic soda, respectively) are commonly
(Re. 1) used for odor removal (Lebrero et al. 2011). Moreover, due
Odors are considered as the main causes of nuisances to the importance of the task, the efficiency of the removal
experienced by the exposed populations. Preventing the process should be monitored regularly.
release of odor from a source is the direct way of control- Air pollution models are regularly used to estimate envi-
ling human exposure to odor. The resultant exposure to a ronmental impact assessments and environmental risk anal-
specific odor can be mainly quantified by the integral effects ysis. Such models cover a wide-ranging scale either from
of emission sources, pathway of dispersion, and character- local or global modeling requirements (Baawain et al. 2017).
istics of receptor. Thus, the reduction of human exposure Many of the available algorithms in the air quality mod-
to an odor needs either eliminating any of these compo- els are based on simplified physical models of the various
nents or separating the connection among them (Southwood dispersion processes in combination with some empirical
et al. 2010; Gujarat Maritime Board 2012). Absorption is a data (Macdonald 2003). Gaussian plume dispersion (GPD)
treatment/elimination method in which one or more compo- is one of the most important models that is extensively used
nents of odor are transferred from the gaseous phase to the to estimate the concentration of air pollutants. GPD model

13
International Journal of Environmental Science and Technology (2019) 16:2721–2732 2723

is a mathematically driven model considering simultaneous 2014a). The plant receives and treats a proportion of tanker
advection and diffusion of gaseous molecules/particles in sewage and pumped effluents from a central pump station
air media. The extreme advantage of GPD models is that (CPS). The current flow from the tanker is 25,000 m3/day,
the models are extremely fast because a single formula is and that from the CPS (a residential area) is 30,000 m3/day
used for every receptor point. However, their fast response (Baawain et al. 2014b). The STP has an inlet facility (head-
depends on the following simplifications (Brusca et  al. work) comprising two covered pre-aeration lagoons, three
2016): constant emission rates, negligible diffusion in the fine screens (3 mm) within open channels, but contained
downwind direction, homogeneous horizontal meteorologi- within an inlet building and twin covered aerated grit and
cal conditions over the space modeled, no wind sheer in the sand removal lagoons that their emissions are sent to OCU to
horizontal/vertical planes, non-reactive gases/aerosol, no be treated. For biological treatment, the flow passes through
deposition/reaction with the surface from the plume, and four activated sludge lagoons and eight downstream sets of
Gaussian distributions in the crosswind and vertical direc- membrane biological reactor units. Sludge treatment com-
tion. This model has been widely implemented in study- prises aeration of surplus activated sludge with a covered
ing the dispersion of odors from different industrial sources holding tank, followed by thickening with gravity belt thick-
(Lehtinen et al. 2012; Llavador Colomer et al. 2012). eners. The OCU is designed as a chemical wet scrubber.
Most of the previous studies have been carried out to Figure 1 shows an aerial picture of the STP.
identify the resultant effects of multiple odor-producing
compounds. However, instead of utilizing extensive odor H2S sampling methodology
removal instruments, the current study enables to produce
knowledge on higher removal efficiency. In light of that, this There are a number of odor sources within the STPs, which
study aims to (1) investigate the potential sources and con- together or independently result in significant emission of
centration levels of ­H2S emissions from an STP; (2) evaluate odors. These sources and their corresponding sampling
the efficiency of the odor control unit or OCU (chemical wet points are shown in Fig. 2.
scrubber); (3) compare the actual receptor data with those
of the GPD model; and (4) identify the dispersion behav- 1. Tanker discharge area (open manholes, vents, and tanker
ior of the emitted H­ 2S using data acquisition receptor. The connections) as shown in Fig. 2a.
samplings for this study were carried out during January 5 2. Covered pre-aeration basins (Fig. 2b). The air from the
to March 25, 2014 in Al-Ansab STP, Sultanate of Oman. basin is ducted into the OCU.
3. Screen building as presented in Fig. 2c.
4. Grit/sand and fog removal channels (Fig. 2d).
Materials and methods
H2S was measured by using a portable multi-gas detector
Site location and description (QRAE-II, RAE Systems, US) designed to provide real-time
exposure measurement of H ­ 2S emissions using an electro-
Al-Ansab STP, a new plant in 8 km southwest of the capital chemical sensor. The detection range of the instrument for
district of Muscat, is currently owned and operated by Haya ­H2S was 0–100 ppm. The instrument was calibrated with
Water Company, which was commissioned in 2010 with an the standard protocols provided by the manufacturer prior
average treatment capacity of 55,000 m3/day (Baawain et al. to the tests. Each measurement was conducted three times a

Fig. 1  Aerial picture of the STP


obtained from Google Maps

13

2724 International Journal of Environmental Science and Technology (2019) 16:2721–2732

Fig. 2  Sources of ­H2S emissions: a tanker discharge area, b covered pre-aeration basins, c screen building, and d aerated grit/sand and fog
removal channels

day in order to check the repeatability of the instrument and


reliability of the ­H2S emission data.

Chemical wet scrubber

The OCU serves the headwork area of the STP and con-
sists of two identical parallel-trains cross-flow chemical wet
scrubbers. The OCU was designed to receive 165,450 m3/
hr wasted air volume. Al-Ansab STP has to comply with the
current ­H2S limits as designed (≤ 1 ppm at the plant bounda-
ries). In order to evaluate the performance of the OCU, ­H2S
emissions from the inlet and outlet points of the unit were
measured.

Fig. 3  Gaussian air pollution dispersion plume


Meteorological conditions

The meteorological conditions used in this study were Table 2  Dispersion coefficients based on Pasquill–Gifford classifica-
tions
based on the same period in 2013. The air temperature was
between 21.3 and 33.1 °C, and the mean temperature was Zref = 1.0 m 𝜎z = Dz X az 𝜎y = Dy X ay
28 °C. The atmospheric relative humidity varied from 47 Stability Dz az Dy ay
to 78% (the mean relative humidity was about 64%). The
Extremely unstable 0.61 0.83 0.58 0.88
wind direction was the mean direction of the year as 60°
weather condition
(mean wind speed, 6 m/s), which was northeast toward the
Moderately stable 0.31 0.58 0.09 0.88
expressway and residential areas. weather condition

Analytical dispersion model Emission dispersion from the tanker discharge


facility
Dispersion of air pollutants in the ambient atmosphere is
simulated by atmospheric dispersion models, which are For the tanker discharge facility, the concentration of ­H2S in
used to estimate/predict the downwind concentration of this premise was estimated as emission rate. The dispersion
air pollutants from pollutant sources as illustrated in Fig. 3 coefficients were calculated based on Pasquill–Gifford stabil-
(Sriram et al. 2006). The dispersion model has various forms ity classification as a function of distance from the source as
depending on the types, source, emission rate, stack height, demonstrated in Table 2. The ­H2S concentrations for the two
assumption, etc. In this study, the GPD model was imple- weather conditions were calculated by the following GPD
mented to estimate the downwind concentrations of ­H2S equation (assuming y = 0 and z = 1.0 m):
from the major sources. One source was the emissions from [ ( )]
the tanker discharge area (untreated H­ 2S), and the second Q − 12 y2 2
+ z2
source was the OCU outlet. The dispersion of ­H2S emissions C(x, y, z) = e 𝜎y2 𝜎z
(1)
2𝜋uc 𝜎y 𝜎z
was analyzed under two different weather conditions, mod-
erately stable and extremely unstable weather conditions.

13
International Journal of Environmental Science and Technology (2019) 16:2721–2732 2725

where x is distance in wind direction, y is lateral distance,


and z is vertical distance; C = H2S concentration at the
atmosphere at specific distance x (μg/m3), Q = source emis-
sion rate (Kg/s), σ = Pasquill–Gifford values in y and z direc-
tions (Macdonald 2003), and uc = corrected horizontal wind
speed (m/s) calculated by:
( )p
uc z
= (2)
uref zref

z = Roughness height; zref denotes a reference value (for cit-


ies = 1.0), uref = horizontal wind speed at zref, and p = cor-
rected wind speed profile for each stability class. The p value
increases with increasing surface roughness.

Emission dispersion from the chemical wet scrubber


outlet Fig. 4  Effective stack height in the plume

Similarly, ­H2S concentration at the OCU outlet was also


estimated as the emission rate under the two weather con- Step 2 Compute the stable crossover point (ΔTstable) by:
ditions. The dispersion coefficients were calculated based
on Pasquill–Gifford stability classifications (Table 2). The  � √�
following form of the GPD model was used to estimate the ΔTstable = 0.019582 Ts vs s (5)
dispersion of H
­ 2S from the source, where height of the stack
is included.
where T s  = stack exhaust temperature (K),
[ ( )]
− 12 y2
+ (z−h)
2 ѵs = stack exhaust velocity (m/s), and s = stability
Q
C(x, y, z;h) = e 𝜎y2 𝜎z2
(3) parameter ­(s−2)
2𝜋uc 𝜎y 𝜎z
Step 3 Determine the dominated plume, if:
where C (μg/m3), Q (Kg/s), uc (m/s), and P were calculated ( ) then
Ts − Ta > ΔTstable ⇒ buoyancy-dominated
the same way as Eq. 1; h = the effective stack height, which
plume
is the summation of the stack height and plume rise as illus-
trated in Fig. 4. ( ) then
 Ts − Ta < ΔTstable ⇒ momentum-dominated
plume
Under moderately stable weather condition Step 4 Calculate the dominated plume flux (where ds is the
stack diameter):
The estimation of ­H2S emissions from the OCU outlet
should be determined, whether the plume is buoyancy domi- 
nated or momentum dominated. This was achieved by the ( ) ( )2
T − Ta d
following procedure: Fb = g s vs s (6)
Ts 2

Step 1 Calculate the stability parameter (S) by Eq. 4 as


follows: 
( )( )2
2 Ta ds
Fm = vs (7)
Ts 2
 ( )
dT∕dz
S=g (4)
Ta
where Fb and Fm are buoyancy and momentum
where g = acceleration due to gravity (9.81 m/s2), Ta = ambi- flux, respectively.
ent air temperature (K), dT/dz = potential temperature gradi- Step 5 Calculate the effective stack height:
ent at stack top (K/m) = 0.05 for moderately stable weather
condition. For buoyancy-dominated plume:

13

2726 International Journal of Environmental Science and Technology (2019) 16:2721–2732

( ) then
 Ts − Ta < ΔTunstable ⇒ momentum-dominated

( )1∕ plume
Fb 3
h� = 2.6 (8) Step 4 Calculate the effective stack height:
Uc s
For buoyancy-dominated plume

For momentum-dominated plume: 


⎛ 3∕4 ⎞
 ⎜F ⎟
h = 21.425⎜ b ⎟ Fb < 55

(13)
� �1∕ ⎜ Uc ⎟
3
Fm (9) ⎝ ⎠
h� = 1.5 √
Uc s

Step 6 Include the effective height in the stack height in 


⎛ 3∕5 ⎞
GPD equation ⎜F ⎟
h� = 38.71⎜ b ⎟ Fb ≥ 55 (14)
⎜ Uc ⎟
⎝ ⎠

 �
h = hs + h (10)
For momentum-dominated plume

Under extremely unstable weather condition  ( )


vs
(15)

h = 3ds
For the unstable plume condition, the following modeling Uc
procedure was applied.
Step 5 Include the effective height of the stack height in
the GPD equation:
Step 1 Calculate the stability parameter (S) by Eq. 4:

where dT/dz is equal to 0.01 (K/m) for extremely



unstable weather condition. h = h s + h� (16)
Step 2 Compute the unstable crossover point (ΔT
unstable):
Results and discussion
Calculate buoyancy flux for buoyancy-dominated
Emissions of ­H2S from the sources
plume, if
As shown in Fig. 5, site survey results during January to March

( )1∕ 2014 revealed the sources with the highest impact on odor
vs 3
ΔTunstable = 0.0297Ts Fb < 55 (11) emissions. From the site survey, ­H2S emission from the tanker
ds2 discharge facility was detected to be very high (100 ppm)
compared to other sources. The root causes of odor were the
development of septicity and the intrinsic nature of the tanker
 raw sewage. Thus, because of the sewage nature, the tanker
( )1∕ discharge facility emitted the highest level of H­ 2S among the
v2s 3
ΔTunstable = 0.00575Ts Fb ≥ 55 (12) other sources in the STP premises. Site survey showed that
ds the tanker connections involved during the discharge process
release high amounts of ­H2S to the ambient. The aerated grit/
Step 3 Determine the dominated plume, if: sand and fog removal channels, screen building, and pre-aera-
tion basin were placed as the other sources of H ­ 2S emissions.
( ) then
Ts − Ta > ΔTunstable ⇒ buoyancy-dominated The minimum and maximum values of H ­ 2S emissions from
plume the aerated grit/sand and fog removal channels were 29 ppm
and 68 ppm, respectively. Open weir results in such variations

13
International Journal of Environmental Science and Technology (2019) 16:2721–2732 2727

(a) Tanker discharge area Aerated grit/sand and fog removal channels Ambient T [⁰C]
100 40

35

80

30

25
Concentraon [ppm]

60

Temperature [ºC]
20

40
15

10

20

0 0
1-07

1-12

1-14

1-19

1-21

1-26

1-28

2-02

2-04

2-09

2-11

2-16

2-18

2-23

2-25

3-02

3-04

3-09

3-11

3-16

3-18

3-23

3-25
1-05

Date (2014)

(b)
2.5
Pre-aeration basin Screen building

2.0
Concentration [ppm]

1.5

1.0

0.5

0.0
1-12

3-11
1-05

1-07

1-14

1-19

1-21

1-26

1-28

2-02

2-04

2-09

2-11

2-16

2-18

2-23

2-25

3-02

3-04

3-09

3-16

3-18

3-23

3-25

Date (2014)

­ 2S emissions from different sources: a tanker discharge area, combined outlet of chemical wet scrubber and aerated
Fig. 5  Concentrations of H
grit/sand and fog removal channel in addition to ambient temperature, b pre-aeration and screen building

in ­H2S emissions from the channels. The open weir prevents flow, which increases the emission of H ­ 2S to the atmosphere.
the skimmed fats, oil, and grease from mixing with the sew- Additionally, the air bubbles during the aeration process might
age flow. The open weir creates turbulence in the sewage intensify emissions of H­ 2S through the opening weir. Besides,

13

2728 International Journal of Environmental Science and Technology (2019) 16:2721–2732

changes in the ambient temperature during the study period the unit ranged between 65 and 170 ppm, while the maxi-
might lead to higher emissions of ­H2S from the sources. mum value as outlet was less than 40 ppm. Thus, the H ­ 2S
The measured H ­ 2S emissions in the pre-aeration basin removal efficiencies by using chemical wet scrubbers were
were less than 1 ppm as shown in Fig. 5b. This narrow range between 80 and 96% as plotted in Fig. 6. The average effi-
was attributed to the installation of a cover over the basin, ciency of the unit with some exceptions was 90.2%, while
which prevented the emissions to the atmosphere. Due to the efficiency during maintenance period was reduced to
the atmospheric changes during the sampling period, the 67%. The average removal efficiency achieved by chemi-
­H2S emissions from the screen building showed some fluc- cal wet scrubber in this study was significantly lower than
tuations. This was due to the fact that the screen building the value reported by Sanchez et al. (2006) using the same
was partially open and some portion of the emissions were device (~ 98%). The orientation of the chemical wet scrubber
released to the atmosphere. Besides, temperature changes used in this study is the reason for this difference. Because
and some possible leakages from the air ducting system may of the deterioration in liquor quality across the flow of air,
also lead to such variations. For example, increasing H ­ 2S the cross-flow orientation of gas inside the scrubber was
emissions during the last month of the study confirmed the inherently less efficient than the countercurrent orientation.
impact of atmospheric temperature on ­H2S emissions. The
­H2S emissions from the screen building ranged from 0.4 to Dispersion of ­H2S from the tanker discharge facility
2.3 ppm, which showed one order of magnitude higher than
that of the pre-aeration basin. The dispersions of ­H2S from the tanker discharge facility
were analyzed using the GPD model under two different
Removal efficiency of ­H2S using chemical wet weather conditions. The estimated results were also vali-
scrubber dated with the experimental data as shown in Fig. 7. The
measured ­H2S concentration after 2 m from the tanker dis-
The combined inlet and outlet concentration of ­H2S from the charge facility was as high as 100 ppm, which was consid-
OCU is shown in Fig. 6. The concentrations of ­H2S input to ered as the emission rate under the two weather conditions.

Combined inlet Combined outlet Removal efficiency


180 100

160 90

80
140

70
120
Concentration [ppm]

60 Efficiency [%]
100

50

80
40

60
30

40
20

20 10

0 0
2-16
1-05

1-07

1-12

1-14

1-19

1-21

1-26

1-28

2-02

2-04

2-09

2-11

2-18

2-23

2-25

3-02

3-04

3-09

3-11

3-16

3-18

3-23

3-25

Date (2014)

Fig. 6  Combined emissions of H
­ 2S from the inlet and outlet of the OCU in addition to its efficiency of removal

13
International Journal of Environmental Science and Technology (2019) 16:2721–2732 2729

The high emission rate from this source was due to the boundaries (x = 165 m). As indicated in Table 3, this value
release of H
­ 2S under septic condition. The dispersion results was attributed at 50 m away from the tanker discharge facil-
showed that the concentration of H ­ 2S was reduced sharply ity under the stable weather condition and 30 m away under
with the increase of distance from the tanker discharge facil- the unstable weather condition. The resultant concentrations
ity under different weather conditions. of ­H2S at the STP boundary under the unstable and stable
The GPD model showed that there is not a significant weather conditions were 0.05 and 0.16 ppm, respectively.
deviation between model predictions under different weather The short distance under the unstable weather condition
conditions, when the distance is greater than 30 m. However, might be ascribed to the dispersion due to the high vertical
predicted values very close to the source (< 30 m) under turbulence of wind. Nevertheless, the ­H2S concentration can
moderately stable weather conditions showed higher val- be detected at nights by the neighborhood when the weather
ues than those of moderately unstable weather conditions. conditions became stable.
The high rate of decreasing H ­ 2S concentration under unsta- Table 4 shows the comparisons between measured and
ble weather condition was due to higher turbulence under predicted values by the GPD model under stable weather
this condition. Table  3 shows the comparisons between conditions. The analysis of the results confirmed that the
the main disparities of the dispersion studies under two relative errors were between 11.7 and 20.8%, which could
weather conditions. Despite the identical source concentra- be attributed to the impact of involved holes used for waste
tions, the emission rate in the unstable weather condition drainage. Although the model is based upon many simpli-
(0.003 kg/s) was higher than that of stable weather condition fying assumptions, the error values are within the expected
(0.0001 kg/s). This was attributed to the high wind velocity limits of accuracy for most GPD models (Beychok 1995).
(turbulence) in unstable weather condition (3.1 m/s).
As per the current STP compliance regulations, the meas-
ured ­H2S concentration should be ≤ 1 ppm at the STP site

150

Stable weather conditions Unstable weather conditions


120
Concentration [ppm]

90

60

30

0
0 20 40 60 80 100
Distance [m]

Fig. 7  Dispersion of H
­ 2S emissions from the tanker discharge facility under two different weather conditions

Table 3  Comparison of Parameter Moderately stable weather Extremely


dispersion studies from tanker condition unstable weather
facility under two weather condition
conditions
Emission rate (Kg/s) 1.0E − 04 3.0E − 03
Corrected wind velocity (m/s) 1.58 3.08
H2S concentration at x = 2 m (ppm) 100 100
Distance at which ­H2S concentration ≤ 1 ppm (m) 50 30
H2S concentration at STP boundary (ppm) 0.16 0.05

13

2730 International Journal of Environmental Science and Technology (2019) 16:2721–2732

Table 4  Validation of H
­ 2S dispersion from the tanker discharge facil- unit in unstable weather condition. The concentration of
ity under stable weather conditions ­H2S was reduced suddenly to preferred value (≤ 1 ppm)
Distance (m) H2S conc. Predicted values Relative under unstable weather condition, while the stable weather
(ppm) (ppm) error condition was unable to disperse the concentration level
(%) below the favored value. The distance at which the con-
5 40 35.5 11.7 centration of ­H2S is attributed to 1 ppm was 8 m under
10 15 12.8 14.4 the unstable weather condition. However, 1 ppm of ­H2S
20 5.9 4.7 20.8 under the stable weather condition was estimated at 50 m.
Table 5 shows the dispersion of emissions from the OCU
under the weather conditions. Despite the same emission
Dispersion of ­H2S from the OCU Outlet rates, the reduction in H­ 2S dispersion under the unstable
weather condition might be attributed to the higher disper-
The combined ­H 2S concentration from the OCU outlet sion of H­ 2S at different directions under this condition.
(12.4 ppm) was used to estimate the emission rate under Moreover, the concentration of H ­ 2S at the STP boundary
two weather conditions. The emission rate from the out- (x = 150 m) under the unstable and stable weather condi-
let (6.0E − 04 ppm) was relatively lower than that of the tions was 0.005 and 0.17 ppm, respectively.
tanker discharge facility, since the released ­H2S had been Table 6 demonstrates the comparisons of the measured
already treated. Figure 8 shows the dispersion of ­H2S from and predicted values from the OCU outlet under the sta-
the OCU outlet versus distance in the wind direction. The ble weather condition. The results were predicted with an
same as stable weather condition, the concentration of acceptable range of 11.5–28.3%. Although the concentration
­H2S was decreased with increasing the distance from the of ­H2S was reduced to 1 ppm at far away from the sources,

15

Stable weather conditions Unstable weather conditions


12
Concentration [ppm]

0
0 25 50 75 100 125 150
Distance [m]

Fig. 8  Dispersion of H
­ 2S emissions from the OCU outlet under different weather conditions

Table 5  Comparison of Parameter Moderately stable Extremely


dispersion studies from the weather condition unstable weather
OCU outlet under two weather condition
conditions
Emission rate (Kg/s) 6.0E − 04 6.0E − 04
Corrected wind velocity (m/s) 8.6 6.8
Source height (m) = stack height plus effective height 28 20
H2S concentration at x = 1 m (ppm) 12.4 12.4
Distance at which ­H2S concentration ≤ 1 ppm (m) 50 8
H2S concentration at STP boundary (ppm) 1.7E − 01 5.0E − 03

13
International Journal of Environmental Science and Technology (2019) 16:2721–2732 2731

Table 6  Validation of ­H2S dispersion from the OCU outlet under the transfer coefficient and reach the best possible performance
stable weather conditions of the unit. Hence, thus far, it is concluded that the plant is
Distance (m) H2S conc. Predicted values Relative complying with the current standard specifications of ­H2S
(ppm) (ppm) error emissions set by the environmental authorities (≤ 1 ppm at
(%) the plant boundaries).
8 14 12.4 11.5
Acknowledgements  The authors would like to acknowledge the field
10 11 9.1 17.5
and logistic support from Haya Water Company. The project was sup-
20 4.6 3.3 28.3 ported by the Ministry of Environment and Climate Affairs, Muscat,
Sultanate of Oman (Project Code: CR/DVC/CESAR/15/01).

the ambient air surrounding the STP was being contami-


nated with a relatively high concentration of H
­ 2S, especially
the area in close proximity to the tanker discharge facility. References
Therefore, the environmental authorities should set a new
Baawain MS, Al-Omairi A, Choudri BS (2014a) Characterization
concentration level for ­H2S emissions either at the outlet of of domestic wastewater treatment in Oman from three different
each unit or at site boundaries. regions and current implications of treated effluents. Environ
Monit Assess 186(5):2701–2716
Baawain MS, Al-Jabri M, Choudri BS (2014b) Characterization of
domestic wastewater sludge in oman from three different regions
Conclusion and recommendations for alternative reuse applications. Iran J
Public Health 43(2):168–177
Odor emissions from the STPs are considered as a significant Baawain M, Al-Mamun A, Omidvarborna H, Al-Jabri A (2017) Assess-
ment of hydrogen sulfide emission from a sewage treatment plant
environmental annoyance due to their direct impacts on both
using AERMOD. Environ Monit Assess 189(6):263
the human health and economy. In this study, the potential Bentzen G, Smith AT, Bennett D, Webster NJ, Reinholt F, Sletholt E,
sources of ­H2S emissions from a STP were examined and the Hobson J (1995) Controlled dosing of nitrate for prevention of
dominant mechanisms of its dispersion were explored. In the H2S in a sewer network and the effects on the subsequent treat-
ment processes. Water Sci Technol 31(7):293–302
STP premises, the main contributors of H ­ 2S emissions were
Beychok MR (1995) Air dispersion modeling: keep predicted values
identified as the tanker discharge facility (100 ppm) followed in perspective. Environ Eng World 1(6):16–18
by aerated sand/grit and oil removal channels (68 ppm) Bhawan P, Nagar A (2008) Guidelines on odour pollution & its control.
and screen building (2.3 ppm). All of the H ­ 2S emissions, Central Pollution Control Board, Ministry of Environmental &
Forests, Govt. of India, Delhi
except the tanker discharge facility, were ducted to the OCU
Brusca S, Famoso F, Lanzafame R, Mauro S, Garrano AMC, Monforte
(maximum value of combined inlet = 170 ppm). The average P (2016) Theoretical and experimental study of gaussian plume
removal efficiency of ­H2S from the sources through the OCU model in small scale system. Energy Procedia 101:58–65
was observed approximately at 90.2%. Burgess JE, Parsons SA, Stuetz RM (2001) Developments in odour
control and waste gas treatment biotechnology: a review. Biotech-
The dispersion of ­H 2S from the two main sources
nol Adv 19(1):35–63
(tanker discharge facility and OCU) was determined using Capelli L, Sironi S, Del Rosso R, Céntola P (2009) Predicting odour
the GPD model with an acceptable error range in overall emissions from wastewater treatment plants by means of odour
(11.5–28.3%). GPD is the core of almost all regulatory dis- emission factors. Water Res 43(7):1977–1985
Davidson S (2003) Know thy scrubber. Water Environ Technol
persion models and is applied to a wide array of dispersion
15(3):48–51
scenarios. The model predicted that H ­ 2S emissions from the EPA (2003) Hydrogen sulfide (CASRN 7783-06-4) Toxicological
tanker discharge facility and the OCU were reduced with the review. Retrieved October 2017 from https​://cfpub​.epa.gov/ncea/
increase of distance from them. The allowable concentration iris/iris_docum​ents/docum​ents/toxre​views​/0061t​r.pdf
Gujarat Maritime Board (2012) EIA for common hazardous waste
of dispersed ­H2S (≤ 1 ppm) was definitely observed at 50 m
incineration at Alang, Bhavnagar Gujarat, final report
away from both the sources under test conditions. Although Huang S, Zhao X, Sun Y, Ma J, Gao X, Xie T, Xu D, Yu Y, Zhao
the concentration of ­H2S was reduced to less than 1 ppm at Y (2016) Spatial distribution of organic pollutants in industrial
plant boundaries, the odor had been detected under the calm construction and demolition waste and their mutual interaction on
an abandoned pesticide manufacturing plant. Environ Sci Process
weather condition outside the plant boundaries. Therefore,
Impacts 18(4):482–492
the new regulation requires that all emissions should be con- Iranpour R, Cox HH, Deshusses MA, Schroeder ED (2005) Literature
veyed to the OCU and combined emissions from the OCU review of air pollution control biofilters and biotrickling filters
outlet should be ≤ 1 ppm. However, the plant is allowed to for odor and volatile organic compound removal. Environ Prog
24(3):254–267
have 1 ppm or less at the fence until all odor sources are
Jehlickova B, Longhurst PJ, Drew GH (2008) Assessing effects of
fully connected to the OCU. This is supposed to be com- odour: a critical review of assessing annoyance and impact on
pleted as soon as possible. Furthermore, countercurrent flow amenity. In: Odours and VOCs: measurement, regulation and con-
arrangement in the OCU is recommended to intensify mass trol techniques, vol 31, p 139

13

2732 International Journal of Environmental Science and Technology (2019) 16:2721–2732

Karageorgos P, Latos M, Kotsifaki C, Lazaridis M, Kalogerakis N Pavilonis BT, O’Shaughnessy PT, Altmaier R, Metwali N, Thorne PS
(2010) Treatment of unpleasant odors in municipal wastewater (2013) Passive monitors to measure hydrogen sulfide near con-
treatment plants. Water Sci Technol 61(10):2635–2644 centrated animal feeding operations. Environ Sci Process Impacts
Koe LC (1985) Ambient hydrogen sulphide levels at a wastewater treat- 15(6):1271–1278
ment plant. Environ Monit Assess 5(1):101–108 Riva G, Elettronica O (1992) Spartan Environmental Technologies.
Kong X, Liu J, Ren L, Song M, Wang X, Ni Z, Nie X (2015) Identifica- Mentor, Ohio, USA
tion and characterization of odorous gas emission from a full-scale Sanchez C, Couvert A, Laplanche A, Renner C (2006) New compact
food waste anaerobic digestion plant in China. Environ Monit scrubber for odour removal in wastewater treatment plants. Water
Assess 187(10):1–14 Sci Technol 54(9):45–52
Latos M, Karageorgos P, Kalogerakis N, Lazaridis M (2011) Disper- Simms KL, Wilkinson S, Bethan S (2000) Odour nuisance and disper-
sion of odorous gaseous compounds emitted from wastewater sion modeling: an objective approach to a very subjective prob-
treatment plants. Water Air Soil Pollut 215(1–4):667–677 lem. Water Sci Technol 41(6):89–96
Lebrero R, Bouchy L, Stuetz R, Munoz R (2011) Odor assessment and Southwood M, Rowland J, Lockwood R, Horrocks D, Pointing J, Lon-
management in wastewater treatment plants: a review. Crit Rev ghurst P, Sneath R, Peirson S, Selwyn T (2010) Odour guidance
Environ Sci Technol 41(10):915–950 for local authorities. Department for Environment, Food and Rural
Lehtinen J, Giuliani S, Zarra T, Reiser M, Naddeo V, Kranert M, Bel- Affairs Nobel House, London
giorno V, Romain AC, Nicolas J, Sówka I, Wang KY (2012) Case Sriram G, Mohan NK, Gopalasamy V (2006) Sensitivity study of
studies for assessment, control and prediction of odour impact. In: Gaussian dispersion models. J Sci Ind Res 65(4):321–324
Belgiorno V, Naddeo V, Zarra T (eds) Odour impact assessment Sucker K, Both R, Winneke G (2009) Review of adverse health effects
handbook. Wiley, New york, pp 205–283 of odours in field studies. Water Sci Technol 59(7):1281–1289
Li J, Yin C, Huo F (2015) Chromogenic and fluorogenic chemosensors Sulfide H (1981) World Health Organization, Geneva
for hydrogen sulfide: review of detection mechanisms since the Van Harreveld AP (2001) From odorant formation to odour nuisance:
year 2009. RSC Adv 5(3):2191–2206 new definitions for discussing a complex process. Water Sci Tech-
Llavador Colomer F, Espinós Morató H, Mantilla Iglesias E (2012) nol 44(9):9–15
Estimation of hydrogen sulfide emission rates at several waste- Zarra T, Naddeo V, Belgiorno V, Reiser M, Kranert M (2008) Odour
water treatment plants through experimental concentration meas- monitoring of small wastewater treatment plant located in sensi-
urements and dispersion modeling. J Air Waste Manag Assoc tive environment. Water Sci Technol 58(1):89–94
62(7):758–766 Zarra T, Giuliani S, Naddeo V, Belgiorno V (2012) Control of odour
Macdonald R (2003) Theory and objectives of air dispersion modeling. emission in wastewater treatment plants by direct and undirected
Department of Mechanical Engineering, University of Waterloo, measurement of odour emission capacity. Water Sci Technol
Wind Engineering MME A, 474 66(8):1627–1633

13

You might also like