Next Article in Journal
Accurate Absorption Energy Calculations in Solution Using the Reference Interaction Site Model Self-Consistent Field Including the Constrained Spatial Electron Density Distribution
Previous Article in Journal
Benchmarking Free Energy Calculations in Liquid Aliphatic Ketone Solvents Using the 3D-RISM-KH Molecular Solvation Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Can a Wastewater Treatment Plant Power Itself? Results from a Novel Biokinetic-Thermodynamic Analysis

by
Mustafa Erguvan
1,* and
David W. MacPhee
2
1
Department of Mechanical Engineering, The University of Alabama, Tuscaloosa, AL 35401, USA
2
Department of Mechanical & Aerospace Engineering, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada
*
Author to whom correspondence should be addressed.
J 2021, 4(4), 614-637; https://doi.org/10.3390/j4040045
Submission received: 20 September 2021 / Revised: 13 October 2021 / Accepted: 14 October 2021 / Published: 21 October 2021
(This article belongs to the Section Engineering)

Abstract

:
The water–energy nexus (WEN) has become increasingly important due to differences in supply and demand of both commodities. At the center of the WEN is wastewater treatment plants (WWTP), which can consume a significant portion of total electricity usage in many developed countries. In this study, a novel multigeneration energy system has been developed to provide an energetically self-sufficient WWTP. This system consists of four major subsystems: an activated sludge process, an anerobic digester, a gas power (Brayton) cycle, and a steam power (Rankine) cycle. Furthermore, a novel secondary compressor has been attached to the Brayton cycle to power aeration in the activated sludge system in order to increase the efficiency of the overall system. The energy and exergy efficiencies have been investigated by varying several parameters in both WWTP and power cycles. The effect of these parameters (biological oxygen demand, dissolved oxygen level, turbine inlet temperature, compression ratio and preheater temperature) on the self-efficiency has also been investigated. It was found here that up to 109% of the wastewater treatment energy demand can be produced using the proposed system. The turbine inlet temperature of the Brayton cycle has the largest effect on self-sufficiency of the system. Energy and exergy efficiencies of the overall system varied from 35.7% to 46.0% and from 30.6% to 33.55%, respectively.

1. Introduction

Wastewater treatment has been particularly important for humanity since the Bronze age (CA 3200–1100 BC) although advanced technologies started in the early 1900s [1,2]. The United Nations Educational, Scientific and Cultural Organization published a detailed report in 2020 to illustrate how climate change will affect water scarcity in the coming decades. According to this report, more than 40% of mankind will suffer from water scarcity by 2050 due to climate change and a high increase in population [3].
The priority in wastewater treatment plants (WWTP) has historically been concerned with meeting certain effluent water standards with little concern for energy consumption [4]. Nevertheless, energy efficiency of WWTPs has been an important subject because of their high energy consumption. According to the US Environmental Protection Agency, 3–4% of total electricity use in the US is consumed by drinking water and wastewater systems [5,6]. A relationship between energy and water, the “water–energy nexus”, has been the focus of attention worldwide due to climate change, population increase and urbanization [7,8]. While there is a high water demand for power generation, WWTP consume large amounts of electricity in most countries [9]. Since the standards of the effluent discharge is strict in WWTPs, energy consumption to treat wastewater can be quite high [7].
There are several ways to treat wastewater using physical, chemical and biological methods. Activated sludge is a biological wastewater treatment method for a secondary wastewater, and it is the most broadly applied treatment technology for municipal and industrial wastewater today [10,11]. Aeration is the most important parameter affecting performance of the activated sludge process, and there are two popular aeration devices utilized in the activated sludges process, namely, surface aerators and blowers/compressors. While surface aerators were highly popular until 1990s, blowers and compressors are the leading technology in today’s activated sludge applications due to increased reliability and higher efficiency [12]. Here, compressed air is forced through membranes or perforated tubes in order to supply aeration basins with air bubbles, facilitating oxygen transfer from air to wastewater.
The purpose of the aeration process is to transfer oxygen from air to the wastewater so that carbonaceous matter may be consumed by organisms undergoing cellular respiration. Aeration is the most energy consuming process in WWTPs, and consumes 50% to 90% of electricity usage [12]. It is not surprising then that a significant number of studies have been conducted to investigate ways to increase performance of aeration devices [13,14,15]. Since compressors or blowers require external power to run, a different approach is envisioned in this study by adding a secondary compressor (also called “bleed air”) to the power cycle so as to eliminate the associated generator inefficiency in the turbine as well as to directly use the shaft power in the Brayton cycle, thereby avoiding motor inefficiency.
In addition to the aforementioned aeration studies, combined heat and power (CHP) systems have been proposed using the biogas generated from anaerobic digestion (AD) process by digesting the activated sludge in WWTPs [16]. Once wastewater is treated in a WWTP, a vast amount of sludge that comes from the primary and secondary treatment processes must be separated from the water, and the disposal of this sludge may represent up to 50% of the total operational cost [17]. There are many different methods to harness the bio-energy in sludge, such as, pyrolysis, gasification, dark fermentation and anaerobic digestion. While gasification converts biomass into a combustible gas mixture—mostly carbon dioxide and hydrogen—anaerobic digestion transforms the volatile suspended solids into biogas [18,19]. Shizas and Bagley [20] claim that the energy potential of a raw wastewater is 9.3 times higher than the required energy to treat it. Appels et al. [17] provide an intensive review to show the potential and principle of the anaerobic digestion of the activated sludge. Another review has been conducted by Shen et al. [21] to show how anaerobic digestion can contribute to WWTPs in the United States as an energy source. The authors claimed that less than 10% of the WWTP in the US produces biogas for beneficial use. Farahbakhsh et al. [22] perform a novel integrated WWTP–CCHP system in order to investigate the performance of the system from the points of view of energy and economy. The authors also introduced a significant number of studies who focus on the integration of CHP on WWTPs. While there are several different approaches to model the CHP system, such as transient simulation using the real gas equation of state [23], the ideal gas law approach is used in this study for simplicity.
Biogas is the final production of the AD process, and mainly consists of methane and carbon dioxide. Since methane has a high lower heating value, it is one of the most common fuels in CHP systems. Self-sufficiency (sometimes called net zero) energy models for WWTP have been an interesting topic as of late, with several studies investigating whether WWTPs can produce enough power to treat the wastewater itself; some claim that this is indeed possible [24,25,26,27,28]. While challenges and opportunities toward net-zero WWTP have been discussed by Shen et al. [21], several self-sufficient full-scale WWTP with AD of sewage sludge have been illustrated. The authors performed a case study and found that 126% of a WWTP’s electricity demand can be met from the WWTP–CHP system. Schwarzenbeck et al. [25] claim that a WWTP in Germany produced 113% of the electricity consumed for plant operations in 2005 using a gas engine. Nowak et al. [27] analyzed two WWTPs in Austria whether the plants can be self-sufficient. The authors found that 180% energy generation compared to energy needs is possible in these two plants using a biogas powered CHP system.
A theoretical net zero energy (NZE) model has been created by Peng et al. [29] to investigate the self-sufficiency of 20 WWTPs in China. The authors implemented CHP systems to each WWTP, and they claim that while 6 of the WWTPs can produce excess energy than required, eight of the WWTPs may achieve 100% energy self-sufficiency by modifying some parameters. The same model has been used in another study [29] to find the effect of temperature and chemical oxygen demand effect on self-sufficiency. It was found that while the self-sufficiency rate is 43.9% under the existing operation in the WWTP, it can be increased up to 110.6% by optimizing the substrate allocations. In 2011, a 700-kW capacity CHP system is integrated into Sheboygan Regional WWTP to produce power for the plant. The WWTP currently generates between 90 and 115% of electricity need in the plant as well as 90% of heating energy on-site [28].
Second law (exergy) analysis has been a common method to investigate energy conversion systems in order to identify locations and magnitude of exergy losses [30,31]. A theoretical study based on the anaerobic digestion of sewage from WWTP has been conducted by Safari and Dincer [32] in order to investigate the thermodynamic efficiencies of a multigeneration system. The authors obtained overall energy and exergy efficiencies as 63.6% and 40%, respectively. Ozdil et al. [33] conducted a detailed exergy and exergoeconomic analyses of a biogas powered electricity production in a WWTP. The authors found that the highest exergy destruction is obtained in the gas engine and an overall efficiency of 69.1%. Abusoglu et al. [34,35,36] performed several studies so as to investigate exergoeconomic analysis of a municipal WWTP in Gaziantep, Turkey. The authors conducted a very detailed exergy analysis for each subsystem. The exergy efficiencies of the primary, secondary, and overall systems there were found to be 53.4%, 14.8% and 34%, respectively.
In this study, thermodynamic analysis of a novel integrated system in a WWTP is conducted. Sludge, one of the final products of a WWTP, is digested in an anerobic digestor to produce biogas. Then, the produced gas is combusted in a biogas-powered gas turbine cycle in order to generate power. There are two main novelties involved in this study. First, this study incorporates a comprehensive analysis including both biological and thermal parameters for WWTP and cogeneration systems, which has not been done for this system before. Secondly, the proposed system includes a secondary compressor connected to the gas turbine cycle in order to provide required air for aeration in the WWTP at reduced system loss. Since the dominant energy consumers in a WWTP is the aerators, the system analyzed herein has the potential to vastly increase efficiency and reduce associated costs and emissions for WWTP operations.

2. Model Description

In this study, a mathematical model has been developed in order to investigate a multigeneration system for a wastewater treatment plant using a net-zero energy building model. The proposed system is an integration of wastewater treatment plant with a combined gas–vapor power cycle. Figure 1 indicates the schematic diagram of the proposed system.

2.1. WWTP

As can be seen from Figure 1, municipal wastewater first passes a bar screen process to remove large items in order to prevent damage to the equipment in the plant followed by the grit chamber to separate solid organic matter from the wastewater. After that, the suspended solids are removed by sedimentation in the primary clarifier. Since a lower biochemical oxygen demand (BOD) level is required to have a high-quality treated water, a secondary treatment is a must in order to reduce BOD concentration. There are various systems for wastewater treatment, and the most common method—activated sludge—is chosen here. The most crucial part of a WWTP from an energy standpoint is the aeration basin, and this process requires a significant amount of oxygen. A new approach here is applied to provide air to the aeration basins using a compressor which is connected to the gas power cycle in order to remove the generator inefficiency of the gas turbine and any motor inefficiency associated with a fan/blower. In a typical WWTP, the air is provided from a separate compressor or blower. Since the sludge from the secondary clarifier has a high volume because of high water content, a gravity thickener is used to increase the solid content. The sludge coming from primary and secondary clarifiers are blended in a blending tank, then the mixture is sent to the anerobic digester to stabilize the sludge. Anaerobic digestion is the most common method to stabilize the sludge, and it is a very efficient way to produce biogas [22]. Although biogas contains many different gas components, the biogas can be approximated by assuming a molar fraction of 60% CH4 and 40% CO2 [34]. Another product of the anaerobic digestion process is digestate, which is sent to a dewatering process in order to reduce the liquid volume. The final product here is called caked sludge, and it can be used as a fertilizer or gasified for syngas production.

2.2. Cogeneration Power Cycle

The bottom cycle in Figure 1 depicts the schematic diagram of the cogeneration power cycle. The produced biogas from the AD process is used as a fuel in a combustion chamber. Here, air is used as an oxidizer and the air fuel ratio is calculated based on the conditions of a complete combustion operation. The compressed air is heated in a regenerator (HX 1) in order to decrease the heat input requirements for the same power production [37]. Once combustion takes place, high-temperature exhaust gas is expanded in a gas turbine to generate power. Afterwards, exhaust passes through two heat exchangers to increase the temperature of air and steam, respectively. The second heat exchanger (HX 2) is used to combine a Rankine cycle with the Brayton cycle. The final temperature of the exhaust gas is set to a minimum value of 400 K in order to avoid corrosive sulfuric acid formation [38]. As explained before, a secondary compressor is attached to the same shaft in the Brayton cycle in order to provide the required air for aeration purpose in the WWTP, which may be called “bleed air”.

3. Model Analysis

This section addresses the modeling of the proposed system, and it is divided into two subsections, WWTP and CHP, which are modeled on mass and energy balance bases, respectively. A steady state assumption exists for both systems.

3.1. WWTP Model

In addition to the mass balance, there are other important parameters to consider such as biochemical oxygen demand (BOD), total suspended solid (TSS) and volatile suspended solid (VSS). Several assumptions have been made to coincide with reasonable estimates and are listed in Table 1.

3.1.1. Volatile Suspended Solid

Although many parameters are calculated during system modeling, the volatile suspended solid (VSS) calculation process is explained here since the total VSS is digested in the anaerobic digester. The sources of VSS are from the primary and secondary clarifiers and its calculation is shown in Equation (1).
V S S 9 = V S S 4 + V S S 7
Here, VSS4 and VSS7 indicate the volatile suspended solids from the primary and secondary clarifiers (subscripts correspond to states in Figure 1). The following equation is used to find VSS4:
V S S 4 = Q × T S S 9 × δ G C × δ P C f p c
Again, all parameters in Equation (2) are defined in Table 1. VSS production from the secondary clarifier is more complex than in the primary clarifier, see Equations (3)–(6).
V S S 7 = ( V S S 5 f s c Q × T S S 9 ) × δ t i c f s c
Here, VSS5 represents the produced suspended solid in the activated sludge process and can be calculated as shown in Equation (4).
V S S 5 = Q × Y o b s × ( B O D 0 S )
Yobs and S are the observed yield and the soluble BOD concentration in the effluent. Based on the assumption listed in Table 1, S can be calculated as follows:
S = B O D 9 f b × T S S 9 × U B O D × f
where fb, UBOD and f are the biodegradable TSS ratio, ultimate biological oxygen demand and a constant value to convert UBOD to effluent BOD concentration, respectively. To calculate observed yield, Equation (6) is used.
Y o b s = Y ( 1 + k d × S R T )
Y and kd are the biomass conversion factor and decay rate of microorganism, respectively.

3.1.2. Oxygen Transfer

As discussed earlier, oxygen addition is the fundamental process of the activated sludge process. Oxygen transfer requires either compressors or blowers, and they consume the highest amount of power in a WWTP. To calculate oxygen requirement in a WWTP, Equation (7) is used.
O T R f = Q ( S 0 S ) B O D 5 / B O D L 1.42 × V S S 5
Here, OTRf and BOD5/BODL denote the actual oxygen transfer rate at the plant and ratio of 5-day BOD to ultimate carbonaceous BOD, and assumed to be 0.68 [40]. While in some sources [39] BOD5/BODL were not taken into consideration, it is in some others [40]. For completeness, this study takes this parameter into consideration and follows [40] for actual oxygen transfer rate calculations. Due to the fact that aerators, blowers and compressors are used to provide oxygen to the activated sludge process, the standard oxygen transfer rate (SOTR) must be calculated using additional parameters as shown below [39]:
S O T R = O T R f α F [ C 20 β ( C s t / C s 20 ) ( P b / P s ) ( C 20 ) C ] [ ( 1.024 ) ( 20 T ) ]
Details and assumed values for various parameters in Equation (8) are seen in Table 2.

3.1.3. Power Requirement

As can be seen from Figure 1, bleed air will be used to provide required air from Compressor 1. There are two different ways to calculate the required power for adiabatic compression, the first is shown below [39]:
W c o m p , 1 = w R T 0 28.97   n e [ ( P y P x ) n 1 ]
where w is the air flow rate and calculated as follows:
w = S O T R ( E × 0.232   O 2 / k g   a i r )
where E denotes the oxygen transfer efficiency, assumed to be 0.25 [39]. The assumptions from [39] are made and listed in Table 3 for power requirement calculations.
However, there is a simpler method to calculate the required power once the air is assumed to be an ideal gas. If the mass flow rate, temperature and pressure of the air is known before and after compression, enthalpy of the streams can be found, hence the power requirement for the compressor.
W c o m p , 1 = m ˙ 24 ( h 25 h 24 )
While aeration consumes the most power, anaerobic digestion heating is the second largest power consumer in a CHP-WWTP. In order for mesophilic anaerobic digestion to take place, a temperature range between 30 and 40 °C is required; it is assumed to be 35 °C in this study [41,42]. In order to calculate the heat requirement for the digestion process, Equation (12) is used:
Q D i g = m ˙ s l u d g e C P , s l u d g e ( T D i g T 0 )
Here CP,sludge represents the specific heat of sludge, and it is assumed to be equal to specific heat of water as 95% of the sludge contains water. Since there is not enough data to calculate the total plant power requirement, and the sum of aeration and anaerobic digestion heat requirement consumes about 72% of the WWTP [43], the below equation is used to estimate the total power requirement for the WWTP.
W T o t a l ,   r e q . = W c o m p , 1 + Q D i g 0.72

3.1.4. Exergy Analysis

Here, both physical and chemical exergy of the WWTP must be taken into consideration since sewage has a high potential in terms of chemical exergy. An expression has been claimed by Tai et al. [44] so as to calculate the specific chemical exergy of sewage using COD concentration as expressed in the following equation:
e x s e w a g e C H = 13.6 × COD
The BOD5 values in the influent and effluent are 375 mg/L and 20 mg/L [39], COD values are 661.32 mg/L and 64.37 mg/L [34], respectively. Sludge is one of the final products of a WWTP, and the exergy of the sludge can be calculated as follows [34,45]:
e x s l C H = ( L H V s l + h e v a p z w ) β + ( e x s u l C H L H V s u l ) z s u l + e x a s h C H z a s h + e x w C H z w
Here, z, hevap, e x C H and β are the mass fractions of each component, enthalpy of water vaporization, specific chemical exergy of each component and atomic ratio, respectively.
An expression for β is as follows when the oxygen to carbon ratio is lower than 0.5 [45]:
β s l = 1.0437 + 0.0140 H C + 0.0968 O C + 0.0467 N C
Here, C, H, O, N, S, and ash represent the molar fractions of the carbon, hydrogen, oxygen, nitrogen, sulfur, and ash of a dried sludge in a municipal WWTP, and their molar fractions are 50, 2.5, 12.5, 1.1, 0.4, 10, respectively [34,39,46].
The last process of the WWTP is the anaerobic digestion, and the chemical exergy for the produced gas mixture is calculated as follows:
e x ¯ g a s ,   m i x t u r e C H = x i e x i 0 , C H + R T 0 x i l n ( x i )
While xi represent the molar fraction of any component in the biogas mixture, e x i 0 , C H denote specific chemical exergy of each component in the biogas mixture at the reference state. Once all calculations have been completed, energy and exergy efficiencies of the WWTP part can be calculated using Equations (18) and (19), respectively.
η W W T P = ( m ˙ t w e t w ) + ( m ˙ b i o g a s e b i o g a s ) + ( m ˙ s l e s l ) ( m ˙ s w e s w ) + W T o t a l ,   r e q .
ψ   W W T P = ( m ˙ t w e x t w ) + ( m ˙ b i o g a s e x b i o g a s ) + ( m ˙ s l e x s l ) ( m ˙ s w e x s w ) + W T o t a l ,   r e q .

3.2. Thermal Modeling

The first and second laws of thermodynamic analyses have been performed for the proposed system. Engineering Equation Solver (EES) is used for the calculations of both energy and exergy efficiencies. While evaluating the thermal modeling, the following assumptions have been made based on the references [38,42,47,48].
  • A steady state condition has been assumed for all processes;
  • Air and combustion products treated according to the ideal gas mixture law;
  • The fuel is assumed to be CH4, and only volume portion of the methane in the biogas mixture is taken into consideration;
  • A 2% heat loss of LHV of natural gas is considered in the combustion chamber, while all other systems are adiabatic;
  • The oxidizer in the combustion chamber is assumed to be air, and it contains 77.48% N2, 20.59 O2, 0.03% CO2 and 1.90% H2O on a volumetric basis [38];
  • Pressure drop in the air preheater, heat recovery steam generator, combustion chamber are considered to be 3%, 5% and 5%, respectively [38];
  • The temperature and pressure of the environment are assumed to be 298.15 K and 1.013 bar, respectively;
  • A temperature of 35 °C is assumed for the anaerobic digester [42].

3.2.1. Gas Turbine Cycle

As can be seen from Figure 1, a Brayton cycle with regeneration has been used in the proposed study. This system consists of a compressor, turbine, a combustion chamber and an air preheater. In this study, three variables, the compression ratio of the compressor, the temperature of the preheated air before the combustor chamber and the turbine inlet temperature, are variable and their effect on the efficiencies have been investigated. Details are listed in Table 4 and Table 5.
As mentioned earlier, biogas is the final product of the AD process, and there are many different ways to calculate the mass of produced biogas. The following equation is used to estimate the produced biogas as in [22,39,49]:
m ˙ b i o g a s = G P × V S D × V S S 9 × ρ r × ρ a i r
where GP, VSD, ρair and ρr represent the VSS destruction rate in an anaerobic digester, production factor of biogas, density of air and the density ratio of produced gas to the density of air, respectively. Since the biogas contains mainly CH4 (~60%) and CO2 (~40%), only CH4 portion in the biogas is considered here as a fuel. The mass flow rate of the air, which is the oxidizer in the combustion chamber, is calculated based on the combustion model which will be discussed later in the combustion chamber section.
Since the temperature, pressure and molar concentration of the air in state 13 are known, enthalpy and entropy values can be determined easily. In order to find the enthalpy value in state 14, the following equation is used.
η A C = h ¯ 14 s h ¯ 13 h ¯ 14 h ¯ 13
The isentropic compression of the air in the compressor is stated as [48]:
s ¯ 14 s s ¯ 13 = X N 2 [ s ¯ 0 ( T 14 s ) s ¯ 0 ( T 13 ) R ¯ l n P 14 P 13 ] N 2 + X o 2 [ s ¯ 0 ( T 14 s ) s ¯ 0 ( T 13 ) R ¯ l n P 14 P 13 ] o 2 + X c o 2 [ s ¯ 0 ( T 14 s ) s ¯ 0 ( T 13 ) R ¯ l n P 14 P 13 ] c o 2 + X H 2 o [ s ¯ 0 ( T 14 s ) s ¯ 0 ( T 13 ) R ¯ l n P 14 P 13 ] H 2 O = 0
The molar compositions of the components in the air are as given in Section 3.2. Once h ¯ 14 is calculated using Equation (21), temperature of the state 14 can be obtained using EES. Now, it is possible to calculate the compressor work using the following equation:
W A C , 2 = m ˙ a i r h ¯ 14 h ¯ 13 M a i r
where m ˙ a i r and M a i r are the mass flow rate and molecular weight of air, respectively. The energy balance for the preheater is as follows:
h ¯ 14 h ¯ 15 = ( 1 + λ ¯ ) h ¯ 18 h ¯ 17
λ ¯ is the ratio of molar flow rates of fuel to air, and will be explained in the next section.
P 15 P 14 = 1 Δ P A P
The mixture of air and produced biogas from AD is combusted in a combustion chamber. The molar flow ratio of the air, fuel and products are determined as follows [38]:
λ ¯ = n ˙ f n ˙ a     ,       1 + λ ¯ = n ˙ p n ˙ a     ,             n ˙ p = n ˙ f + n ˙ a
where n ˙ f , n ˙ a , n ˙ p are the molar flow rates of fuel, air and combustion products, respectively. The chemical equation of the combustion process on a per mole of air basis is described as below [38]:
λ ¯ C H 4 + [ 0.7748   N 2 + 0.2059   O 2 + 0.0003   C O 2 + 0.019   H 2 O ] λ ¯ [ 1 + λ ¯ ] [ X N 2   N 2 + X O 2   O 2 + X C O 2   C O 2 + X H 2 O   H 2 O ]
where,
X N 2 = 0.7748 1 + λ ¯ ,     X O 2 = 0.2059 2 λ ¯ 1 + λ ¯ ,     X C O 2 = 0.0003 + λ ¯ 1 + λ ¯ ,     X H 2 O = 0.019 + 2 λ ¯ 1 + λ ¯
Since a heat loss of a 2% of the lower heating value of the fuel is assumed, energy balance for the combustion chamber is as follows [38]:
0.02 λ ¯ L H V ¯ C H 4 + n ˙ f h ¯ f n ˙ p h ¯ p + n ˙ a h ¯ a = 0
or, equivalently,
0.02 λ ¯ L H V ¯ C H 4 + h ¯ a λ ¯ h ¯ f ( 1 + λ ¯ ) h ¯ p = 0
Methane has been used as a fuel here, and its lower heating (LHV) and enthalpy values are 802,361 kj/kmol and −74,872 kj/kmol, respectively [38]. To determine the enthalpy values of the air and combustion products, the ideal gas law mixture principle has been used as shown below:
h ¯ a = [ 0.7748 h ¯ N 2 + 0.2059 h ¯ O 2 + 0.0003 h ¯ C O 2 + 0.019 h ¯ H 2 O ] T = T a i r , i n = T 15
( 1 + λ ¯ ) h ¯ p = [ 0.7748 h ¯ N 2 + ( 0.2059 2 λ ¯ ) h ¯ O 2 + ( 0.0003 + λ ¯ ) h ¯ C O 2 + ( 0.019 + 2 λ ¯ ) h ¯ H 2 O ] T = T p , o u t = T 16
where
λ ¯ = 0.7748 Δ h ¯ N 2 + 0.2059 Δ h ¯ O 2 + 0.0003 Δ h ¯ C O 2 + 0.019 Δ h ¯ H 2 O h ¯ f 0.02 L H V ¯ ( 2 h ¯ O 2 + h ¯ C O 2 + 2 h ¯ H 2 O ) T = T 16
The pressure drop in the combustion chamber can be calculated as follows:
P 16 P 15 = 1 Δ P c c
Using the isentropic efficiency of the turbine, the enthalpy value of the exhaust gas from the gas turbine can be determined from the following equation:
η G T = h ¯ 16 h ¯ 17 h ¯ 16 h ¯ 17 s
Since the entropy values of the inlet and outlet of the turbine ( s ¯ 17 s s ¯ 16 = 0 ) is equal for an isentropic expansion, enthalpy and temperature of the stream 17 can be found. Using energy balance of the gas turbine, the produced power from the gas turbine can be expressed as follows:
W G T = m ˙ 16 h ¯ 16 h ¯ 17 M p

3.2.2. Gas Turbine Cycle Exergy Analysis

On the basis of the first and second law of thermodynamics, neglecting kinetic and potential exergy, the total exergy of a stream is calculated including physical and chemical exergies:
e x i = e x p h + e x c h
e x p h = ( h i h 0 ) T 0 ( s i s 0 )
e x ¯ g a s ,   m i x t u r e C H = x i e x i 0 , C H + R T 0 x i l n ( x i )
The energy and exergy efficiencies of the gas turbine cycle is expressed as the following equations.
η B r a y t o n = W G T W A C , 2 ( m ˙ b i o g a s L H V b i o g a s )
ψ   B r a y t o n = W G T W A C , 2 ( m ˙ b i o g a s e x b i o g a s )

3.2.3. Rankine Cycle

As explained earlier, the temperature of the leaving exhaust gas from the gas turbine cycle is assumed to be above 400 K. For this reason, a Rankine cycle is used to produce more power, and water has been used here as a working fluid. As can be seen from Figure 1, heat is transferred through a heat exchanger from state 20 to 21. Then, the superheated steam is expanded to produce power. The assumptions made for the Rankine cycle are listed in Table 6.
Applying energy balance for the heat recovery steam generator (HX2) assuming no heat loss to ambient, the mass flow rate of the water in the Rankine cycle can be calculated as shown below:
m ˙ 18 ( h 18 h 19 ) + m ˙ 20 ( h 20 h 21 ) = 0
To calculate energy and exergy efficiencies of the Rankine cycle for each component, the following equations are used:
W P = m ˙ 20 ( h 20 h 23 )
where Wp is the pump power. To calculate the power production from the steam turbine, Equation (44) is used.
W S T = m ˙ 21 ( h 22 h 21 )
Now, energy and exergy efficiencies of the Rankine cycle can be calculated as follows:
η R a n k i n e = W n e t ,   R a n k i n e m ˙ s t e a m ( h 21 h 20 )
ψ R a n k i n e = W n e t ,   R a n k i n e m ˙ s t e a m ( e x 21 e x 20 )
Once the calculations have been determined for both gas turbine and Rankine cycles, the efficiencies of the cogeneration system are expressed as follows including the compressor work due to WWTP aeration:
W T o t a l , C o g = W n e t ,   B r a y t o n + W n e t ,   R a n k i n e
η C o g = W T o t a l , C o g W A C , 1 ( m ˙ b i o g a s L H V b i o g a s )
ψ C o g = W T o t a l , C o g W A C , 1 ( m ˙ b i o g a s e x b i o g a s )

3.2.4. Overall Efficiencies

In the proposed multigeneration system, the useful outputs are considered to be the power production from the gas turbine and Rankine cycle, treated wastewater, and digested sludge. The inputs to the overall system are the influent sewage in the WWTP as well as the required power in the WWTP. The overall energy and exergy efficiencies of the overall system are expressed as follows:
η O v e r a l l = W T o t a l , C o g W T o t a l , r e q . + ( m ˙ t w e t w ) + ( m ˙ s l e s l ) ( m ˙ s w e s w ) + W T o t a l ,   r e q .
ψ O v e r a l l = W T o t a l , C o g W T o t a l , r e q . + ( m ˙ t w e x t w ) + ( m ˙ s l e x s l ) ( m ˙ s w e x s w ) + W T o t a l ,   r e q .
As explained earlier, the main purpose of this study is to determine whether the proposed cogeneration system can produce sufficient power to treat wastewater for a specified effluent standard. So as to examine this self-sufficiency of the proposed system, the following equation has been used, which represents the ratio of produced power from the cogeneration cycle to the sum of the power requirement of wastewater treatment and cogeneration system.
S S R = W T o t a l , C o g W T o t a l , r e q .

4. Results

In this section, a base and a parametric study have been performed varying a significant number of variables in order to investigate the performance of the wastewater treatment plant, and combined gas-vapor cycle from the point of view of first and second law efficiencies. In addition, power requirement for the WWTP, including the power requirement of aeration and heat requirement for anaerobic digestion, is evaluated for different scenarios.

4.1. Case Study

A case study has been conducted here using data from a well-known study [39]. Furthermore, several parameters listed in Table 7 are varied to calculate the efficiencies of each system as well as power requirement for the WWTP.
Before starting the discussion of energy and exergy efficiencies of each subsystem, energy transfer rate, exergy destruction rate and exergy efficiencies for Brayton cycle components have been listed in Table 8. It is obvious from the table that the highest energy transfer and exergy destruction take place in the combustion chamber due to a high entropy generation during the combustion process. Furthermore, heat exchanger II is found to have the second highest exergy destruction rate due to a high heat transfer from exhaust gas to the Rankine cycle. Exergy efficiencies of the Brayton cycle components varied from 55.9% to 92.9%. Heat exchanger I and compressor II have been found to have the lowest and highest exergy efficiencies, respectively.
Figure 2 indicates the energy and exergy efficiencies of each subsystem. As can be seen, overall energy and exergy efficiencies for the case study are found to be 41.2% and 32.2%, respectively. While energy and exergy efficiencies are 28.96% and 28.19% for the Brayton cycle, they are 28.41% and 68.44% for the Rankine cycle. Since chemical exergy is quite high in the influent of the wastewater, exergy efficiency is higher than energy efficiency in WWTP. Figure 2b illustrates the power requirement for aeration and the total for the WWTP as well as the power production from both Brayton and Rankine cycles. Also, while a power of 219.5 kW is produced from the cogeneration system, a total power of 226.1 kW is required for the WWTP. Hence, it can be said that 97% of the total energy requirement of the WWTP can be provided using the multigeneration system.
Table 9 indicates the thermodynamic properties of each stream including mass flow rate, pressure, temperature, enthalpy, specific entropy and total specific exergy for the case study.
In the following section, a parametric study has been conducted varying important variables for both WWTP, and cogeneration systems. While biological oxygen demand in the effluent and dissolved oxygen level in the WWTP varied, turbine inlet temperature, compression ratio as well as air preheater temperature changed for the cogeneration system. The variables have been listed in Table 7.

4.2. Parametric Study—WWTP

4.2.1. Biological Oxygen Demand

Since activated sludge is chosen here as a treatment method, the two important variables of this process, total effluent BOD and DO level, are varied in order to investigate the soluble BOD in the effluent, standard oxygen transfer rate (SOTR), total power requirement for the multigeneration system, and energy and exergy efficiencies of each subsequent system. While the BODe is varied from 15 to 25 mg/L, DO concentration is changed from 2 to 4 mg/L. As Riffat [50] stated, a minimum DO concentration of 2 mg/L is required in order for most aquatic plants and animals to survive. Furthermore, game fish and other higher life-forms require 4 mg/L or more for survival [50]. For this reason, a minimum level of 2 mg/L DO level is chosen for most calculations.
Figure 3a shows the variation of Se and SOTR with changing BODe. It is not surprising that, Se increased from 1.4 to 11.4 mg/L when BODe is increased from 15 mg/L to 25 mg/L. As can be seen from Equation (5), the soluble BOD (Se) depends on only total effluent BOD in the effluent and the BOD in the suspended solid; hence, there is a linear relation between BODe and Se. Figure 3a also shows the standard transfer oxygen rate with changing BODe. Since higher BODe results in lower quality treated water, oxygen requirement will decrease when the BODe is higher; hence, Se is higher. SOTR decreased from 396 to 380 kg/h when BODe is changed from 15 to 25 mg/L.
As discussed earlier, aeration shares the highest portion of power requirement in a WWTP. Figure 3b shows the power requirement for the entire WWTP, including the power requirement for aeration. As is obvious from the graph, there is an inverse relation between BODe and power requirement since SOTR is lower when BODe is higher. Power production using the cogeneration system, also indicated in Figure 3b, and the power production decreased with increasing BODe since the amount of produced biogas is lower when the BODe is higher. It is found here that about 108% of the required power in WWTP can be produced using the proposed cogeneration system (see Figure 3c). However, it should be noted that DO level in this case is 2 mg/L, which is the lowest acceptable concentration in a WWTP. More detailed analysis varying the DO level is discussed in the next section.
Energy and exergy efficiencies of the WWTP and overall system have been indicated in Figure 3c. A change in the BODe level did not show a high influence on the thermodynamic efficiencies. While the energy efficiency of the WWTP varied from 43.18% to 43.55% in the WWTP, overall energy efficiency increased from 44.92% to 43.55% when the BODe level is varied from 15 to 25 mg/L. As mentioned earlier, more power can be produced than required for WWTP using the activated sludge as shown in Figure 3c. The self-sufficiency ratio (SSR) varied from 106.2% to 108.1% for the given BODe range.

4.2.2. Dissolved Oxygen Concentration

A detailed parametric study has been carried out to investigate the effect of DO concentration on the thermodynamic efficiencies and oxygen requirement. The BODe level is assumed to be constant at 20 mg/L in this case. Looking at Figure 4a, standard oxygen transfer rate (SOTR) variation with DO concentration is quite obvious. When the DO concentration increased from 2 to 4 mg/L, SOTR jumps from 388 to 555 kg/h. The increment is about 43%, and SOTR is fundamentally dependent on DO concentration. An increase in the DO level causes a substantial increase in aeration power requirement; hence total power requirement for the WWTP. As obvious from the Figure 4b that, while the required power for aeration increased from 87 to 124 kW, the total WWTP power requirement increased from 204 to 256 klW when the DO concentration varied from 2 to 4 mg/L. Since DO level does not have an effect on the biogas production rate, produced power from the cogeneration system will be constant at 219 kW for this case. Therefore, the energy efficiencies of WWTP and overall will decrease with DO level increment. While the energy efficiency of WWTP decreased from 43.4% to 40.5%, overall energy efficiency decreased from 45.3% to 35.7% when the DO concentration is varied from 2 to 4 mg/L. As considered earlier, DO level has a significant effect on the total power requirement because of a higher oxygen requirement. As shown in Figure 4b, self-sufficiency ratio decreased from 107% to 85% when the DO concentration increased from 2 to 4 mg/L. Therefore, the proposed system is not able to provide its power itself for higher DO concentrations.

4.3. Parametric Study—Cogeneration Power System

In this section, performance of the proposed cogeneration energy system will be investigated varying the turbine inlet temperature, compression ratio and the preheater temperature in the Brayton cycle. The values of the parameters for the parametric study have been provided in Table 7.

4.3.1. Gas Turbine Inlet Temperature (TIT)

In the literature, the proposed cogeneration energy system is referred to with different names such as integrated biogas-based power generation system [48] and biogas fueled power generation system [47]. In this paper, we refer to the proposed system as cogeneration (cog) power system. Based on literature the turbine inlet temperature should not exceed 1277 °C [38]. For this reason, the inlet temperature in this study is varied from 700 to 1200 °C. Except TIT, all other parameters are assumed to be constant for both Brayton and Rankine cycle as shown in Table 7 for the base study except the DO level, which is assumed to be 2 mg/L.
Figure 5 indicates the effect of TIT on the air to fuel ratio and the exhaust gas mass flow rate. As seen, exhaust gas mass flow rate decreases with increasing the TIT since fuel mass flow rate will be constant and the air fuel ratio will be lower when TIT is higher. While air fuel ratio decreases from 72 to 26, exhaust gas mass flow rate decreases from 1.27 to 0.46 kg/s when TIT varies from 700 to 1200 °C.
TIT effect is one of the most important factors in evaluating the performance of Brayton cycles. While Figure 6a shows the variation of the mass flow rates of air and combustion products in the Brayton cycle and water in the Rankine cycle, Figure 6b shows the variation of power generation from Brayton cycle, Rankine cycle and the total power production of this integrated system as well as the self-sufficiency ratio. While mass flow rate of the exhaust gas decreases with increasing TIT, mass flow rate of the water in the Rankine cycle increases because more heat is transferred in HX2 when TIT is higher, as seen in Figure 6a. Although there is a tremendous decrease in the exhaust gas mass flow rate, net powers from Brayton and Rankine cycles increase because of higher TIT. Despite the fact that the unit power production from Rankine cycle does not change when the TIT varies, the mass flow rate of working fluid in the Rankine cycle increases with increasing TIT, as shown in Figure 6a. Hence, the net power production increment in the Rankine cycle is not surprising. Since BODe and DO levels are kept constant here, the total power requirement for the WWTP is calculated to be 204.8 kW. As can be seen from Figure 6b, a total power of 219.5 kW can be produced from the cogeneration system when the TIT is 1200 °C. However, it should be noted that the self-sufficiency ratio varies from 76.6% to 107.2% when TIT changes from 700 °C to 1200 °C. Therefore, it can be said that TIT has the highest effect on the self-sufficiency of the proposed system.
Energy and exergy efficiency change of cogeneration and overall systems with TIT variation have been illustrated in Figure 7a,b, respectively. While energy and exergy efficiencies of WWTP are 28.4% and 68.4%, the efficiencies for Rankine cycles are 43.3% and 53.2%, respectively. Since the efficiency of the WWTP and Rankine cycle did not change with TIT variation, they were not included to the graphs. Whereas energy efficiency of the cogeneration system increased from 14% to 27%, it increased from 36% to 45% for the overall system when TIT varied from 700 to 1200 °C. The reason behind of these increments can be explained by increase on the power generation from both Brayton and Rankine cycles. Exergy efficiencies of the cogeneration system increased from 14% to 26% due to exergy destruction reduction in the Brayton cycle.

4.3.2. Compression Ratio (Rp)

Compression ratio effect on the efficiencies and power production is investigated here. As shown in Table 7, TIT and air preheater temperature is assumed to be constant here; hence AF is constant. Figure 8a,b show energy and exergy efficiencies of the Brayton cycle, cogeneration system as well as overall system. While energy efficiency of the cogeneration system increased from 18% to 27%, the overall system energy efficiency increased from 39.6% to 45.6% when compression ratio is varied from 3 to 15. The main reason for the energy efficiency increment can be explained by increase in power production with higher compression ratios. As can be seen from Figure 8c, while the Brayton cycle is able to produce 144.9 kW of power, the Rankine cycle produces 76.9 kW when the compression ratio is set to 15. It is obvious from this figure that when the compression ratio is lower, the power production from the Rankine cycle is higher due to the fact that the turbine outlet temperature of the exhaust gas is too high; hence more heat is transferred from Brayton cycle to Rankine cycle. Therefore, the mass flow rate of the working fluid in the Rankine cycle is too high. Self-sufficiency ratio is also shown in Figure 8c, and it is varied from 86% to 108% when compression ratio is changed from 3 to 15.

4.3.3. Preheater Temperature

In a gas turbine cycle, regeneration plays an important role in order to increase the temperature of the air entering to the combustion chamber. As shown in Figure 9a, the exhaust mass flow rate increases with increasing preheater temperature since more air is consumed for a constant fuel amount in the combustor as the temperature difference between the inlet and outlet of the combustion chamber is lower for higher preheater temperature. Figure 9b indicates the power production from Brayton and Rankine cycles. While the Brayton cycle is able to produce more power when the preheater temperature is increased, the Rankine cycle power production tends to decrease due to a decrease in the working fluid mass flow rate. Obviously, this decrement is a result of temperature drop in the turbine outlet exhaust gas. Energy and exergy efficiencies are shown in Figure 9c,d, and it can be said that the change in the overall efficiencies with preheater temperature variation are very small. While Brayton cycle energy and exergy efficiencies vary from 28.1% to 30.5% and 27.4 % to 29.7%, cogeneration energy and exergy efficiencies change from 26.7% to 28.1% and from 26% to 27.3%, respectively.

5. Conclusions

In this paper, a theoretical thermodynamic study has been conducted to investigate the self-sufficiency of a WWTP from the standpoint of energy. A novel multigeneration system has been developed by adding a secondary compressor to the combined gas–vapor cycle to provide air for the WWTP aeration as well as eliminate the generator efficiency in the gas turbine. The energy and exergy efficiency variations with several parameters for both WWTP and power cycles are investigated. The key findings in this study are:
  • While energy efficiency of the overall system varied from 35.7% to 46.0, exergy efficiency changed from 30.6% to 33.55%;
  • The highest efficiencies were obtained for a turbine inlet temperature, air preheater temperature, compression ratio, effluent biological oxygen demand and dissolved oxygen level of 1200 °C, 427 °C, 10, 20 mg/L and 2 mg/L, respectively;
  • Self-sufficiency ratio varied between 76.6% to 109.4%. Hence, a self-sufficient system is possible using the proposed multigeneration system;
  • Self-sufficiency ratio can be increased by 42% and 12.7% by choosing the optimal parameters compared to the least efficient system and case study, respectively;
  • While an increase in the desired effluent BOD result in an increase in the oxygen requirement, a small increase in the desired DO level cause a tremendous rise in the oxygen requirement;
  • Similarly, while efficiencies do not change considerably varying effluent BOD, the self-sufficiency ratio varies from 85% to 107% when the DO level is varied. Therefore, the proposed system is found to be self-sufficient when the DO level is lower than 3 mg/L;
  • A power production increase of up to 52.9% can be realized by integrating a Rankine cycle into the Brayton cycle for lower compression ratios;
  • While turbine inlet temperature and DO level are found to be the most dominant decision variables for self-sufficiency ratio variations, desired effluent BOD and air preheater temperature are found to be the least dominant parameters;
  • While the most important factor on the self-sufficiency ratio is found to be turbine inlet temperature, DO concentration also has a high effect on the self-sufficiency ratio since it has the highest effect on the required power for aeration process.
In the future, different power cycle models with increased complexity will be investigated to increase the self-sufficiency of the WWTP, with the goal of providing energy back to the grid at an increased rate including lifecycle assessment and cost analysis. Additionally, since bio-energy from WWTPs are considered a renewable resource, the authors plan to incorporate carbon capture from exhaust streams, which could possibly result in WWTPs becoming not only self-sufficient but also negative carbon emitters. The authors concede that many simplifying assumptions have been made, especially for subsystem components (e.g., ideal gases, adiabatic processes, ideal turbine in the Rankine cycle, etc.). However, the proposed model does obtain results very close to some analytical [38,39] and numerical studies [34], and may serve as a very reasonable first approach to modeling and analyzing self-sufficiency of WWTPs.

Author Contributions

Conceptualization, M.E. and D.W.M.; methodology, M.E.; software, M.E.; validation, M.E., and D.W.M.; formal analysis, M.E.; data curation, M.E.; writing—original draft preparation, M.E. and D.W.M.; writing—review and editing, M.E. and D.W.M.; visualization, M.E.; supervision, D.W.M.; project administration, D.W.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

General
ADAnaerobic digestion
BODBiological oxygen demand (mg/L)
COperating DO (mg/L)
CHPCombined heat and power
CODChemical oxygen demand (mg/L)
CpSpecific heat (kj/kgK)
DODissolved oxygen (mg/L)
EAeration efficiency
eSpecific energy (kj/kg)
exSpecific exergy (kj/kg)
FDiffuser fouling factor
GPVSS destruction rate (m3/kg VSS destroyed)
hSpecific enthalpy (kj/kg)
HXHeat exchanger
kdDecay rate of microorganism (1/day)
LHVLower heating value (kj/kmol)
MMolecular weight (kg/kmol)
m ˙ Mass flow rate (kg/s)
n ˙ Molar flow rate (kmol/s)
NZENet zero energy
OTRActual oxygen transfer rate (kg/h)
PPressure (bar)
QFlow rate (m3/d)
RUniversal gas constant (J/mol.K)
RpCompression ratio
SSoluble BOD (mg/L)
sSpecific entropy (kj/kgK)
SOTRStandard oxygen transfer rate (kg/h)
SRTSolid retention time (day)
SSRSelf-sufficiency ratio
TTemperature (°C)
TITTurbine inlet temperature (°C)
TSSTotal suspended solid (mg/L)
UBODUltimate biological oxygen demand
VSDBiogas production factor
VSSVolatile suspended solid (mg/L)
WPower (kW)
wAir flow rate (kg/h)
WENWater energy nexus
WWTPWastewater treatment plant
XMolar concentration
YBiomass conversion factor
zMass fraction
δRemoval efficiency
ΔPPressure change
λFuel to air ratio
ρDensity (kg/m3)
Greek Symbols
αRelative transfer rate to clean water
βRelative DO saturation to clean water
ηEnergy efficiency
ψExergy efficiency
Subscripts
aAir
acAir compressor
chChemical
compCompressor
DeDewatering
digDigester
fFuel
gcGrit chamber
gtGas turbine
pProduct
pcPrimary clarifier
phPhysical
scSecondary clarifier
slSludge
sulSulfur
swSewage
ticThickener
twTreated wastewater
wWater

References

  1. Angelakis, A.N.; Zheng, X.Y. Evolution of water supply, sanitation, wastewater, and stormwater technologies globally. Water 2015, 7, 455–463. [Google Scholar] [CrossRef] [Green Version]
  2. Salgot, M.; Folch, M. Wastewater treatment and water reuse. Curr. Opin. Environ. Sci. Health 2018, 2, 64–74. [Google Scholar] [CrossRef]
  3. United Nations Educational Scientific and Cultural Organization. United Nations World Water Development Report 2020: Water and Climate Change; United Nations Educational, Scientific and Cultural Organization: Paris, France, 2020. [Google Scholar]
  4. Rojas, J.; Zhelev, T. Energy efficiency optimisation of wastewater treatment: Study of ATAD. Comput. Chem. Eng. 2012, 38, 52–63. [Google Scholar] [CrossRef]
  5. US. EPA. Energy Efficiency in Water and Wastewater Facilities; U.S. Environment Protection Agency: Washington, DC, USA, 2014; p. 49.
  6. Qandil, M.D.; Abbas, A.I.; Salem, A.R.; Abdelhadi, A.I.; Hasan, A.; Nourin, F.N.; Abousabae, M.; Selim, O.M.; Espindola, J.; Amano, R.S. Net zero energy model for wastewater treatment plants. J. Energy Resour. Technol. 2021, 143, 122101. [Google Scholar] [CrossRef]
  7. Lee, S.; Esfahani, I.J.; Ifaei, P.; Moya, W.; Yoo, C. Thermo-environ-economic modeling and optimization of an integrated wastewater treatment plant with a combined heat and power generation system. Energy Convers. Manag. 2017, 142, 385–401. [Google Scholar] [CrossRef]
  8. Olsson, G. Water and Energy: Threats and Opportunities; IWA Publishing: London, UK, 2012. [Google Scholar]
  9. Ali, B.; Kumar, A. Development of life cycle water footprints for gas-fired power generation technologies. Energy Convers. Manag. 2016, 110, 386–396. [Google Scholar] [CrossRef]
  10. Sotomayor, O.; Park, S.; Garcia, C. A simulation benchmark to evaluate the performance of advanced control techniques in biological wastewater treatment plants. Braz. J. Chem. Eng. 2001, 18, 81–101. [Google Scholar] [CrossRef]
  11. Jafarinejad, S. Cost estimation and economical evaluation of three configurations of activated sludge process for a wastewater treatment plant (WWTP) using simulation. Appl. Water Sci. 2017, 7, 2513–2521. [Google Scholar] [CrossRef] [Green Version]
  12. Drewnowski, J.; Remiszewska-Skwarek, A.; Duda, S.; Łagód, G. Aeration process in bioreactors as the main energy consumer in a wastewater treatment plant. Review of solutions and methods of process optimization. Processes 2019, 7, 311. [Google Scholar] [CrossRef] [Green Version]
  13. Sánchez, F.; Rey, H.; Viedma, A.; Nicolás-Pérez, F.; Kaiser, A.; Martínez, M. CFD simulation of fluid dynamic and biokinetic processes within activated sludge reactors under intermittent aeration regime. Water Res. 2018, 139, 47–57. [Google Scholar] [CrossRef]
  14. Mohseni, E.; Herrmann-Heber, R.; Reinecke, S.; Hampel, U. Bubble generation by micro-orifices with application on activated sludge wastewater treatment. Chem. Eng. Process. Process. Intensif. 2019, 143, 107511. [Google Scholar] [CrossRef]
  15. Mahmud, R.; Erguvan, M.; Macphee, D.W. Performance of closed loop venturi aspirated aeration system: Experimental study and numerical analysis with discrete bubble model. Water 2020, 12, 1637. [Google Scholar] [CrossRef]
  16. Kim, M.; Yoo, C. Multi-objective controller for enhancing nutrient removal and biogas production in wastewater treatment plants. J. Taiwan Inst. Chem. Eng. 2014, 45, 2537–2548. [Google Scholar] [CrossRef]
  17. Appels, L.; Baeyens, J.; Degrève, J.; Dewil, R. Principles and potential of the anaerobic digestion of waste-activated sludge. Prog. Energy Combust. Sci. 2008, 34, 755–781. [Google Scholar] [CrossRef]
  18. Calli, O.; Colpan, C.O.; Gunerhan, H. Performance assessment of a biomass-fired regenerative ORC system through energy and exergy analyses. In Exergetic, Energetic and Environmental Dimensions; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 253–277. [Google Scholar]
  19. Aydin, E.S.; Yucel, O.; Sadikoglu, H. Numerical investigation of fixed-bed downdraft woody biomass gasification. In Exergetic, Energetic and Environmental Dimensions; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 323–339. [Google Scholar]
  20. Shizas, I.; Bagley, D. Experimental determination of energy content of unknown organics in municipal wastewater streams. Am. Soc. Civ. Eng. 2004, 130, 45–53. [Google Scholar]
  21. Shen, Y.; Linville, J.L.; Urgun-Demirtas, M.; Mintz, M.M.; Snyder, S.W. An overview of biogas production and utili-zation at full-scale wastewater treatment plants (WWTPs) in the United States: Challenges and opportunities towards energy-neutral WWTPs. Renew. Sustain. Energy Rev. 2015, 50, 346–362. [Google Scholar] [CrossRef] [Green Version]
  22. Farahbakhsh, M.T.; Chahartaghi, M. Performance analysis and economic assessment of a combined cooling heating and power (CCHP) system in wastewater treatment plants (WWTPs). Energy Convers. Manag. 2020, 224, 113351. [Google Scholar] [CrossRef]
  23. Mauger, G.; Tauveron, N.; Bentivoglio, F.; Ruby, A. On the dynamic modeling of Brayton cycle power conversion systems with the CATHARE-3 code. Energy 2019, 168, 1002–1016. [Google Scholar] [CrossRef]
  24. Jenicek, P.; Kutil, J.; Benes, O.; Todt, V.; Zabranska, J.; Dohanyos, M. Energy self-sufficient sewage wastewater treatment plants: Is optimized anaerobic sludge digestion the key? Water Sci. Technol. 2013, 68, 1739–1744. [Google Scholar] [CrossRef]
  25. Schwarzenbeck, N.; Pfeiffer, W.; Bomball, E. Can a wastewater treatment plant be a powerplant? A case study. Water Sci. Technol. 2008, 57, 1555–1561. [Google Scholar] [CrossRef]
  26. Nowak, O.; Keil, S.; Fimml, C. Examples of energy self-sufficient municipal nutrient removal plants. Water Sci. Technol. 2011, 64, 1–6. [Google Scholar] [CrossRef]
  27. Nowak, O.; Enderle, P.; Varbanov, P. Ways to optimize the energy balance of municipal wastewater systems: Lessons learned from Austrian applications. J. Clean. Prod. 2015, 88, 125–131. [Google Scholar] [CrossRef]
  28. Willis, J.; Stone, L.; Durden, K.; Beecher, N.; Hemenway, C.; Greenwood, R. Barriers to Biogas Utilization for Renewable Energy; IWA Publishing: London, UK, 2012. [Google Scholar]
  29. Yan, P.; Shi, H.-X.; Chen, Y.-P.; Gao, X.; Fang, F.; Guo, J.-S. Optimization of recovery and utilization pathway of chemical energy from wastewater pollutants by a net-zero energy wastewater treatment model. Renew. Sustain. Energy Rev. 2020, 133, 110160. [Google Scholar] [CrossRef]
  30. Erguvan, M.; MacPhee, D.W. Second law optimization of heat exchangers in waste heat recovery. Int. J. Energy Res. 2019, 43, 5714–5734. [Google Scholar] [CrossRef]
  31. Erguvan, M.; Macphee, D.W. Energy and exergy analyses of tube banks in waste heat recovery applications. Energies 2018, 11, 2094. [Google Scholar] [CrossRef] [Green Version]
  32. Siddiqui, O.; Dincer, I.; Yilbas, B.S. Development of a novel renewable energy system integrated with biomass gasification combined cycle for cleaner production purposes. J. Clean. Prod. 2019, 241, 118345. [Google Scholar] [CrossRef]
  33. Ozdil, N.F.T.; Tantekin, A. Exergy and exergoeconomic assessments of an electricity production system in a running wastewater treatment plant. Renew. Energy 2016, 97, 390–398. [Google Scholar] [CrossRef]
  34. Abusoglu, A.; Demir, S.; Kanoglu, M. Thermoeconomic assessment of a sustainable municipal wastewater treatment system. Renew. Energy 2012, 48, 424–435. [Google Scholar] [CrossRef]
  35. Abusoglu, A.; Demir, S.; Kanoglu, M. Thermodynamic analysis and assessment of a wastewater treatment plant in scope of anaerobic sludge digestion and on-site electricity production from biogas. In Proceedings of the 2nd International Exergy, Life Cycle Assessment, and Sustainability Workshop & Symposium (ELCAS2), Nisyros, Greece, 19–21 June 2011. [Google Scholar]
  36. Abusoglu, A.; Demir, S.; Kanoglu, M. Exergoeconomic assessment of a municipal primary and secondary sewage treatment. Int. J. Exergy 2012, 11, 387. [Google Scholar] [CrossRef]
  37. Cengel, Y.A.; Boles, M.A. Thermodynamics an Engineering Approach, 8th ed.; McGraw-Hill Education: New York, NY, USA, 2015. [Google Scholar]
  38. Bejan, A.; Tsatsaronis, G.; Moran, M. Thermal Design and Optimization; John Wiley & Sons: New York, NY, USA, 1995. [Google Scholar]
  39. Metcalf and Eddy Inc. Wastewater Engineering: Treatment and Resource Recovery, 5th ed.; McGraw-Hill Education: New York, NY, USA, 2014. [Google Scholar]
  40. Qasim, S.R.; Zhu, G. Wastewater Treatment and Reuse: Theory and Design Examples; CRC Press: Boca Raton, FL, USA, 2017; Volume 1. [Google Scholar]
  41. Weiland, P. Biogas production: Current state and perspectives. Appl. Microbiol. Biotechnol. 2010, 85, 849–860. [Google Scholar] [CrossRef]
  42. Safari, F.; Dincer, I. Development and analysis of a novel biomass-based integrated system for multigeneration with hydrogen production. Int. J. Hydrogen Energy 2019, 44, 3511–3526. [Google Scholar] [CrossRef]
  43. Gu, Y.; Li, Y.; Li, X.; Luo, P.; Wang, H.; Robinson, Z.P.; Wang, X.; Wu, J.; Li, F. The feasibility and challenges of energy self-sufficient wastewater treatment plants. Appl. Energy 2017, 204, 1463–1475. [Google Scholar] [CrossRef] [Green Version]
  44. Tai, S.; Matsushige, K.; Goda, T. Chemical exergy of organic matter in wastewater. Int. J. Environ. Stud. 1986, 27, 301–315. [Google Scholar] [CrossRef]
  45. Szargut, J.; Morris, D.R.; Steward, F.R. Energy Analysis of Thermal, Chemical, and Metallurgical Processes; Hemisphere Publishing: New York, NY, USA, 1988. [Google Scholar]
  46. Cartmell, E.; Gostelow, P.; Riddell-Black, D.; Simms, N.; Oakey, J.; Morris, J.; Jeffrey, P.; Howsam, P.; Pollard, S.J. Biosolids—A fuel or a waste? An integrated appraisal of five co-combustion scenarios with policy analysis. Environ. Sci. Technol. 2006, 40, 649–658. [Google Scholar] [CrossRef] [Green Version]
  47. Gholizadeh, T.; Vajdi, M.; Mohammadkhani, F. Thermodynamic and thermoeconomic analysis of basic and modified power generation systems fueled by biogas. Energy Convers. Manag. 2019, 181, 463–475. [Google Scholar] [CrossRef]
  48. Hosseini, S.E.; Barzegaravval, H.; Wahid, M.A.; Ganjehkaviri, A.; Sies, M.M. Thermodynamic assessment of integrated biogas-based micro-power generation system. Energy Convers. Manag. 2016, 128, 104–119. [Google Scholar] [CrossRef]
  49. Erguvan, M.; MacPhee, D.W. Analysis of a Multigeneration Energy System for Wastewater Treatment. In Proceedings of the ASME 2021 Power Conference, ASME 2021 Power Conference, Virtual, Online, 20–22 July 2021. [Google Scholar] [CrossRef]
  50. Riffat, R. Fundamentals of Wastewater Treatment and Engineering; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
Figure 1. Proposed System.
Figure 1. Proposed System.
J 04 00045 g001
Figure 2. (a) Energy and exergy efficiencies for the case study, (b) Power requirement, power production and self-sufficiency ratio.
Figure 2. (a) Energy and exergy efficiencies for the case study, (b) Power requirement, power production and self-sufficiency ratio.
J 04 00045 g002
Figure 3. BOD effect on: (a) Soluble BOD in the effluent and SOTR, (b) Efficiencies, (c) Power requirement and power production.
Figure 3. BOD effect on: (a) Soluble BOD in the effluent and SOTR, (b) Efficiencies, (c) Power requirement and power production.
J 04 00045 g003aJ 04 00045 g003b
Figure 4. Dissolved oxygen level effect on: (a) Efficiencies and SOTR, (b) Power requirement and power production.
Figure 4. Dissolved oxygen level effect on: (a) Efficiencies and SOTR, (b) Power requirement and power production.
J 04 00045 g004
Figure 5. Turbine inlet temperature effect on air fuel ratio and exhaust gas mass flow rate.
Figure 5. Turbine inlet temperature effect on air fuel ratio and exhaust gas mass flow rate.
J 04 00045 g005
Figure 6. Turbine inlet temperature effect on: (a) Mass flow rate of exhaust gas, air, and water (b) self-sufficiency ratio and power production (total power requirement for the WWTP is 184.3 kW).
Figure 6. Turbine inlet temperature effect on: (a) Mass flow rate of exhaust gas, air, and water (b) self-sufficiency ratio and power production (total power requirement for the WWTP is 184.3 kW).
J 04 00045 g006
Figure 7. Turbine inlet temperature effect on: (a) energy efficiencies (b) exergy efficiencies.
Figure 7. Turbine inlet temperature effect on: (a) energy efficiencies (b) exergy efficiencies.
J 04 00045 g007
Figure 8. Compression ratio effect on: (a) energy efficiencies, (b) exergy efficiencies, (c) power production, and self-sufficiency ratio (total required power for the WWTP is 184.3 kW).
Figure 8. Compression ratio effect on: (a) energy efficiencies, (b) exergy efficiencies, (c) power production, and self-sufficiency ratio (total required power for the WWTP is 184.3 kW).
J 04 00045 g008aJ 04 00045 g008b
Figure 9. Combustor inlet temperature effect on: (a) air fuel ratio and exhaust gas mass flow rate, (b) power production, and self-sufficiency ratio (total required power for the WWTP is 184.3 kW), (c) Energy Efficiency, (d) Exergy Efficiency.
Figure 9. Combustor inlet temperature effect on: (a) air fuel ratio and exhaust gas mass flow rate, (b) power production, and self-sufficiency ratio (total required power for the WWTP is 184.3 kW), (c) Energy Efficiency, (d) Exergy Efficiency.
J 04 00045 g009aJ 04 00045 g009b
Table 1. WWTP modeling assumptions [39].
Table 1. WWTP modeling assumptions [39].
DescriptionParameterUnitValue
Wastewater flow rateQ(m3/d)22,600
Influent BODBOD0(mg/L)375
Influent TSS concentrationTSS1(mg/L)400
Grit removal efficiencyδGC-0.9
Primary clarifier removal efficiencyδPC-0.7
Thickener removal efficiencyδTic-0.9
Dewatering removal efficiencyδDe-0.93
Effluent total BODBOD9(mg/L)20
Effluent TSSTSS9(mg/L)22
Fraction of biodegradable solidsfb-0.65
Solid retention timeSRTdays10
Biomass conversion factorY-0.5
Decay rate of microorganism kd(1/day)0.06
Ultimate biological oxygen demandUBOD-1.42
Conversion factor UBOD to BODef-0.68
Volatile fraction in primary clarifierfpc-0.683
Volatile fraction in secondary clarifierfsc-0.8
Biogas production factorVSD-0.5
Density of airρairkg/m31.204
Density ratio of biogas/density of airρr-0.86
VSS destruction rateGPm3/kg of VSS destroyed0.95
Table 2. Aeration process assumptions [39].
Table 2. Aeration process assumptions [39].
DescriptionParameterUnitValue
Relative transfer rate to clean waterα-0.50
Relative DO saturation to clean waterβ-0.95
Diffuser fouling factorF-0.90
Saturated DO value at sea level and 20 °C for diffused aeration (mg/L) C 20 (mg/L)10.64
Saturated DO value at sea level and 20 °C C s 20 (mg/L)9.09
Saturated DO at sea level and operating temperature (at 25 °C) C s t (mg/L)8.263
Pressure correction factor P b / P s -0.94
Operating DO in basinC(mg/L)2.0
Aeration basin temperatureT°C25
Table 3. Power requirement assumptions [39].
Table 3. Power requirement assumptions [39].
DescriptionParameterUnitValue
Universal gas constantRJ/mol·K8.314
Absolute inlet temperatureT0K298
Absolute inlet pressurePxkPa101.3
Absolute outlet pressurePykPa156.5
Compressor efficiencye-0.85
Oxygen transfer efficiencyE-0.25
Specific heat ration-0.285
Table 4. Decision variables [38,48].
Table 4. Decision variables [38,48].
DescriptionParameterUnitRange
Compression ratioRp-3–15
Air temperature before combustorT15°C347–427
Turbine inlet temperatureT16°C700–1200
Table 5. Constant values for the Cogeneration system [38,48].
Table 5. Constant values for the Cogeneration system [38,48].
ParameterUnitRange
T1°C25
P1Bar1.013
T15°C377
T16°C927
Rp-10
ηAC-0.86
ηGT-0.86
T19°C153
Table 6. Constant values for the Rankine cycle [37].
Table 6. Constant values for the Rankine cycle [37].
ParameterUnitRange
Steam Turbine inlet temperature, T21°C500
Steam Turbine inlet pressure, P21Bar30
Condenser pressure, P22Bar0.75
Table 7. Case Study Parameters and their variation in parametric study.
Table 7. Case Study Parameters and their variation in parametric study.
ParameterUnitBase
Study
Variation
[39,42,48]
WWTPEffluent total BOD, BOD9mg/L2015–25
Dissolved Oxygen Level, DOmg/L32–4
Gas Turbine
Cycle
Gas Turbine inlet temperature, T16°C1200700–1200
Compression ratio, Rp-103–15
Air preheater temperature, T15°C377347–427
Table 8. Thermodynamic analysis of Brayton cycle components for the base study.
Table 8. Thermodynamic analysis of Brayton cycle components for the base study.
ComponentPower/Heat
Transfer Rate (kW)
Exergy Destruction
Rate (kW)
Exergy
Efficiency (%)
Compressor 1103.113.586.9
Compressor 2147.210.492.9
Combustion chamber371.7160.968.0
Gas turbine288.814.995.1
Heat exchanger 119.076.455.9
Heat exchanger 2274.138.074.9
Table 9. Thermodynamic properties of the base study.
Table 9. Thermodynamic properties of the base study.
StateFluid m ˙
(kg/s)
P
(bar)
T
(°C)
h
(kj/kg)
s
(kj/kgK)
ex
(kj/kg)
0Water01.0025.0104.90.3670
1Sewage2501.0025.5107.00.3748.996
2Sewage2451.1525.6107.40.3769.012
3Sewage245.51.1825.7107.90.3775.021
4Sludge1.0510.9025.1106.30.369771.1
5Sewage241.71.7125.3106.20.3713.064
6Sludge4.18913.7725.1106.50.36857.7
7Sludge0.41438.2025.1106.00.368514.1
8Treated Water237.51.7125.3106.20.3710.9473
9Sludge1.4646.0025.1105.60.368698
10Digestate1.0281.0532.0134.20.464642.1
11Digestate0.20511.1027.0113.30.3952824
12Biogas0.0097761.0035.0−4627.211.69151,382
13Air0.45131.0125.0−164.56.9540
14Air0.451310.13337.7161.67.032303
15Air0.45139.62376.9203.97.114320.8
16Exhaust gas0.46119.141200.079.48.3151054
17Exhaust gas0.46111.10706.6−546.98.424395.6
18Exhaust gas0.46111.07672.6−588.28.390364.4
19Exhaust gas0.46111.01153.8−1182.67.50135.15
20Water0.0892830.0091.9387.51.21332.82
21Water0.0892830.00500.03457.27.2361307
22Water0.089280.7591.82582.27.236431.9
23Water0.089280.7591.8384.41.21329.79
24Air2.1871.0125.0−164.56.9540
25Air2.1871.5771.2−117.36.97540.97
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Erguvan, M.; MacPhee, D.W. Can a Wastewater Treatment Plant Power Itself? Results from a Novel Biokinetic-Thermodynamic Analysis. J 2021, 4, 614-637. https://doi.org/10.3390/j4040045

AMA Style

Erguvan M, MacPhee DW. Can a Wastewater Treatment Plant Power Itself? Results from a Novel Biokinetic-Thermodynamic Analysis. J. 2021; 4(4):614-637. https://doi.org/10.3390/j4040045

Chicago/Turabian Style

Erguvan, Mustafa, and David W. MacPhee. 2021. "Can a Wastewater Treatment Plant Power Itself? Results from a Novel Biokinetic-Thermodynamic Analysis" J 4, no. 4: 614-637. https://doi.org/10.3390/j4040045

Article Metrics

Back to TopTop