[go: up one dir, main page]

Next Article in Journal
The Co-Inhibiting Effect of Managerial Myopia on ESG Performance-Based Green Investment and Continuous Innovation
Previous Article in Journal
Are European Trees Prepared for Being Limited by Other Nutrients than Nitrogen?
Previous Article in Special Issue
Potential Power Output from Vehicle Suspension Energy Harvesting Given Bumpy and Random-Surfaced Roads
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Estimating Sludge Deposition on the Heat Exchanger in the Digester of a Biogas Plant

by
Tomasz Janusz Teleszewski
1,* and
Leszek Hożejowski
2
1
Department of HVAC Engineering, Bialystok University of Technology, 15-351 Bialystok, Poland
2
Faculty of Management and Computer Modelling, Kielce University of Technology, al.Tysiąclecia P.P. 7, 25-314 Kielce, Poland
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(18), 7981; https://doi.org/10.3390/su16187981
Submission received: 27 June 2024 / Revised: 4 September 2024 / Accepted: 10 September 2024 / Published: 12 September 2024
(This article belongs to the Special Issue Sustainable and Renewable Thermal Energy Systems)

Abstract

:
The presented research addresses a problem occurring in a biogas plant, which we know plays an important role in sustainable development. The sludge deposited on the walls of the digester’s heat exchanger impairs heat transfer to the substrate. It leads to a temperature drop inside the biogas plant and threatens its correct operation. The thickness of the sludge layer cannot be directly measured when the plant is operating. Therefore, the aim of this work was to develop and then validate a method for estimating, based on the operating parameters of the exchanger, the thickness of the sludge layer and to give theoretical foundations for designing an automatic sludge monitoring system. Two mathematical models (and methods) were developed: one- and two-dimensional. The former model was solved analytically while the latter by the Trefftz method. The numerical results from these two approaches showed very good agreement with each other and with the actual measurement taken directly after removing the substrate from the fermentation chamber. According to the calculation results, the growth of the sludge layer was linear with time, and its rate was 0.0064 mm per day. Finally, a schematic diagram of an intended sludge monitoring system was proposed. It could optimize biogas plant operation and thus become a step towards more sustainable energy production.

1. Introduction

The sustainable development of both rural and urban areas is increasingly related to the production of electricity and heat in biogas plants [1,2,3,4,5]. Biogas plants also constitute an important element of the circular bioeconomy [6] and can be used to stabilize the energy system [7]. A typical agricultural biogas plant in Poland, most often used for production of electricity and heat, consists of the following elements: preparation tank, anaerobic digester, biogas storage, and cogeneration unit [8]. The basis for the proper operation of the fermentation tank is a constant temperature inside the chamber, which is usually around 40 °C [9]. A drop or an excessive increase in temperature in the digester causes a reduction in methane production, which may also affect the stability of the energy system [10].
The most common problems with maintaining the right temperature may occur during winter in temperate climates [11]. The greatest heat losses are generated by the pneumatic roof of the gas chambers and amount to approximately 90% of all heat losses [12]. Insulating a pneumatic roof is problematic. In order to ensure the required temperature inside the digester chamber of the biogas plant, special attention should be paid to the operation of the heat exchanger used to heat the substrate.
The most commonly used heat exchangers in fermentation chambers are pipe-shaped heat exchangers, which are mounted on the walls of the fermentation chambers using clamps. In the process of designing a biogas plant, the size of the heat exchanger is selected to match the size of the fermentation chamber by determining the appropriate surface area of the heat exchanger walls and determining the heat exchanger’s supply temperature. It should be noted here that the supply temperature of the heat exchanger cannot be too high because high temperatures may lead to the death of methane-producing microorganisms. A serious problem in the operation of the heat exchanger is the sludge deposited on its outer walls, which causes additional thermal resistance to heat exchange [12]. In the literature on the subject, we have not found works reporting such problems concerning digesters. There are many publications devoted to problems with biogas plants connected with temperature, not fouling [13,14]. Additionally, unlike in the present paper, researchers have focus their attention mostly on the deposits inside a heat exchanger, e.g., [15,16,17,18,19,20,21]. One paper [22] stands out from these publications as it addresses the problem of the sediment occurring both inside and outside the digester. However, the study assumed a constant (time-independent) thickness of the outer sludge layer and determined its growth inside the exchanger, so comparison to the present research would be problematic.
When the thickness of the sludge deposit on the walls of the heat exchanger is so large that further heating of the substrate will be significantly hampered by the thermal resistance of the sludge, the heat exchanger is cleaned. Cleaning the heat exchanger involves stopping the operation of the biogas plant and replacing the entire substrate inside the fermentation chamber.
Due to the fact that, in a working digester, it is not possible to measure the thickness of the sludge on the walls of the heat exchanger, a reasonable solution is a mathematical calculation based on the operating conditions of the heat exchanger. In terms of applied mathematics, such a problem is referred to as an inverse geometric problem.
To solve the aforementioned problem, the presented research aimed to propose and then validate with actual data a computational method for estimating the thickness of the sludge deposit on the outer walls of the heat exchanger inside a fermentation chamber during the fermentation process. The method is intended for the design of an automatic control system for monitoring the growth of the sludge layer. The installation of such a system is being considered in the future. It would certainly help to optimize the production of biogas and, to a certain extent, make the process of production less energy-consuming, thus contributing to the sustainable development of biogas plants.
The following sections outline the remaining content of this paper. Section 2 describes in detail the fermentation chamber in the biogas plant. Then, two mathematical models for the considered problem are built, each with the appropriate solution method for determining the sludge layer thickness. Additionally, the overall heat transfer coefficient is determined. Section 3 presents the numerical results and their discussion. The last section contains the conclusions of this work.

2. Materials and Methods

2.1. Subject of This Research

The subject of this research was the heat exchanger of the fermentation chamber (Figure 1 and Figure 2) of a 1 MW biogas plant located in north-eastern Poland, which has a temperate climate, where winters can be both frosty and mild, and summers can be hot or rainy. The tested biogas plant normally operated continuously for a three-year period, from the moment the biogas plant was turned on until it was turned off, in order to clean the fermentation chamber and the heat exchanger. Fermentation chambers were filled with substrate consisting of corn silage, poultry manure, and potato pulp. Due to problems with maintaining the proper temperature in the tested digesters during the winter, modernization works were carried out in August 2017 by adding additional heat exchanger loops inside the digester.
The length of a single loop of the tubular heat exchanger was 90 m, with an internal diameter d1 and external diameter d2 equal to 56.3 mm and 60.3 mm, respectively. The heat exchanger pipe was made of austenitic corrosion resistant steel with a thermal conductivity kp of 15 W/(m K). The average values of the thermal conductivity kw, the dynamic viscosity coefficient μw, and the specific heat capacity of the heating medium (water) cpw in the heat exchanger were assumed to be 0.64 W/(m K), 0.000509 kg/(m s), and 4180 J/(kg K), respectively. Based on the literature [23,24,25,26], the following substrate and sludge parameters were assumed: substrate thermal conductivity ks equal to 0.62 W/(m K), substrate specific heat cps equal to 4184 J/(kg K), substrate velocity vs equal to 0.005 m/s, substrate dynamic viscosity coefficient μs equal to 0.03 kg/(m s), and thermal conductivity of the sludge ksl equal to 0.8 W/(m K). Due to the use of demineralized water in the heating installation, it was assumed that there was no fouling layer inside the heat exchanger.
The heat exchanger’s operating parameters were measured using a Kamstrup heat meter with the following accuracies: temperature (PT500 temperature sensors manufactured by Kamstrup (Skanderborg, Denmark)) ± 0.3 °C, flow (Ultraflow flowmeter manufactured by Kamstrup) with a maximum error of 5% of the measured value. The average flow rate of the heating medium in winter through the heat exchanger loop was 0.29 kg/s.
Figure 2 shows a diagram of the fermentation chamber, photographs of a fragment of the heat exchanger with sludge after a three-year period of operation, and the same fragment after cleaning.
To assess the thickness of the sludge on the outer wall of the heat exchanger, measurements were made at 30 points, using an Insize 1141 depth gauge. Figure 3 shows a boxplot for the measurement data. Based on this, the average value of the deposits layer on the outer walls of the heat exchanger was 7.1 mm (August 2020). It should be noted that the measurement was made after removing the substrate from the fermentation chamber because measuring the sludge thickness during the operation of the biogas plant was not possible.

2.2. Model of Heat Transfer in the Heat Exchanger

The mathematical model adopted to describe thermal processes in the considered heat exchanger had the following assumptions: steady-state conditions, fully developed flow, incompressible liquid (substrate), and negligible viscous dissipation. Heat exchange between the fermentation chamber and its surroundings was not taken into account.
The heating pipe that runs along the walls of the fermentation chamber creates loops so large in diameter that the pipe can be thought of as straight. The geometric model of the considered heat exchanger was therefore a long hollow cylinder with the outer shell being a layer of sludge, as shown in Figure 4.

2.2.1. Determining the Sludge Layer Thickness: One-Dimensional Approach

In the one-dimensional approach, it was assumed that the temperature varies only in the radial direction. In this case, the following heat equation was adopted:
d d r r d T d r = 0 ,
and it is satisfied by both the steel pipe temperature (Tp) and the sludge temperature (Tsl).
The following boundary conditions were assumed for Equation (1), corresponding to the physics of the problem:
T p = T w      at r = r 1 ,
k p d T p d r = q      at r = r 1 ,
T p = T s l      at r = r 2 ,
k p d T p d r = k s l d T s l d r    at r = r 2 ,
where Tw is water temperature, q is heat flux density, and kp and ksl are the thermal conductivity of the steel pipe and the sludge, respectively.
It should be noted that strict compliance with the model assumptions requires that the water temperature must be constant. The temperature actually drops as water flows through the pipe. Therefore, to ensure a constant Tw in boundary condition (2), the arithmetic mean of the temperatures at the inlet (Tw,i) and outlet (Tw,o) of the heat exchanger was adopted.
The total heat flux Q is given by the following equation:
Q = m c p w Δ T = m   c p w ( T w , o T w , i ) ,
where m and cpw are the mass flow and specific heat of the heating medium, respectively. For simplicity, it was assumed that the heat flux density q is constant and is described by the following relationship:
q = Q A ,
where A denotes the inner surface of the pipe.
Analytically integrating the governing Equation (1), the following solutions were obtained:
T p = C 1 ln r + C 2 ,
T s l = D 1 ln r + D 2 ,
The constants C1, C2, D1, and D2 are specified using boundary conditions (2)–(5).
After determining the temperature Tsl, the sludge thickness s can be determined on the basis of the convective boundary condition at the interface between the sludge layer and the substrate at temperature Ta:
k s l d T s l d r = h 0   ( T s l T a ) at r = r 2 + s ,
The external convective heat transfer coefficient h0 present in Equation (10) can be determined using the Churchill–Bernstein correlation [27]:
h 0 = N u k s d 3 at d 3 = 2 r 2 + 2 s ,
N u = 0.3 + 0.62 Re 1 / 2 Pr 1 / 3 1 + 0.4 / Pr 2 / 3 1 / 4 1 + Re 282000 5 / 8 4 / 5 , Re Pr 0.2 ,
h 0 = N u k s d 3 at d 3 = 2 r 2 + 2 s ,
Re = ρ s v s d 3 μ s ,
Pr = c p s μ s k s ,
where Re denotes the Reynolds number, and Nu is the Nusselt number.
Combining (9) and (10) gives
k s l D 1 r 2 + s = h 0   ( D 1   ln ( r 2 + s ) T a )
Equation (16) should be solved for s, and it can be solved only numerically. The preliminary results concerning the solutions to Equation (16) were presented in [28].

2.2.2. Determining the Sludge Layer Thickness: Two-Dimensional Approach

In the second approach, reference was made to the previous mathematical model in terms of geometric assumptions (Figure 4) but without neglecting temperature changes in the axial direction (z). Assuming an exponential water temperature profile, the following formula for Tw can be written:
T w ( z ) = T w , i T w , o T w , i z / L
Now, considering a small pipe element with length ∆z, it was assumed that Equations (6) and (7) refer locally to this tubular element, with Δ T = T w ( z + Δ z ) T w ( z ) and A = 2 π   r 1 Δ z . In this case, Equation (7) gives the average q over the length ∆z of the pipe element. Taking the limit ∆z → 0, the heat flux at point z was obtained:
q ( z ) = lim Δ z 0   m   c   Δ T   2 π   r 1 Δ z = m   c   T w , i   T w ,   o T w ,   i z / L ln T w , o T w , i 2 π   r 1 L
The proposed two-dimensional mathematical model of heat transfer is described by
2 T r 2 + 1 r T r + 2 T z 2 = 0
which is satisfied by both Tp and Tsl so that T has two meanings, depending on r. The boundary conditions for Tp and Tsl are as follows:
T p = T w      at r = r 1   , 0 < z < L ,
k p T p r = q      at r = r 1   , 0 < z < L ,
T p = T s l      at r = r 2   , 0 < z < L ,
k p T p r = k s l T s l r    at r = r 2   , 0 < z < L ,
Similar to the first approach, the sludge thickness is determined on the basis of convective boundary condition (10), with the difference being that temperature Tsl now also depends on z.
It was decided to choose the Trefftz method for solving (19)–(23). The reason for this choice was in the numerous advantages of this method as well as its proven effectiveness on a wide range of heat transfer problems [29,30,31,32]. The Trefftz method, named after the German mathematician Erich Trefftz, applies to linear partial differential equations. It allows for approximating solutions through certain basis functions that satisfy the governing equation of a considered problem but at the cost of approximate fulfillment of the boundary conditions.
Two types of basis functions are used for the governing Equation (19) [31], Vn(1)(r, z), and Vn(2)(r, z), given by the following formulas:
V n ( 1 ) ( r   ,   z ) = k = 0 [ n / 2 ] ( 1 ) k 4 k ( k ! ) 2 ( n 2 k ) ! r 2 k z n 2 k
and
V 0 ( 2 ) ( r   ,   z ) = ln r V 1 ( 2 ) ( r   ,   z ) = z ln r V n ( 2 ) ( r   ,   z ) = 2 n 1 n 2 z V n 1 ( 2 ) ( r   ,   z ) r 2 + z 2 n 2 V n 2 ( 2 ) ( r   ,   z ) + 2 r n 2 V n ( 1 ) ( r   ,   z ) r
In order to find an approximate solution to problem (19)–(23), temperature Tp was expressed as a linear combination of the basis functions as follows:
T p T ˜ p = n = 0 N C n V n ( 1 ) + n = 0 N D n V n ( 2 )
The coefficients Cn and Dn were determined by best matching the solution (26) to the assumed conditions (20) and (21) at the boundaries. “Matching” should be understood in the variational sense as minimization of the error functional. The functional to be minimized has the form
Φ ( T ˜ p ) = Φ ( C 1 , , C N , D 1 , , D N ) = T ˜ p ( r 1 , z ) T w ( z ) 2 + q ( z ) k p T ˜ p ( r 1 , z ) r 2
After approximating Tp, temperature Tsl was approximated in the same way to best meet boundary conditions (22) and (23).
In the two-dimensional model, condition (10) should be rewritten using partial differentiation:
k s l T s l r ( r 2 + s , z ) = h 0   ( T s l ( r 2 + s , z ) T a )
Equation (28) needs to be solved with respect to s, assuming z is fixed. The solutions of Equation (28) for different values of z represent the sludge thickness varying throughout the exchanger pipe. In the last step, the obtained s should be averaged to result in a single value, as we can expect from the actual measurement.

2.2.3. Overall Heat Transfer Coefficient

After determining the thickness of the sludge layer, the overall heat transfer coefficient can be determined:
U = 1 h i + d 1 2 k p ln d 2 d 1 + d 1 2 k s l ln d 3 d 2 + d 1 d 3 1 h 0 1 ,
The internal convective heat transfer coefficient hi is determined from the Dittus–Boelter correlation [27]:
h i = N u d 1 k w d 1 ,
N u d 1 = 0.023 Re d 1 4 / 5 Pr 0.3 ,     Re d 1 10000 ,      0.6 Pr 16000 ,
Pr = c p w μ w k w ,
Re d 1 = 4 m π d 1 μ w ,
Obviously, the formula for U applies to both the one-and two-dimensional approaches.

3. Results and Discussion

Significant heat losses through the walls of the fermentation chamber and especially through the pneumatic roof [11] occur in the winter months. Therefore, the heat exchanger in the fermentation chamber operates at maximum thermal power during this period. This is the reason why the calculations and validation of the proposed mathematical models are based on data from the winter months, namely January, February, and December in 2017–2020.
Table 1 shows the results of the sludge thickness estimation based on the measured average monthly temperatures inside the fermentation chamber and average monthly operating temperatures of the heat exchanger. During the period under study, the circulation pump worked at the maximum constant rotational speed of the pump impeller, and the average flow of water in a single loop of the heat exchanger was 0.29 kg/s.
As can be seen in Table 1, both mathematical models give similar results for the sludge layer thickness, with absolute differences of 0.17 mm at most.
Figure 5 shows the calculated thickness of the sludge on the walls of the digester heat exchanger in the months of January, February, and December in 2017–2020. Also, it contains the point of starting the digester after thoroughly cleaning the heat exchanger (August 2017) and the point of turning off the digester and cleaning the heat exchanger (August 2020). It must be noted that the sludge thickness at the initial time (August 2017) was assumed to be zero (s = 0), while at the point of shutting down the digester for cleaning (August 2020), the measured thickness was 7.1 mm on average. The thickness of 7.1 mm compared to 12 mm reported in the previous study [28] is 40.8% thinner. Two reasons may explain this difference. Firstly, in 2017, two additional loops were added to the heat exchanger in order to lower the supply temperature in the heat exchanger, and, as shown in [28,33], a lower temperature translates into thinner layer of sludge deposit. Secondly, the present study referred to the operation time of the biogas plant, which was shorter than that before 2017.
Based on the calculations performed, one can observe an almost linear increase in the sludge thickness on the heat exchanger walls. This conclusion follows not from an observation alone but primarily from the coefficient of determination R2 being greater than 0.96 (a very good fit of the computed values to the regression line). The linear function describing this relationship was obtained by the least squares method, and it reads
s = 0.0064 t ,
where t denotes time measured in days.
The coefficient of t in (34) was estimated as 0.0064, which means that each increase in time by one day caused an increase in the sludge layer thickness by 0.0064 mm. Furthermore, Formula (34) enables the prediction of the sludge thickness. The value of s is equal to 7.03 mm for t = 1096 days, which agrees with the measurement (see Figure 3).
Figure 6 shows the results of calculating the overall heat transfer coefficient of the heat exchanger in the digester for various sludge thicknesses. The calculations assumed a constant temperature of 39.5 °C inside the fermentation chamber and a supply temperature of 55 °C. As the thickness of the sludge on the piping walls increases, the overall heat transfer coefficient decreases. The value of the overall heat transfer coefficient is approximately two times lower for a sludge thickness of 7.1 mm than for the wall without sludge. The relationship between the overall heat transfer coefficient and the thickness of the deposit on the heat exchanger walls can be well approximated with a third-degree polynomial, which has the following form for the 2D model:
U = 0.0641 s 3 + 1.3197 s 2 13.321 s + 115.84
The value of the overall heat transfer coefficient for the case without sludge deposit was 115.93 W/(m2 K); after three years of using the digester, it decreased to 67.8 W/(m2 K).
A paper [22] refers to a digester located in Italy, and it presents the results of calculating the overall heat transfer coefficient as a function of the fouling thickness inside the pipe, assuming a constant thermal resistance of the fouling outside the pipe of 0.002 m2 K/W. As shown by the authors of [22], the general trend in the overall heat transfer coefficient decreases with an increasing amount of fouling, obviously caused by the additional thermal resistance of the deposit. Furthermore, the growth in the outer fouling alone resulted in a drop in the overall heat transfer coefficient by 27%. To make a comparison to the present study, the same percentage decrease in the overall heat transfer coefficient would be caused by a 3.2 mm sludge layer on the outer surface of the pipe.
Figure 7 shows the thermal power of a single heat exchanger loop and the temperature at the heat exchanger outlet versus the thickness of the sludge on the outer walls. As the thickness of the sludge on the heat exchanger walls increases, the temperature at the exchanger’s outlet increases, while the heating power decreases, which is consistent with the trends in the thermal power and outlet temperature shown in [12]. Monitoring the sludge layer on heat exchanger walls would help maintain a relatively stable temperature inside the digester; this should be desirable because too low or too high temperature in the digester causes a decrease in methane production [10,34,35]. Additionally, fluctuating changes in temperature might kill the methane-producing microorganisms [35].
Currently, the intensive development of automatic control systems equipped with microprocessor controllers has allowed the use of various algorithms controlling the operation of heating devices and heat transport pipelines. An exemplary algorithm for controlling deposits inside heat exchangers in a city heating network is presented in [18]. The algorithm presented in this work can also be used to constantly monitor the level of sludge in the heat exchanger of a biogas plant. Figure 8 shows the proposed scheme for controlling the sludge layer thickness on the heat exchanger walls. The main elements of the sludge monitoring system are a flow meter (1), sensors for measuring the substrate temperature (2), the supply temperature (3), and the return temperature (4) from the heat exchanger piping. The parameters of the substrate, sludge, and water flowing through the pipe should also be entered into the controller (5).

4. Conclusions

The presented research concerns the problem of identifying the thickness of the sludge layer accumulated on the outer wall of a heat exchanger—a problem rarely addressed in the literature. It is worth emphasizing that the results obtained and shown in this paper are based on the operating parameters of an actually existing biogas plant, not on simulations alone.
From a computational point of view, the Trefftz method demonstrated its potential applicability to inverse geometric problems. But, more importantly, from an engineer’s perspective, each of the approaches to the considered problem (one- and two-dimensional) proved to be effective in estimating the sludge layer thickness. The results provided by them are consistent. Since the calculations yielded a mathematical (linear) dependence between the heating time and the sludge layer thickness, it can be used to predict, in quantitative terms, the growth in the sludge deposited on the walls of a heat exchanger. This in turn provides a biogas plant operator with the information necessary to make decisions about cleaning the digester.
Monitoring the thickness of the sludge deposited on the walls of the heat exchanger in a fermentation chamber is particularly important in the winter when there is a risk of underheating of the fermentation chamber, which reduces biogas production. The findings of this study allow for developing a framework for a monitoring system. A schematic diagram of such a system is included in this work.
Controlling the degree of contamination of the heat exchanger surface is one of the tools for planning the stable operation of the biogas plant as a renewable energy source, which is an integral element of sustainable development.

Author Contributions

Conceptualization, L.H. and T.J.T.; methodology, L.H.; software, L.H.; validation, L.H. and T.J.T.; formal analysis, L.H. and T.J.T.; investigation, L.H. and T.J.T.; resources, L.H. and T.J.T.; data curation, L.H. and T.J.T.; writing—original draft preparation, L.H. and T.J.T.; writing—review and editing, L.H.; visualization, L.H. and T.J.T.; supervision, L.H. and T.J.T.; project administration, T.J.T.; funding acquisition, T.J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was executed with resources of the statutory work financed by the Ministry of Science and Higher Education in Poland (WZ/WB-IIŚ/8/2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Ainner surface of the pipe
cpsspecific heat capacity of the substrate (J/(kg K))
cpwspecific heat capacity of the heating medium (water) (J/kg/K)
Cncoefficient (real number)
ddiameter (m)
Dncoefficient (real number)
hiinternal convective heat transfer coefficient (W/(m2 K)
hoexternal convective heat transfer coefficient (W/(m2 K)
kpthermal conductivity of the steel pipe (W/(m K))
kslthermal conductivity of the sludge (W/(m K))
ksthermal conductivity of the substrate (W/(m K))
kwthermal conductivity of the heating medium (water) (W/(m K))
Llength of heat exchanger in a single loop (m)
mmass flow (kg/s)
nindex (nonnegative integer)
Nnumber of basis functions
NuNusselt number
PrPrandtl number
qheat flux density (W/m2)
Qtotal heat flux (W)
rradius, radial coordinate (m)
ReReynolds number
R2coefficient of determination
ssludge thickness (m or mm)
ttime (day)
Tasubstrate temperature (°C)
Tppipe temperature (°C)
Tslsludge temperature (°C)
Twwater temperature (°C)
Tw,iaverage temperature at the inlet to the heat exchanger (°C)
Tw,oaverage temperature at the outlet from the heat exchanger (°C)
Uoverall heat transfer coefficient (W/(m2 K))
Vn(1)1st kind basis function
Vn(2)2nd kind basis function
vssubstrate velocity (m/s)
zaxial coordinate (m)
Greek symbols
μsdynamic viscosity coefficient of the substrate (kg/(m s))
μwdynamic viscosity coefficient of the heating medium (water) (kg/(m s))
ρssubstrate density (kg/m3)
Φfunctional

References

  1. Ouedraogo, N.I.G.; Konaté, Y.; Sawadogo, B.; Beré, E.; Sodré, S.; Karambiri, H. Characterization and Methanogenic Potential Evaluation of Faecal Sludge: Case of the Kossodo Biogas Plant in Ouagadougou. Sustainability 2023, 15, 16401. [Google Scholar] [CrossRef]
  2. Asensio-Ramos, M.; Melián, G.V.; Padrón, E.; Hernández, P.A.; Pérez, N.M.; Peraza Cano, J.L. Tracer Gas Method Evaluation for Assessing the Energy Potential of Biogas from Chicken Farms in the Canary Islands. Sustainability 2024, 16, 4168. [Google Scholar] [CrossRef]
  3. Ferreira, J.; Santos, L.; Ferreira, M.; Ferreira, A.; Domingos, I. Environmental Assessment of Pig Manure Treatment Systems through Life Cycle Assessment: A Mini-Review. Sustainability 2024, 16, 3521. [Google Scholar] [CrossRef]
  4. Dębowski, M.; Zieliński, M.; Kazimierowicz, J.; Walery, M. Aquatic Macrophyte Biomass Periodically Harvested Form Shipping Routes and Drainage Systems in a Selected Region of Poland as a Substrate for Biogas Production. Appl. Sci. 2023, 13, 4184. [Google Scholar] [CrossRef]
  5. Czubaszek, R.; Wysocka-Czubaszek, A.; Wichtmann, W.; Zając, G.; Banaszuk, P. Common Reed and Maize Silage Co-Digestion as a Pathway towards Sustainable Biogas Production. Energies 2023, 16, 695. [Google Scholar] [CrossRef]
  6. Meisterl, K.; Sastre, S.; Puig-Ventosa, I.; Chifari, R.; Martínez Sánchez, L.; Chochois, L.; Fiorentino, G.; Zucaro, A. Circular Bioeconomy in the Metropolitan Area of Barcelona: Policy Recommendations to Optimize Biowaste Management. Sustainability 2024, 16, 1208. [Google Scholar] [CrossRef]
  7. Arasto, A.; Chiaramonti, D.; Kiviluoma, J.; van den Heuvel, E.; Waldheim, L.; Maniatis, K.; Sipilä, K. Bioenergy’s Role in Balancing the Electricity Grid and Providing Storage Options—An EU Perspective; Task 41P6; IEA Bioenergy: Paris, France, 2017; Volume 2017. [Google Scholar]
  8. Gadirli, G.; Pilarska, A.A.; Dach, J.; Pilarski, K.; Kolasa-Więcek, A.; Borowiak, K. Fundamentals, Operation and Global Prospects for the Development of Biogas Plants—A Review. Energies 2024, 17, 568. [Google Scholar] [CrossRef]
  9. Uzodinma, E.O.U.; Ofoefule, A.U.; Eze, J.I.; Onwuka, N.D. Optimum Mesophilic Temperature of Biogas Production from Blends of Agro-Based Wastes. Trends Appl. Sci. Res. 2007, 2, 39–44. [Google Scholar]
  10. Wang, S.; Ma, F.; Ma, W.; Wang, P.; Zhao, G.; Lu, X. Influence of Temperature on Biogas Production Efficiency and Microbial Community in a Two-Phase Anaerobic Digestion System. Water 2019, 11, 133. [Google Scholar] [CrossRef]
  11. Teleszewski, T.J.; Żukowski, M. Analysis of Heat Loss of a Biogas Anaerobic Digester in Weather Conditions in Poland. J. Ecol. Eng. 2018, 19, 242–250. [Google Scholar] [CrossRef]
  12. Teleszewski, T.; Żukowski, M. The Influence of Sludge on Thermal Performance of Heat Exchanger Tubes Inside in an Anaerobic Digester. Annu. Set Environ. Prot. 2018, 20, 763–779. [Google Scholar]
  13. Makamure, F.; Mukumba, P.; Makaka, G. An analysis of bio-digester substrate heating methods: A review. Renew. Sustain. Energy Rev. 2021, 137, 110432. [Google Scholar] [CrossRef]
  14. Wystalska, K. Chapter 5, Creation of a systematic, descriptive review of the main sewage sludge biowaste processing technologies. In Environmental Safety of Biowaste in the Circural Economy; Neczaj, E., Grosser, A., Eds.; Wydawnictwo Politechniki Częstochowskiej: Częstochowa, Poland, 2022; pp. 91–117. [Google Scholar]
  15. Berce, J.; Zupančič, M.; Može, M.; Golobič, I. A Review of Crystallization Fouling in Heat Exchangers. Processes 2021, 9, 1356. [Google Scholar] [CrossRef]
  16. Rehman, O.u.; Ramasamy, M.G.; Rozali, N.E.M.; Mahadzir, S.; Ghumman, A.S.M.; Qureshi, A.H. Modeling Strategies for Crude Oil-Induced Fouling in Heat Ex-changers: A Review. Processes 2023, 11, 1036. [Google Scholar] [CrossRef]
  17. Ben-Mansour, R.; El-Ferik, S.; Al-Naser, M.; Qureshi, B.A.; Eltoum, M.A.M.; Abuelyamen, A.; Al-Sunni, F.; Ben Mansour, R. Experimental/Numerical Investigation and Prediction of Fouling in Multiphase Flow Heat Exchangers: A Review. Energies 2023, 16, 2812. [Google Scholar] [CrossRef]
  18. Romanowicz, T.; Taler, J.; Jaremkiewicz, M.; Sobota, T. Determination of Heat Transfer Correlations for Fluids Flowing through Plate Heat Exchangers Needed for Online Monitoring of District Heat Exchanger Fouling. Energies 2023, 16, 6264. [Google Scholar] [CrossRef]
  19. Wang, J.; Sun, L.; Li, H.; Ding, R.; Chen, N. Prediction Model of Fouling Thickness of Heat Exchanger Based on TA-LSTM Structure. Processes 2023, 11, 2594. [Google Scholar] [CrossRef]
  20. Sritham, E.; Nunak, N.; Ongwongsakul, E.; Chaishome, J.; Schleining, G.; Suesut, T. Development of Mathematical Model to Predict Soymilk Fouling Deposit Mass on Heat Transfer Surfaces Using Dimensional Analysis. Computation 2023, 11, 83. [Google Scholar] [CrossRef]
  21. Aikhuele, D.O.; Ighravwe, D.E.; Sorooshian, S. Intelligent Model for the Reliability of the Non-Intrusive Continuous Sensors Used for the Detection of Fouling-Layer in Heat Exchanger System. Appl. Sci. 2023, 13, 3028. [Google Scholar] [CrossRef]
  22. Calise, F.; Cappiello, F.L.; Dentice d’Accadia, M.; Infante, A.; Vicidomini, M. Modeling of the Anaerobic Digestion of Organic Wastes: Integration of Heat Transfer and Biochemical Aspects. Energies 2020, 13, 2702. [Google Scholar] [CrossRef]
  23. Poloski, A.P.; Bredt, P.R.; Schmidt, A.; Swoboda, R.G.; Chenault, J.W.; Gano, S.R. Thermal Conductivity and Shear Strength of K Basin Sludge; Prepared for the U.S. Department of Energy under Contract DE-AC06- 76RL01830; U.S. Department of Energy: Washington, DC, USA, 2002. [Google Scholar]
  24. Dewil, R.; Appels, L.; Baeyens, J. Improving the heat transfer properties of waste activated sludge by advanced oxidation processes. In Proceedings of the European Congress of Chemical Engineering (ECCE-6), Copenhagen, Denmark, 16–20 September 2007. [Google Scholar]
  25. Wu, B. Advances in the use of CFD to characterize, design and optimize bioenergy systems. Comput. Electron. Agric. 2013, 93, 195–208. [Google Scholar] [CrossRef]
  26. Terradas-III, G.; Pham, C.H.; Triolo, J.M.; Martí-Herrero, J.; Sommer, S.G. Thermic Model to Predict Biogas Production in Unheated Fixed-Dome Digesters Buried in the Ground. Environ. Sci. Technol. 2014, 48, 3253–3262. [Google Scholar] [CrossRef] [PubMed]
  27. Incropera, F.P.; Dewitt, D.P.; Bergman, T.L.; Lavine, A.S. Fundamentals of Heat and Mass Transfer, 6th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2006. [Google Scholar]
  28. Hożejowski, L.; Teleszewski, T.J. Determining the thickness of sludge on the heat exchanger tube inside an anaerobic digester. EPJ Web Conf. 2019, 213, 02028. [Google Scholar] [CrossRef]
  29. Grabowski, M.; Hożejowska, S.; Maciejewska, B.; Płaczkowski, K.; Poniewski, M.E. Application of the 2-D Trefftz Method for Identification of Flow Boiling Heat Transfer Coefficient in a Rectangular MiniChannel. Energies 2020, 13, 3973. [Google Scholar] [CrossRef]
  30. Piasecka, M.; Hożejowska, S.; Maciejewska, B.; Pawińska, A. Time-Dependent Heat Transfer Calculations with Trefftz and Picard Methods for Flow Boiling in a Mini-Channel Heat Sink. Energies 2021, 14, 1832. [Google Scholar] [CrossRef]
  31. Hożejowska, S.; Piasecka, M. Numerical Solution of Axisymmetric Inverse Heat Conduction Problem by the Trefftz Method. Energies 2020, 13, 705. [Google Scholar] [CrossRef]
  32. Frąckowiak, A.; Wróblewska, A.; Ciałkowski, M. Trefftz numerical functions for solving inverse heat conduction problems. Int. J. Therm. Sci. 2022, 177, 107566. [Google Scholar] [CrossRef]
  33. Pugh, S.J.; Hewitt, G.F.; Müller-Steinhagen, H. Fouling During the Use of “Fresh” Water as Coolant—The Development of a “User Guide”. Heat Transf. Eng. 2009, 30, 851–858. [Google Scholar] [CrossRef]
  34. Kosobucki, P.; Buszewski, B.; Górski, Ł. The Influence of Temperature on the Process of Dynamic Methane Fermentation of Sewage Sludge. Pol. J. Environ. Stud. 2008, 17, 369–375. [Google Scholar]
  35. Bień, J.; Wystalska, K. Sewage Sludge: Theory and Practice; Publishing House of Czestochowa University of Technology: Częstochowa, Poland, 2011. (In Polish) [Google Scholar]
Figure 1. General view of the biogas plant fermentation chamber.
Figure 1. General view of the biogas plant fermentation chamber.
Sustainability 16 07981 g001
Figure 2. Diagram of the digester with a photo of a fragment of the heat exchanger before and after cleaning the heat exchanger walls.
Figure 2. Diagram of the digester with a photo of a fragment of the heat exchanger before and after cleaning the heat exchanger walls.
Sustainability 16 07981 g002
Figure 3. Boxplot of the measured sludge thickness on the outer walls of the heat exchanger.
Figure 3. Boxplot of the measured sludge thickness on the outer walls of the heat exchanger.
Sustainability 16 07981 g003
Figure 4. Diagram of the heat exchanger adopted for the model.
Figure 4. Diagram of the heat exchanger adopted for the model.
Sustainability 16 07981 g004
Figure 5. Calculated thickness of the sludge layer on the heat exchanger walls in the winter months of 2017–2020 with added start and end points.
Figure 5. Calculated thickness of the sludge layer on the heat exchanger walls in the winter months of 2017–2020 with added start and end points.
Sustainability 16 07981 g005
Figure 6. Dependence of the overall heat transfer coefficient on the thickness of the deposits on the walls of the heat exchanger (based on the 2D model).
Figure 6. Dependence of the overall heat transfer coefficient on the thickness of the deposits on the walls of the heat exchanger (based on the 2D model).
Sustainability 16 07981 g006
Figure 7. Dependence of the thermal power of a single heat exchanger loop and the temperature at the outlet of the heat exchanger as a function of the sludge thickness.
Figure 7. Dependence of the thermal power of a single heat exchanger loop and the temperature at the outlet of the heat exchanger as a function of the sludge thickness.
Sustainability 16 07981 g007
Figure 8. Schematic diagram of sludge monitoring on the heat exchanger in a digester: 1—flow meter, 2—substrate temperature sensor inside the digester, 3—heating medium temperature sensor on the supply pipe of the heat exchanger, 4—heating medium temperature sensor on the return from the heat exchanger, 5—heat exchanger operation monitor, 6—supply pipe, 7—return pipe, 8—fermentation chamber wall.
Figure 8. Schematic diagram of sludge monitoring on the heat exchanger in a digester: 1—flow meter, 2—substrate temperature sensor inside the digester, 3—heating medium temperature sensor on the supply pipe of the heat exchanger, 4—heating medium temperature sensor on the return from the heat exchanger, 5—heat exchanger operation monitor, 6—supply pipe, 7—return pipe, 8—fermentation chamber wall.
Sustainability 16 07981 g008
Table 1. Heat exchanger operating parameters and results of calculations of the thickness of sludge on the heat exchanger walls in the winter months of 2017–2020.
Table 1. Heat exchanger operating parameters and results of calculations of the thickness of sludge on the heat exchanger walls in the winter months of 2017–2020.
DateTaTw,iTw,oS (1D Model)S (2D Model)
-°C°C°Cmmmm
December 201740.755.044.10.000.00
January 201841.955.245.80.300.14
February 201839.155.644.20.540.38
December 201842.055.848.03.543.37
January 201941.359.149.03.443.27
February 201942.058.049.03.603.44
December 201942.256.949.24.894.73
January 202042.058.150.05.675.50
February 202041.558.050.06.436.26
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Teleszewski, T.J.; Hożejowski, L. Estimating Sludge Deposition on the Heat Exchanger in the Digester of a Biogas Plant. Sustainability 2024, 16, 7981. https://doi.org/10.3390/su16187981

AMA Style

Teleszewski TJ, Hożejowski L. Estimating Sludge Deposition on the Heat Exchanger in the Digester of a Biogas Plant. Sustainability. 2024; 16(18):7981. https://doi.org/10.3390/su16187981

Chicago/Turabian Style

Teleszewski, Tomasz Janusz, and Leszek Hożejowski. 2024. "Estimating Sludge Deposition on the Heat Exchanger in the Digester of a Biogas Plant" Sustainability 16, no. 18: 7981. https://doi.org/10.3390/su16187981

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop