You are on page 1of 21

Journal of Building Engineering 85 (2024) 108475

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: www.elsevier.com/locate/jobe

An experimental study and numerical simulation of natural


ventilation in a semi-arid climate building using a wind catcher
with evaporative cooling system and solar chimney
Nadia Saifi a, b, *, Aymen Baadi a, b, Rebha Ghedamsi a, b, Abdessamed Guerrout c,
Noureddine Settou a, d
a
VPRS Laboratory, Kasdi Merbah University of Ouargla, BP 511, Ouargla, 30000, Algeria
b
Department of Mechanical Engineering, Kasdi Merbah University of Ouargla, BP 511, Ouargla, 30000, Algeria
c
Laboratory of Numerical and Experimental Modeling of the Mechanical Phenomena, University of Mostaganem, Kharouba, Mostaganem, 27000,
Algeria
d
Department of Physic, Kasdi Merbah University of Ouargla, BP 511, Ouargla, 30000, Algeria

A R T I C L E I N F O A B S T R A C T

Keywords: The energy demands of buildings are increasing, underlining the need for sustainable energy
Wind catcher sources. Passive cooling methods, such as wind towers and solar chimneys, are growing due to
Evaporative cooling system their capacity to minimize emissions and costs through natural ventilation. The main objective of
Solar chimney this study was to determine the efficiency of a passive solar cooling system, consisting of a wind
Experimental validation catcher, an evaporative cooling system, and a solar chimney, to reduce temperatures in a test
Computational fluid dynamics (CFD)
room in the semi-arid climate of Ouargla, Algeria. The system’s performance was evaluated
through an experimental study involving a series of tests to measure air velocity, temperature,
and humidity at different locations within the system. Concurrently, a numerical study was
conducted using computational Fluid Dynamics (CFD) analysis to assess the airflow within the
test cell, the wind catcher, and the solar chimney system. The results demonstrated that the wind
tower increased air velocity within the cell during windy conditions, exceeding exterior air ve­
locity, reaching up to 1.8 m/s. The solar chimney facilitated natural cooling within the room
during the day, preventing temperature stratification within the cell. The internal cell tempera­
ture decreased by approximately 3–7 ◦ C, indicating the efficiency of the passive solar cooling
system in reducing heat. The study concludes the passive solar cooling system’s success in
creating a comfortable, well-ventilated environment in a semi-arid climate. It offers insights into
energy-efficient comfort and demonstrates a practical, sustainable solution for minimizing
building energy needs.

Nomenclature

Cμ , C1ε , C2ε , C3ε Constants of the turbulence model [1/m]


Dt Diffusion coefficient of effective mass due to turbulence
u, v, w Speed component in x, y and z directions [m /s]

* Corresponding author. VPRS Laboratory, University of Kasdi Merbah Ouargla, BP 511, Ouargla, 30000, Algeria.
E-mail address: saifi.nadia2009@gmail.com (N. Saifi).

https://doi.org/10.1016/j.jobe.2024.108475
Received 29 September 2023; Received in revised form 10 December 2023; Accepted 5 January 2024
Available online 28 January 2024
2352-7102/© 2024 Elsevier Ltd. All rights reserved.
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Sh Heat source, including heat of reaction [J /mol]


SMx , SMy , SMz x, y and z components of body forces [N]
W width [m]
αε Turbulent Prandtl number for the energy dissipation velocity equation
αk Turbulent Prandtl number for turbulent Kinect energy equation
αt Vortex thermal diffusivity [m2 /s]
β Thermal expansion coefficient [1 /K]
μ Dynamic viscosity [Kg /m.s]
μeff Effective dynamic viscosity [Kg /m.s]
V Kinematic viscosity [m2 /s]
ρ Density [kg /m3 ]
λ Thermal conductivity [W/m.K]
ε Dissipation rate of turbulent kinetic energy [m2 /s3 ]

Abbreviation
ACH Air change per hour
CFD Computational Fluid Dynamics
SIMPLE Semi-Implicit Method for Pressure-Linked Equations

Subscripts and superscripts


Inl-c Inlet of solar chimney
Inl-Wc Inlet of Wind Catcher
Mid-c Middle of the solar chimney
Mid-Wc Middle of the wind catcher
Out-c Outlet of solar chimney
Out-Wc Outlet of wind catcher

1. Introduction
In semi-arid regions, wind tower systems and solar chimney systems contribute to improving natural ventilation within buildings.
Both solar chimneys and wind towers depend on internal and external components, impacting their efficiency. External factors such as
solar radiation, external wind, and local climate conditions notably influence their performance [1–4]. Optimal airflow within a solar
chimney relies on the intensity of incoming radiation. Higher incident radiation levels correspond to increased airflow within the
chimney. This correlation highlights the direct influence of radiation intensity on the ventilation efficiency of the solar chimney.
Essentially, stronger sunlight results in greater airflow due to resulting temperature variations [5–7]. However, the effectiveness of
solar chimneys and wind catchers is also affected by wind speed and direction [8,9], Even at low speeds, an opposing wind direction
can diminish the volumetric airflow rate of the chimney. Furthermore, the impact of wind blowing in the opposite direction can reduce
the volumetric airflow rate of a chimney, even when wind speeds are low [7]. Imran, A. A. et al. [10] devised an experimental and
simulation-based prototype of a solar chimney to assess its efficiency within Iraq’s environmental and lighting conditions. Within an
inclined solar chimney, natural convection generates a continuous 2D turbulent flow. This flow underwent numerical analysis across
inclination angles of 15◦ –60◦ , solar heat fluxes ranging from 150 to 750 W/m2, and chimney thicknesses of 50-100-150 mm. The
numerical model indicated that a 60-degree chimney inclination offered the most favorable ventilation rate, approximately 20 %
higher than at 45◦ . Saifi, N. et al. [11] conducted an experimental and numerical exploration of a solar chimney with varying in­
clinations. Experimental research investigated the impacts of inclinations at 30◦ and 45◦ , highlighting that 45-degree inclined
chimneys demonstrated superior thermal efficiency. Through an analysis covering inclination angles from 30 to 90◦ , the optimal angle
of 45◦ was identified to maximize the flow rate. Harris, D. et al. [12] utilized computational fluid dynamics (CFD) simulations to
explore how the inclination angle impacts the efficiency of solar chimneys. Their findings highlighted that the effectiveness of solar
chimneys is significantly affected by their tilt. Optimal airflow was attained at an angle of 67.5◦ , presenting an 11 % increase in ef­
ficiency compared to a vertical chimney. Hirunlabh, J.et al.l [13] conducted a study on home heat evacuation utilizing a Metallic Solar
Wall (MSW) in Bangkok. Comprising various components like a glass cover, an air gap, a black metal panel, microfiber insulation, and
plywood, the MSW demonstrated efficient natural ventilation when measuring 1 m in height. Among the tested configurations, the
metallic solar wall, made of a metallic material and featuring an air gap of 14.5 cm with a surface area of 2 m2, showcased the highest
mass airflow rate, ranging between 0.01 and 0.02 kg/s. AboulNaga, M. et al. [14] introduced a theoretical analysis of integrated
wall-to-roof solar chimneys aimed at enhancing nighttime ventilation in buildings. Employing a computational algorithm for a
parametric investigation, the study aimed to determine the optimal arrangement of the chimney connecting the wall and roof. Their
research revealed that a single rooftop solar chimney exposed to an average incident solar radiation of 850 W/m2 could generate an
airflow rate of 0.81 m3/s. Ong K. et al. [15] proposed a mathematical model for the solar stack and conducted experiments using a
physical model with cavity sizes ranging from 0.1 to 0.3 m. The experiments were conducted outdoors, revealing a better qualitative
alignment between experimental and theoretical results in the 0.3 m cavity area within radiation strengths of 200–650 W/m2, with no

2
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

observed airflow in the opposite direction within this range. Gan, G [16]. employed Computational Fluid Dynamics (CFD) software to
assess the efficiency of a glazed solar chimney featuring heat exchange in naturally ventilated buildings. Their study demonstrated that
as the temperature of the chimney wall increased, the expected ventilation rate also rose. Researchers examined how different types of
glass impacted solar heat absorption, revealing that using double or triple windows increased ventilation rates during cold winters.
Lee, K. H. et al. [17] introduced a new component in the Energy Plus software to simulate and compute the thermal chimneys’ energy
impact. Their findings suggested that a rise in solar absorptivity from 0.25 to 1.0 resulted in a remarkable 57 % enhancement in airflow
due to increased surface temperatures of the absorber wall. Pillai, P. et al. investigated the impact of glazing, solar flux, emissivity, and
absorber surface absorptivity on the catcher’s performance. Their experiments conducted at plate temperatures of 60 ◦ C and 100 ◦ C
indicated a direct correlation between solar absorptivity and efficiency, particularly at lower solar flux levels ranging from 200 to 600
W/m2, where solar absorptivity exceeding 0.8 demonstrated optimal performance. Wind catchers, as architectural marvels, exemplify
the harmonious coexistence of natural and man-made environments [18], ingeniously tailored to suit their respective climatic contexts
[19]. Elmualim, A. A. et al. [20] examined the impact of two cross-sectional forms (square and circular) on the efficacy of wind towers
at various wind speeds and directions. According to their findings, the ventilation performance of both types of wind towers was
primarily influenced by wind speed and the prevailing wind direction. Moreover, when exposed to the same external wind speed, it
was observed that the efficacy of a wind catcher with a circular cross-section was inferior to that of a wind tower with a square
cross-section. Gage, S. et al. [21] experimented with comparing the ventilation efficiency of two modern wind towers with different
cross-sectional shapes, specifically square (four-sided) and hexagonal, under varying wind speeds and directions. One of the wind
catcher designs had a 1:10 scale model connected to a test room positioned beneath the wind tunnel. The findings demonstrated that
the ventilation performance of the hexagonal cross-section wind tower was more consistent and reliable across locations with varying
wind angles. Conversely, the four-sided wind catcher performed well at a 45-degree angle of air incidence under dominant wind
conditions. Maneshi, M. M. et al. [22] carried out a numerical analysis to evaluate the performance of modern four-sided wind towers
with circular and square cross-sectional geometry at two wind speeds and directions, 0◦ and 45◦ . Similarly, the study revealed that the
wind tower with a square cross-sectional shape exhibited heightened efficacy when facing an air incidence angle of 45◦ . Goudarzi, N
[23]. conducted a study on indoor flow structures, crossflow ventilation, and thermal comfort in a wind catcher using numerical
simulations based on a mixed-climate scenario. The study focused on two key analyses: varying room outlet elevations (top, middle,
front inlet, floor) and altering outlet area ratios to the wall (10 %–40 %). The results indicated that reducing the room outlet elevation
significantly increased outlet mass flow rates. Optimal thermal comfort was achieved with front and floor outlet elevations. Addi­
tionally, increasing the outlet area ratio at specific elevations enhanced mass flow rates, with the 30 % aspect ratio producing desirable
temperature stratification and velocity distribution. Farouk, M [24]. investigated the airflow performance of a modern wind catcher
affected by three distinct cross-sectional shapes: square (with four sides), circular, and hexagonal. Computational Fluid Dynamics
(CFD) data revealed that altering wind direction didn’t impact the efficacy of the hexagonal wind tower. Additionally, the circular
wind trap exhibited the slowest air renewal rate and the lowest airspeed within. Ghadiri, M. H. et al. [25] assessed the ventilation
effectiveness of traditional wind towers in the arid region of Yazd, Iran, utilizing CFD tools. The study simulated the impact on air
velocity and internal temperature distribution using a four-sided square wind catcher model, ranging in height from 3.5 to 10.5 m.
Calculations demonstrated that the tower height significantly influenced the wind catcher’s ventilation performance, suggesting an
optimal construction height of 6 m for traditional wind towers in Yazd. Elmualim A. A [26]. analyzed a cylinder-shaped cross-section
wind catcher with various aperture configurations. Experimental conditions were employed to compare wind towers with
cross-sections of 2, 4, 6, and 12 segments. This experiment indicated that a wind catcher with two faces exhibited the most effec­
tiveness. Moreover, the study highlighted that an increased number of outlets in a wind tower reduced the impact of wind angle on the
airflow entering the tower. Attia, S [27]. conducted a vapor visualization test within a wind tunnel to observe airflow patterns in a
room equipped with a single wind catcher at different wind directions. At an external wind speed of 2 m/s, the results revealed that a
single wind catcher with an exit wall featuring an opening ratio of 0.6 achieved up to 4 Air Changes per Hour (ACH). Furthermore, with
two upwind and two downwind wind catchers at the same speed and opening ratio, the setup achieved up to 5.6 ACH. Ghadiri, M. H.
et al. [28] investigated cross-ventilation induced by the wind in a structure furnished with a bilateral wind tower and an operable
window across various wind directions (0◦ –90◦ ). The study analyzed parameters including the turbulence model, domain size, and grid
resolution. Results favored the SSG Reynolds Stress turbulence model, showing the highest correlation with experimental data ob­
tained from wind tunnel tests. Zarandi, M. M [29] conducted a field survey to analyze the typology of existing traditional wind towers
in Yazd, Iran. The floor plans and arrangements of 53 different rectangular wind catchers, featuring various internal partition shapes
(X, I, +, H, and K), informed the categorization process. Numerical findings revealed that the + partition-shaped wind catcher was
most effective in reducing indoor wind speed and enhancing resident comfort. Previous studies in different climatic conditions have
explored wind catchers combined with solar chimneys, indicating that this combination enhances passive ventilation efficiency and
augments airflow [30,31]. Foroozesh, J. et al. [32] employed computational fluid dynamics (CFD) simulations and various configu­
rations, such as inclined bottoms and baffles, to propose a new wind-catcher design integrated with evaporative cooling for
eco-friendly HVAC in buildings. Results highlighted the efficacy of a 60-degree baffle in reducing indoor temperatures by 17.4◦ Celsius
and improving thermal comfort. Amid increasing concerns about building energy consumption and the environment, this study un­
derscores the potential role of natural ventilation systems in addressing these challenges. Nouanégué, H. et al. [33] conducted a
numerical analysis on the usage of a solar chimney alongside a wind catcher during periods of minimal wind. The upper entrance was
positioned counter to the wind direction, leading to a prolonged presence in the recirculation zone. The study revealed the necessity of
adjusting the aspect ratio to achieve the highest ventilation efficiency in such scenarios. Moosavi, L. et al. [34] conducted a numerical
and experimental analysis of a combined wind catcher and solar chimney system for a two-story office building in humid and arid
climates. The collected data indicated that the system achieved approximately 9 Air Changes per Hour (ACH) within the structure.

3
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Furthermore, it is anticipated that energy conservation during peak summer demand periods could reach nearly 90 %. Amatalraof, A
[35]. developed a novel passive cooling system for buildings inspired by the cooling mechanism found in camel nasal conchae,
morphologically and physiologically. In Yemen’s desert climate, clay cylinders resembling onion rings, covered with dual layers of jute
fibers, function as the core cooling units within a windbreak. Results demonstrated the success of this biologically-inspired design,
reducing air temperatures by an average of 14.6◦ Celsius and increasing relative humidity by an average of 57.5 %. Compared to other
high-stack cooling designs within wind towers, this 1.20-m-high design achieved a remarkable temperature reduction of 19.8◦ Celsius,
signifying significant cooling efficiency. Krishnan, H. et al. utilized numerical simulations and experimental setups to assess the
performance of a solar heating unit (SHU) integrated with a wind collector. Computational Fluid Dynamics (CFD) techniques revealed
that the SHU effectively elevated incoming air temperatures within the comfortable range by 5 ◦ C–23 ◦ C. Flow rate analyses
demonstrated maximum airflow (0.45 m3/s) at a 45-degree incident angle, indicating efficient ventilation. The integration of evap­
orative cooling, rooted in ancient practices, emerges as an effective strategy when combined with natural ventilation and innovative
cooling systems, such as demonstrated in studies utilizing water for low-energy cooling systems like the Windcatcher. Erell, E. et al.
[36] introduced a multi-stage vertical draft evaporative cooling tower system. This system involves water dispersion in the upper
section of the tower to cool the surrounding arid atmospheric air. The cooled, denser fresh air descends within the tower, combining
with other air currents at the secondary inlet. Further cooling occurs by evaporation when the mixed air encounters additional air
streams at the secondary entrance. The article presents empirical evidence and theoretical analysis, offering experimental findings
regarding temperature reduction, water usage, cooling power, and refrigeration power consumption. Janajreh, I. et al. [37] proposed
utilizing a wind tower to enhance occupant comfort through Trans Evaporative Cooling. As the wind tower collects wind, it introduces
water into it, leading to a decrease in temperature and an increase in humidity. By capturing the wind, moisture is absorbed, resulting
in cooler and denser air. Noroozi, A [38]. described a four-way wind tower featuring two chambers for evaporative cooling along with
a solar chimney. Through experiments and computer simulations, a prototype of the wind tower was created to assess the efficiency of
the solar chimney as a ventilation system. The research indicated that, under closed conditions, a wind speed of 3 m/s provided optimal
cooling. Additionally, the cooling loads were approximately doubled when the wind tower’s height ranged between 2.5 and 3.5 m and
the wind speed dropped below 3 m3/s within an enclosed space. In scenarios where no wind was present but the sun was shining at
1000 W/m2, the airflow rate of an evaporative cooling system averaged around 0.021 kg/s. These studies suggest that a wind catcher
integrated with an evaporation system, functioning as a zero-energy cooling system, could serve as a viable alternative to
energy-intensive mechanical refrigeration equipment.
The objective of this study is to investigate the behavior of a wind catcher equipped with an evaporative cooling system and its
impact on indoor air properties in an experimental cell equipped with a solar chimney, by calculating the air flow rate in each system
and its variation based on ambient temperature and solar radiation intensity. An experimental model on a small scale has been
developed which has been used to make several flow rate predictions for all the systems studied, even those which include connected
channels and their common flow structures. Two methods are proposed for this study, one experimental and the other numerical, using
the ANSYS 2020 version R1 calculation code. A comparison between our numerical results and the experimental study’s results
verified the findings. Finally, the impact of changes in wind speed, outside air temperature, and humidity on the performance and
efficiency of each passive cooling system connected to the experimental cell has been analyzed and discussed.

2. Methodology
This section outlines the methodology employed to investigate the integration of Wind Collector and Solar Chimney Systems to
optimize ventilation and energy efficiency within a hot and arid climate. A cooling system by evaporation has been installed at the
wind sensor level to provide an additional cooling effect. The system involves installing devices within the wind sensor, which helps
reduce the air temperature passing through it, providing a flow of fresh and refreshing air for the building. To assess the performance of
these systems, an experimental study was conducted based on measuring physical parameters. Additionally, a numerical study was
conducted following the acquisition of experimental results. A prototype structure including an integrated system was erected to
replicate conditions in the real world. Wind flow, temperature differences, and ventilation rates were measured using catchers, while
computational fluid dynamics (CFD) simulations were conducted to evaluate the system’s performance under different climatic

Fig. 1. Descriptions of the cells.

4
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

conditions. The collected data and simulation results were subjected to statistical analysis to assess the efficacy of the system.

3. Experimental procedure
3.1. Cell description
The study focused on a cell that measured 80 cm in height and had a surface area of 1 square meter. This cell’s size was determined
by halving a standard room. It features a small door measuring 47.5 × 29 cm2 and a window sized at 39.5 × 24.5 cm2. Each component
has specific requirements, including orientation and construction materials. The interior and exterior walls consist of hollow bricks
coated with mortar. The slab and floor are constructed using reinforced concrete. Additionally, there is a 50 mm thick insulating layer
made from palm fibers covering the outer concrete surface (Fig. 1).
In Table 1, we summarize the thermal properties of wall components.

3.2. Windcatcher
The wind catcher has dimensions of 200 cm in height, 15 cm in width, and 90 cm in length (Fig. 2). It is in the shape of a
parallelepiped, with walls made of 2 mm thick wood and covered with a layer of polystyrene insulation. It is open at the upper part (air
inlet section) oriented towards the east, with dimensions of 200 × 800, as well as at the lower part (air outlet section), which measures
100 × 80. The air gap inside the wind catcher has a thickness of 100.
The interior of the wind catcher was treated with wood varnish to safeguard it against moisture. Varnish, known for its moisture-
resistant properties, forms a uniform coating inside the wind catcher, slowing the wood’s degradation caused by high humidity
exposure.

3.3. Solar chimney


The primary components of the solar chimney include an iron sheet housing measuring 1 × 92 m. This chimney incorporates a
single transparent glass cover, 4 mm thick, and a slender galvanized steel absorber plate, painted matte black and 0.4 mm thick.
Positioned behind the insulator is another galvanized steel plate, with a thickness of 0.4 mm. The gap between the clear cover and the
absorber plate measures 10 mm. Rear insulation is achieved by employing two distinct polystyrene plates—one with a thickness of 40
mm and the other measuring 20 mm (Fig. 3).

3.4. Evaporative cooling system


Natural cooling is provided by wind catchers in two ways: displacement and evaporation [36]. When an airflow passes through a
humid type of media or a water source, evaporative cooling occurs naturally.
Inside the main ducts of the wind tower, a vertical fabric curtain, and water spray pipes are designed to spray water onto the
curtains. The remaining water is stored in a reservoir located at the bottom of this duct (Fig. 4).
The inlet duct of the wind catcher was covered with a rectangular curtain of fabric (Fig. 5). The curtain was made from a 0.60 m
wide and 1.60 m long jute fabric. The curtain is placed in the duct, in the direction of airflow. The curtain is positioned parallel to the
inlet duct and is installed correctly to ensure optimal wind capture performance.
Table 2 presents the main physical characteristics of jute fibers.
Water sprayers, installed at the top of the curtain, wet the curtain. The sprayers were made using medical syringes and a transparent
PVC tube. Several steps were followed to attach the medical syringes to the plastic tube. First, a small hole was made in the tube. The
syringes were glued into the holes using silicone to prevent water leakage and ensure complete sealing (Fig. 6).
At the base of the wind tower, a reservoir was placed to collect water that infiltrated through the jute fabric. The reservoir is well-
positioned to ensure efficient water accumulation. It is designed with the right size and capacity to hold the accumulated water and
allow air to enter the cell.

3.5. Sensor placement


To study the airflow behavior in the prototype, temperature and air velocity distributions, as well as humidity, were monitored at
half-hour intervals. The Fig. 7 shows the arrangement of towers for different cases, including temperature and humidity catchers.

4. Numerical part
4.1. Physical model
Within this prototype’s design (Fig. 8), both a solar chimney and a wind catcher are integrated to enhance the ventilation system’s
efficiency. The insulation, composed of polystyrene with an intervening air layer, effectively separates and prevents interactions

Table 1
Thermal characteristics of the Cell’s walls.

Walls Thickness (cm) Thermal conductivity (W/m.K)

Interior and exterior cement mortar cladding 1.5 1.4


Concrete brick 10 0.48
Reinforced concrete platform and slab 7 2.4

5
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 2. Wind catcher.

Fig. 3. Solar chimney.

Fig. 4. Evaporative cooling system.

between the two systems. The wind tower is designed to include an evaporative cooling system within its structure. To ensure optimal
function, a humidifying device regularly evaporates water within the tower. As air circulates within the prototype, it cools due to heat
extraction from the surrounding air during the water evaporation process. An independent water supply system provides the necessary
water to the humidifying device. The interior of the wind tower experiences natural cooling as air passes through a humid medium or
water source (Fig. 9). This phenomenon occurs because the air in proximity to a moist surface can absorb significant amounts of water
vapor, resulting in reduced temperature and increased air density, ultimately generating a downdraft.
To make the physical description more manageable, it is necessary to simplify certain aspects and make several assumptions to
facilitate calculations. In this study, we chose a three-dimensional domain and established the following assumptions:
A three-dimensional flow is occurring.

• The airflow is turbulent.

6
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 5. Jute fabric.

Table 2
Important physical properties of jute fibers [39].

Fiber duration 2.5 [mm]


Density 1.48–1.50 [g/cm3 ]
Moisture absorption (20 ◦ C–65 % HR) 16-18 [ %]
Water sorption rate 25.4 [ %]
Dry toughness 0.3–0.6 [N. Tex− 1 ]
Loss of wet toughness 15-25 [ %]
UV resistance Medium
Resistance to micro-organisms Medium

Fig. 6. Direct evaporative cooling system.

7
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 7. Location of speed, temperature, and humidity measurements.

Fig. 8. Physical model.

Fig. 9. Location of the evaporative cooling system.

• Fluid under consideration (humid air) is a combination of two gases—dry air and water vapor.
• Composition: incompressible Newtonian fluid.
• No chemical reaction exists between the constituents of the mélange.
• Specific heat capacity of the mixture: the average mass fraction of the two species (water vapor and air) is used to formulate the law
of mixtures.
• The absorber and the glass will always be in a parallel position.

4.2. Boundary conditions


Boundary conditions are those that are imposed on the outside of the domain under study. They’re obligatory for getting the math
problem solved accurately. There are three types of boundary conditions: inlet, outlet, and wall.

8
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

• The conditions at the system inlet mirror those at the wind catcher inlet. The condition is fixed and is proportional to the velocity of
the wind. It’s a Dirichlet, so it locks in the input value of the variable.
• The absorber, the glass, and the outlet are the three boundaries that the solar chimney must adhere to. The Discrete Transfer
Radiation Model (DTRM) technique is used to account for boundary conditions at the walls of a solid. Integrating the intensity of
radiation that reaches a given point on the wall surface yields the incident radiative heat flux. The pressure at the solar chimney’s
outlet is equal to the surrounding air.
• The absence of heat transfer to or from the environment is what makes other walls adiabatic.
These boundary conditions are essential for a complete definition of the problem and a rigorous mathematical solution.

4.3. Governing equations


The continuity equation, the conservation of momentum equation, and the conservation of energy equation are used to characterize
a heat transfer and fluid flow problem.
Continuity Equation
∂u ∂v ∂w
+ + =0 (1)
∂x ∂y ∂z
Equation of Conservation of Momentum
[ ( )]
∂u ∂u ∂u 1 ∂p ∂2 u ∂2 u ∂2 u
According to (OX) : u +v +w = − + μeff + + + SMx (2)
∂x ∂y ∂z ρ ∂x ∂x2 ∂y2 ∂z2
[ ( )]
∂v ∂v ∂v 1 ∂p ∂2 v ∂2 v ∂2 v
According to (OY) : u +v +w = − + μeff + + + SMy (3)
∂x ∂y ∂z ρ ∂y ∂x2 ∂y2 ∂z2
[ ( )]
∂w ∂w ∂w 1 ∂p ∂2 w ∂2 w ∂2 w
According to (OZ) : u +v +w = − + μeff + + + + SMz (4)
∂x ∂y ∂z ρ ∂z ∂x2 ∂y2 ∂z2
Equation of Conservation of Energy.
The following is an expression for the species transport equation for water vapor in the airflow:
( ) ( 2 )
∂T ∂T ∂T ∂ T ∂2 T ∂2 T
ρh u + v + w =k + + + Sh (5)
∂x ∂y ∂z ∂x2 ∂y2 ∂z2

Where u, v, and w are the velocity components in the x, y, and z directions, respectively. T represents temperature, k is thermal
conductivity, and Sh is the heat source, including the heat of the reaction [40,41].
Turbulence model.
To simulate the turbulent regime, we used the standard K-epsilon model with two transport equations.
Turbulent kinetic energy equation:
[ ]
∇(ρKu) = ∇ αk μeff ∇K + Gk + Gb − ρε (6)

Energy dissipation rate equation:


[ ] ε ε2
∇(ρεu) = ∇ αε μeff ∇K + C1ε (Gk + C3ε Gb ) − C2ε ρ (7)
K K
C1, C2ε and C3ε representing empirical model constants (Table 3).
Species Transport Equation for Water Vapor:
The species transport equation for water vapor in the airflow is expressed as follows [41,42]:
[ ] ε ε2
∇(ρεu) = ∇ αε μeff ∇K + C1ε (Gk + C3ε Gb ) − C2ε ρ (8)
K K

With.
• YH2 O : The concentration of water vapor
• SH2 O :: The water vapor source term is due to evaporation.
• wwateraterwwateraterwwateraterwwateraterDH2 O : Water vapor’s diffusion coefficient in air.

Table 3
Constants of k epsilon model.

C ε1 C ε2 Cμ σk σε
1.44 1.92 0.09 1.00 1.3

9
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

• SCt = ρμDt t : The Schmidt number for turbulent flow.


• μt : The turbulent viscosity.
• Dt : The effective mass diffusion coefficient due to turbulence.
The efficiency of air distribution can be partly evaluated by measuring the air change rate. Air changes per hour (ACH) are
expressed as the volumetric air flow rate through the space divided by the space volume, allowing for determining how many times the
air is replaced in the room.
In many air distribution systems, the air is neither uniform nor perfectly mixed. The actual percentage of air exchanged in an
enclosure over time depends on the airflow rate of the enclosure and the ventilation methods used. In a well-mixed ventilation sce­
nario, approximately 63.2 % of the air will be exchanged after 1 h with an air change rate (ACH) of 1. To achieve pressure equilibrium,
the amount of air leaving the space must be equal to the amount entering the space [43].
3600Q
ACH = (9)
V

4.4. Mesh size design


The precision required to solve a problem depends on the number of cells used. The mesh must be tight in all areas where the
variables present strong gradients, but especially in the proximity of the walls (Fig. 10). To select the optimum mesh for obtaining
acceptable numerical results, the number of cells is varied, and it is then necessary to observe the evolution of residuals. The
convergence criterion is constituted by the relative deviations between the model’s inlet and outlet flows.
According to the data presented in Table 4, the mesh containing 2,473,609 elements was selected for which the number of cells was
acceptable to obtain the greatest precision in the results. An irregular, repetitive mesh was also used on the walls.

5. Results and discussion


5.1. Variation of global radiation
Fig. 11 presents the variation of global radiation over time.
The measurements were conducted during a clear day corresponding to (May 11, 2023). The curve of solar radiation variation over
time takes a Gaussian shape, with a maximum radiation value of 534 W/m2 at 1 p.m. and a minimum value of 114 W/m2 at 18h00, at
the end of the measurements.

5.2. Evaluation of air temperature


Fig. 12 illustrates the variation of different temperatures inside the wind catcher as a function of time on May 11, 2023.
Fig. 12 illustrates temperature curves at three selected measurement points displaying similar trends. From 9:00 to 17:00, an
overall temperature increase was observed. At the wind catcher’s inlet, the temperature measured 300.2 K at 9:00, peaking at 309.3 K
by 14:00. Subsequently, between 14:00 and 17:10, the values continued a gradual rise, reaching 309.2 K by 17:10. This rise is
attributed to solar radiation exposure on an adjacent building wall to the experimental prototype. The wall absorbed and stored solar
energy as thermal energy from 9:00 to 14:00. Emitting this stored heat, combined with decreasing solar radiation intensity and rising
environmental temperatures, moderately increased the air temperature at the tower’s inlet. Within the wind tower, temperatures
reached 302.6 ◦ C and 302.3 ◦ C between 14:00 and 17:00. Notably, the temperature drops significantly as air contacts the wet jute
cloth. The temperature gap between the tower’s inlet and midsection was 2.4 ◦ C at 10:00, increasing to 6.7 ◦ C by 16:20. Subsequently,
temperatures decreased further, hitting 26.5 ◦ C at the wind tower’s outlet by 14:00. The maximal temperature differential between the
inlet and outlet, recorded at 14:00, reached 10 ◦ C. The slow airflow through the damp jute cloth allowed moisture absorption, leading
to the lowest achievable temperature. Solar radiation entering the solar chimney warms the glass but cools the absorber, leading to a

Fig. 10. Mesh generation.

10
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Table 4
Difference between the mass flow in and out of the model and the number of iterations.

Mesh size Nodes Elements Relative difference in mass flow of air entering and exiting

0.75 338693 1748091 3.438139E-06


1 470841 2473609 2.935097E-07
1.25 666905 3502817 1.081675E-06
1.5 777789 4102885 1.068647E-06

Fig. 11. Variation of radiation over time.

Fig. 12. Variation of different temperatures inside the wind catcher over time.

temperature difference between the solar chimney’s inlet and outlet. Fig. 13 shows this to be true.
The air temperature at the inlet gradually rises from 9:00 to 13:00, maintaining stability until 14:00, and subsequently decreases
due to reduced solar radiation. After 15:00, it begins to rise again. Analysis of the temperature difference between the chimney’s entry
and exit points reveals a correlation with solar radiation; as solar input increases, so does the internal temperature. This relationship
stems from the absorber’s properties, responsible for absorbing and storing thermal energy. The temperature disparity between the
entry and exit peaks at 18.1 ◦ C around 13:00. This accumulated energy elevates the absorber’s temperature, a significant portion of
which transfers to the air within the chimney.
In Fig. 14, the temperature variation curves over time within the cell are represented.
The curve represents the temperature variation inside the cell at two locations. The first location (T7) is situated 10 cm from the
wall near the entrance of the solar chimney, at a height of 40 cm above the ground. The second location (T5) is in the middle of the
chamber.
It can be observed that both temperature values continue to increase throughout the day. Between 09h00 and 14h20, there is an
almost linear increase in temperatures as solar radiation intensity rises. Additionally, the temperature difference between the two
locations within the test cell remains almost constant during the measurement period mentioned earlier, ranging from 0.8 ◦ C at the
beginning of the measurement to 0.3 ◦ C at 14h20. The temperature difference starts to diminish after 14h20, which is attributed to the

11
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 13. Variation of different temperatures inside the solar chimney over time.

Fig. 14. Variation of different temperatures inside the cell and ambient temperature over time.

effect of the solar chimney system, where the air temperature decreases slowly due to the reduction in solar radiation and the amount
of heat stored by the absorber.

5.3. Variation of Relative Humidity


Fig. 15 depicts the change in humidity over time at both the inlet and outlet of the wind collector. It also displays the humidity value
at the midpoint of the wind collector, which is where the wet jute cloth that serves as the evaporation system is located.

Fig. 15. Variation of relative humidity inside the wind collector over time.

12
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

The relative humidity values at the wind collector’s inlet exhibit slight variations throughout the 8 h, reaching a peak of 38.3 % at
09:40 and dropping to 28.3 % by 18:00. Inside the wind collector, the relative humidity is notably higher, hitting 46.8 % at 10:20 and
42.4 % at 13:00. This discrepancy is attributed to the airflow alongside the wet jute cloth, which significantly impacts humidity levels.
Notably, the relative humidity at the wind collector’s outlet consistently records the highest values. The evaporation zone, facilitated
by the wet jute cloth, notably contributes to temperature reduction. The difference in relative humidity between the inlet and outlet of
the wind collector reaches 24 % at 11:50, peaking at 32.9 % by 17:40, which could result in considerable temperature variations within
the wind collector.
Fig. 16 presents the evolution of the relative humidity rate at the inlet and outlet of the solar chimney, showing that the shape of the
two curves remained the same.
Through the analysis of the curves above (Fig. 16), it can be observed that the relative humidity curves at the inlet of the solar
chimney over time follow the evolution of the relative humidity curves at the inlet; the curves are nearly parallel. It is noticeable that
the value of relative humidity decreases until 13h40, reaching a value of 29.6 % at the inlet and 13 % at the outlet, then increases until
the end of the measurement. The most important factor is an increase in temperature within the solar chimney’s air layer, which causes
a corresponding decrease in humidity.
The respective evolutions of relative humidity within the test cell are represented in Fig. 17.
Specifically, the curves depict the development of relative humidity at two points inside the cell: one (RH7) located 10 cm from the
wall, next to the inlet of the solar chimney, and another (RH2) located 40 cm in height from the floor. The second one (RH5) is located
in the middle of the cell. It can be observed that the relative humidity inside the cell at the two measurement points is nearly the same
at 13h10, around 42.6 %, while there was a difference in the relative humidity value at the beginning of the day.

5.4. Evaluation of air velocity


Fig. 18 illustrates the variation in velocity at the inlet of the wind collector over time.
It can be observed that the velocity varies abruptly, and sometimes a significant change in velocity occurs at the wind collector’s
inlet. Several climatic factors, including variations in solar output and wind velocity, have been linked to this shift.
Fig. 19 represents the variation in velocity in the lower part of the wind collector and near the inner wall, at a distance of 10 cm.
It can be observed that the airflow velocity decreases in the region near the inner wall of the wind collector, as the airflow moves
downward and is forced to turn horizontally towards the cell. The velocity measurement was conducted in an elbow-shaped region,
which allowed for the measurement of lower velocity and reduced turbulence.
Fig, 20 depicts the variation in velocity at the inlet and outlet of the chimney over time.
As the velocity at the chimney inlet and the amount of solar radiation both increases, so does the velocity at the chimney outlet, as
shown in Fig. 20. The temperature difference between the glass and the absorber causes a natural convection phenomenon. Chimney
air temperature rises as it absorbs the incoming energy, resulting in a greater rate of airflow.

6. Numerical simulation
6.1. Validation
In this study, experimental results and simulation results were compared to validate the numerical simulation model. The tem­
perature and relative humidity data in Fig. 21 illustrate this comparison. At 09:00, with Vinl = 0.7 m/s at the wind collector inlet, the
system was validated.
Regarding temperature values, both experimental and numerical values are similar at the time of measurement, where the air
velocity at the wind collector inlet was 0.7 m/s. The average relative error is 0.7 %, and the maximum error is 2.89 %, indicating a good
agreement between the two obtained results. The comparison of relative humidity values revealed an average relative error of 13.36 %
and a maximum error of 22.34 %.

Fig. 16. Variation of relative humidity at the inlet and outlet of the solar chimney over time.

13
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 17. Variation of relative humidity inside the test cell over time.

Fig. 18. Velocity at the inlet of the wind collector.

Fig. 19. Velocity at the outlet of the wind collector and near the inner wall of the wind collector.

The second step involves validating the numerical simulation by comparing the experimental results obtained at 14h00 with the
numerical results, the air velocity was Vinl = 0.9 m/s.
Fig. 22 depict the variation of temperature and relative humidity values at different measurement points. The results show that the
temperature variation is strongly dependent on solar radiation. It can be noted that the results are close at most measurement points,
but there are also discrepancies among some results. This difference is due to temporary weather conditions such as solar radiation or

14
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 20. Variation in velocity at the inlet and outlet of the chimney.

Fig. 21. Comparison between Experimental and Simulation Results at Different Measurement Locations for (a) Temperature and (b) Relative humidity (9h00, Vinl =
0.7 m/s).

Fig. 22. Comparison between Experimental and Simulation Results at Different Measurement Locations for (a) Temperature and (b) Relative humidity (à 14h00 et Ve
= 0.9 m/s).

15
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

wind speed. The simulation estimated the air temperature with a small margin of about 0.84 % as the average relative error and 19.06
% for the relative humidity simulation results.

6.2. Evaluation of results within the cell


6.2.1. Velocity Profiles in the cell
Fig. 23 illustrates the variation of velocity in the median vertical plane inside the test cell.
The curve illustrates a clear correlation between wind speed variations and velocity changes. Starting at 0.7 m/s at 09:00, the speed
rises to 0.96 m/s by 14:00, then gradually declines. Notably, the region near the wind catcher’s opening consistently records the
highest wind velocities at different times. This increase in wind velocity is attributed to the air’s moisture absorption and its associated
weight, which enhances air velocity and subsequently strengthens the negative buoyancy force. At the wind catcher’s exit section, the
velocity value mirrors the incoming air speed. As a result, the streamlines converge more closely, leading to an increase in speed.
Fig. 24 depicts the variation of velocity along the (y) axis in the middle part of the cell as a function of position along the (x) axis.
Very low-velocity values were obtained at the center of the cell and gradually increased to reach 1.7 m/s. This number represents
how fast air is moving as it enters the wind catcher through the outlet. The air continues to move along the centerline and turns in the
cell’s horizontal plane. In the right part, it’s noticeable that the air changed its direction and headed toward the entrance section of the
chimney. The velocity values decrease significantly at 18h00 due to the minimum wind speed.

6.2.2. Dynamic fields


Fig. 25: (a); (b); and (c)) depict the velocity fields in the studied model during different times when measurements were taken. The
air enters through the wind catcher’s entry surface, flows and distributes itself at different levels inside the cell, and eventually exits
through the solar chimney’s exit surface.
At 9h00 the velocity is also higher at the bottom of the wind tower’s exit, reaching 2.25 m/s. This main flow is distributed
throughout most parts of the room and forms two circular cells, one in the upper part of the test cell and the other in the middle area at
the bottom of the cell, as shown in Fig. 25a and b. Part of the air rises, exits the cell, and goes outward through the solar chimney. This
direction is mainly influenced by the buoyancy forces created in the solar chimney. Fig. 25c shows that due to the airflow entering the
wind tower through the entry, there are significant velocity variations near the inner wall opposite the entry, the velocity reaches 2.01
m/s. The main flow enters the test cell with a semi-laminar flow and exits with a turbulent flow.
The airflow moves vertically through the wind catcher’s exit section before turning vertically towards the chimney’s entry section.
Inside the cell, the airflow continues to move with low velocity.
There was an increase in speed and solar radiation at 14:00. The highest value for the air velocity inside the tower was 2.78 m/s,
while the lowest value was 0.2 m/s. The center of the cell displayed a recirculation zone where the air velocity was quite slow, away
from the chimney’s entry. The velocity of the air in this zone ranged from 0 m per second to 0.4 m/s. The air recirculation zone in­
creases with the wind tower’s entry velocity. The airflow circulation in the cell exhibits a large recirculation zone in the lower parts
with faster flows, and the airflow velocity varies from 0.78 m/s to 2.19 m/s. It can be deduced from this that higher air velocity at the
tower’s entry leads to better ventilation within the cell. The air velocities along the warm (absorber) wall are high and reach their
maximum in the same section, decreasing as one moves away from this wall. Heat transfer between the absorber and the chimney’s air
results in heat diffusion between these two media. Thermal gradients create density gradients, leading to vertical buoyancy forces
(Archimedes’ buoyant force) caused by fluid expansion. Consequently, as solar radiation intensity increases, the air velocity within the
solar chimney rises., rapidly evacuating air from the cell to the outside. Air movement from the wind tower toward the cell’s interior
space can be observed. A clear air velocity difference is observed in the lower XY plane adjacent to the floor. When air enters the cell
through the wind catcher’s exit, a velocity difference occurs between the bottom and center of the cell. The same observation can be
made in the XY plane near the roof and starting from the solar chimney’s entry. The velocity variation is more pronounced than that
observed at 09h00. In the central XY plane, a relative velocity difference is noticeable. It can be observed that the air velocities at
18h00 are much lower than at 09h00 and 14h00 due to the low air velocities at the wind catcher’s entry. Variations in air velocity also
occur along the X directions. Three XZ planes were considered: one in the middle and the other two adjacent to the cell’s side walls.

Fig. 23. Velocity profiles - variation of velocity at the entrance of the cell in the vertical mid-plane.

16
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 24. Velocity profiles - variation of velocity at the entrance of the cell in the horizontal mid-plane.

Fig. 25. Dynamic field at different planes.

Velocity variations were more significant near the wind catcher’s exit and less pronounced near the opposite side walls, where the air
distribution area was larger as the air made direct contact with the wall, contributing to its distribution inside the cell. No changes were
observed in the air velocity value in other interior spaces. In the lower part, the velocity reached its highest value of 0.65 m/s, while it
reached 0.33 m/s in the upper part, and it remained nearly zero in the middle part, with a value ranging from 0 to 0.38 m/s.

6.2.3. Thermal fields


The temperature profile in the wind catcher, test cell, and solar chimney for air velocities of 0.7 m/s, 0.9 m/s, and 0.4 m/s,
respectively, are depicted in Fig. 26: (a); (b); and (c), respectively. Fig. 26(a), (b), and (c) show the temperature within the wind
catcher, the test cell, and the solar chimney at air velocities of 0.7 m/s, 0.9 m/s, and 0.4 m/s at the entry of the wind catcher,
respectively.
At 9:00 a.m., the test cell exhibits nearly uniform temperature throughout the space, indicating air recirculation and a minimal
temperature gradient between incoming and outgoing air, fostering a consistent airflow pattern. The inclusion of a damp jute cloth
within the wind tower’s evaporation system notably reduces the room temperature by approximately 1.5 ◦ C. This cooling effect stems

17
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Fig. 26. Thermal field at different planes.

from heat exchange between incoming air and sprayed water, leading to water evaporation. Analysis shows a temperature decrease in
the air passing through the wind tower, dropping from 28.85 ◦ C at entry to 24.85 ◦ C midway. The deliberate slow passage of air
through the wet jute cloth ensures that the entering air attains the lowest feasible temperature. This temperature uniformity at 9:00 a.
m. within the test cell suggests that the temperature differences between entering and exiting air prompt natural convection, likely
resulting in consistent airflow distribution across the room.
Because the air moves so slowly through the wet jute cloth, the air that enters the cell has the coolest temperature that is physically
possible. The temperature inside the testing chamber was brought down to 25.55 ◦ C, which is 1.5 ◦ C lower than the temperature
outside the chamber. This occurrence takes place as a consequence of the transfer of heat between the air inside and outside the tower.
To analyze the results obtained, three planes running along the XZ direction were taken into consideration: one near the wind tower’s
exit, one in the middle, and one next to the back wall. On the first plane, the effect of the air that was being drawn in by the wind
catcher could be seen very plainly. Fig. 26a demonstrates that there is a slight temperature difference between the interior space of the
cell and the back wall as one moves closer to the back wall.
How much solar heat a thermal storage wall absorbs is affected by the solar chimney (absorber). As the sun’s rays get stronger, the
solar chimney heats up and, in turn, warms the air around it. At 14:00, researchers measured wind speed and solar radiation and found

Fig. 27. Flow patterns (streamlines): (a) At 09h00 a.m. and Vinl = 0.7 m/s; (b) at 14h00 and Vinl = 0.9 m/s; (c) at 18h00 and Vinl = 0.4 m/s.

18
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

that the two factors contributed to a 5.5 ◦ C temperature gap between the interior and exterior of the cell. Fig. 26 a and b present the
cell’s temperature distribution. More pronounced temperature variations were observed near the wind tower’s exit, while smaller
variations were noted near the opposite walls, where the temperature was 81.85 ◦ C.
at 18h00, at the end of the measurement. the temperature at the entrance and exit of the wind catcher was 27.45 ◦ C, having been
31.55 ◦ C at the entrance. An internal temperature distribution is observed, with higher temperatures at the solar chimney level and
lower temperatures at the bottom part. Their distribution is homogeneous with close values. The temperature ranged between 29.35 ◦ C
and 30.45 ◦ C. Here, there was only a 0.5–2.0 ◦ C temperature disparity between the top and lower regions. The temperature ranged
from 29.85 ◦ C to 31.35 ◦ C at a height of 0.5 m.

6.2.4. Flow streams (streamlines)


Fig. 27: (a); (b); and (c)) show the flow pattern. The incoming air descends to the bottom of the cell at different moments.
The incoming air descends to the bottom of the test cell, having passed through the wind catcher equipped with a wet jute cloth. It
then passes through the wind tower, descends to the bottom, and reaches the floor of the cell. It rises to the cell’s top due to buoyant
force and exits through the solar chimney. The increase in solar radiation boosts the air velocity within the solar chimney, enhancing
the air circulation area within the room, mainly due to intensified buoyant force.
The model allows for selecting the most suitable ventilation speed at different times. The incoming air into the cell interacts with
the air movement generated by natural convection in the chimney’s air layer. Depending on the airflow location within the cell, the
primary flow at 14h00 yields the best result. Indeed, at this time, it produces an air distribution throughout the cell without the
presence of air vortexes found in the primary flow at 09h00.

6.2.5. Relative humidity


Fig. 28: (a); (b); and (c)) show the distribution of relative humidity in the model. The relative humidity profiles obtained from the
CFD analysis at different times of the day are presented.
Evaporative coolers are a type of cooling system that works by lowering both the temperature and the relative humidity of the air
within the hive. This cooler works by converting liquid into vapor using the thermal energy that is removed from the air as it moves
through it. This reduces the temperature of the air while simultaneously raising the air’s relative humidity. Thus, the increase in air
humidity is one of the advantages of the system installed inside the wind catcher in hot and dry regions. As there are sources of
humidity production inside the tower, the difference in the relative humidity value compared to outdoor conditions will be significant.
We observe a continuous increase in the value of relative humidity. Its value inside the cell at the beginning of the day (09h00) was
45.2 %, and at the exit of the tower, it was 60.7 %. The lowest value was recorded at the exit of the solar chimney, where it reached
27.3. In the lower level adjacent to the wind tower exit, no relative humidity difference was observed. A more significant variation in
relative humidity is observed at 14h00 near the wind tower exits and a less significant variation near the entrance of the solar chimney.
A slight variation in the value of relative humidity is observed in other interior spaces as we go deeper, as shown in Fig. 26. Its value
inside the cell varies between 46.7 % and 49.4 %. The relative humidity value was very high at 18h00 and its distribution occurred in
all points within the cell.

6.3. ACH evolution


The following table indicates the air exchange rate (ACH) and the airflow rate exiting the chimney.
Table 5 shows that the solar radiation value influences the air exchange rate (ACH). Indeed, an increase in solar radiation leads to a
rise in absorber temperature, which in turn increases the driving force and buoyancy force. An increase in wind velocity results in a
higher volumetric air flow rate within the cell, significantly improving the ACH. The required ACH value typically falls between 4 and
30 1/h, so the results obtained in this study are acceptable.

7. Conclusion
Research highlights that ventilation through solar chimneys and windbreaks can lead to an increase in internal temperature. To

Fig. 28. Relative humidity inside the cell: (a) At 09h00 a.m. and Vinl = 0.7 m/s; (b) at 14h00 and Vinl = 0.9 m/s; (c) at 18h00and Vinl = 0.4 m/s.

19
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

Table 5
Effect of radiation and air velocity on air exchange rate.

Time (s) Speed (m/s) Radiation (W/m2) Mass flow (kg/s) Volume flow (m3/s) ACH (1/h)

09h00 0.7 409 0.09 0.08 11.86


14h00 0.9 487 0.16 0.15 20.44
18h00 0.4 114 0.07 0.06 9.13

ensure favorable thermal conditions and air circulation throughout the room, a ventilation system consisting of a solar chimney and a
wind collector equipped with a water spray system must be installed, to improve the efficiency of ventilation and thermal control. The
primary goal of this research is to evaluate the effectiveness of the wind catcher system and improve its contribution to airflow dis­
tribution within a test cell subjected to actual weather conditions. To improve ventilation and cooling potential in hot and semi-arid
conditions, the system incorporates a solar chimney and a direct evaporative system. our study finds that an increase in wind velocity
leads to a proportional rise in air velocity within the wind tower. When air directly contacts the wall opposite the entrance, there’s a
substantial increase in velocity, reaching up to 227.51 % near the entrance and 228.57 % in the lower regions. This heightened velocity
results in a faster decrease in outgoing air temperature from the wind catcher to the test cell, significantly reducing the cooling load. It
notes a temperature reduction of up to 7 ◦ C with a wind velocity of 0.9 m/s when water is introduced. Furthermore, the humidity level
demonstrates fluctuations with outdoor temperature variations, showing a minimum at 9 a.m. and a maximum at 6 p.m. The tem­
perature disparity between the wind tower outlet and inlet showcases a significant difference of 8.6 ◦ C. This temperature variation
significantly impacts air movement within the test cell, demonstrating a notable decrease in temperature compared to ambient
conditions and a considerable rise in relative humidity. These temperature changes create noticeable fluctuations in relative humidity
values. This research confirms that the volume and inflow rate of outside air entering the cell also alters relative humidity values. This
suggests that the quantity and flow rate of external air play a role in influencing the humidity conditions within the test cell. The results
of this study demonstrate the significant impact of wind speed, evaporative cooling, and the interaction between the wind catcher and
the test cell on temperature reduction and humidity control. They provide valuable insights into optimizing airflow distribution and
cooling efficiency, especially in hot and semi-arid conditions, offering potential strategies to improve ventilation and thermal comfort
in such environments.
Furthermore, we can propose the following recommendations based on the results and conclusions of our study.
• Study the solar chimney combined with a wind catcher in multi-story buildings to analyze airflow distribution and cooling
efficiency.
• Conduct further studies on this system to evaluate its performance in other regions and under different climatic conditions.

Availability of data and materials


Not applicable.

CRediT authorship contribution statement


Nadia Saifi: Conceptualization, Formal analysis, Investigation, Methodology, Software, Writing – original draft. Aymen Baadi:
Funding acquisition, Methodology, Software, Data curation. Rebha Ghedamsi: Investigation, Writing – review & editing. Abdes­
samed Guerrout: Investigation, Methodology, Writing – review & editing. Noureddine Settou: Conceptualization, Supervision,
Writing – review & editing.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

References
[1] P.K. Sangdeh, N. Nasrollahi, Windcatchers and their applications in contemporary architecture, Energy and Built Environment 3 (2022) 56–72, https://doi.org/
10.1016/j.enbenv.2020.10.005.
[2] C.A. Varela-Boydo, S.L. Moya, R. Watkins, Analysis of traditional windcatchers and the effects produced by changing the size, shape, and position of the outlet
opening, J. Build. Eng. 33 (2021) 101828, https://doi.org/10.1016/j.jobe.2020.101828.
[3] L. Godoy-Vaca, M. Almaguer, J. Martínez, A. Lobato, M. Palme, Analysis of solar chimneys in different climate zones - case of social housing in Ecuador, IOP
Conf. Ser. Mater. Sci. Eng. 245 (2017) 072045, https://doi.org/10.1088/1757-899X/245/7/072045.
[4] S. Nateghi, M.H. Jahangir, Performance evaluation of solar chimneys in providing the thermal comfort range of the building using phase change materials,
J. Clean. Mater. 5 (2022) 100120, https://doi.org/10.1016/j.clema.2022.100120.
[5] E. Hadji, B. Ndiogou, S. Tigampo, A. Thiam, D. Azilinon, Optimization of a solar chimney with a horizontal absorber for building ventilation: a case study, Fluid
Dynam. Mater. Process. 19 (2022) 901–910, https://doi.org/10.32604/fdmp.2022.021973.

20
N. Saifi et al. Journal of Building Engineering 85 (2024) 108475

[6] A.M. Hassan, Solar chimney performance driven air ventilation promotion: an investigation of various configuration parameters, J. Buildings. 13 (2023) 2796,
https://doi.org/10.3390/buildings13112796.
[7] H. Zhang, Ventilation Performance of a Solar Chimney in Naturally Ventilated Multiple Chambers, Thesis (School of Engineering College of Science, Technology,
Engineering, and Maths), RMIT University, Australia, 2023.
[8] L. Shi, G. Zhang, W. Yang, D. Huang, X. Cheng, S. Setunge, Determining the influencing factors on the performance of solar chimney in buildings, J. Renew. Sust.
Energ. Rev. 88 (2018) 223–238, https://doi.org/10.1016/j.rser.2018.02.033.
[9] L.D.O. Neves, F. Marques Da Silva, Simulation and measurements of wind interference on a solar chimney performance, J. Wind Eng. Ind. Aerod. 179 (2018)
135–145, https://doi.org/10.1016/j.jweia.2018.05.020.
[10] A.A. Imran, J.M. Jalil, S.T. Ahmed, Induced flow for ventilation and cooling by a solar chimney, Renew. Energy 78 (2015) 236–244, https://doi.org/10.1016/j.
renene.2015.01.019.
[11] N. Saifi, N. Settou, B. Dokkar, B. Negrou, N. Chennouf, Experimental study and simulation of airflow in solar chimneys, Energy Proc. 18 (2012) 1289–1298,
https://doi.org/10.1016/j.egypro.2012.05.146.
[12] D.J. Harris, N. Helwig, Solar chimney and building ventilation, J. Appl Energy 84 (2007) 135–146, https://doi.org/10.1016/j.apenergy.2006.07.001.
[13] J. Hirunlabh, W. Kongduang, P. Namprakai, J. Khedari, Study of natural ventilation of houses by a metallic solar wall under tropical climate, J. Renewable
Energy 18 (1999) 109–119, https://doi.org/10.1016/S0960-1481(98)00783-6.
[14] M.M. AboulNaga, S.N. Abdrabboh, Improving night ventilation into low-rise buildings in hot-arid climates exploring a combined wall–roof solar chimney,
J. Renewable Energy 19 (2000) 47–54, https://doi.org/10.1016/S0960-1481(99)00014-2.
[15] K.S. Ong, C.C. Chow, Performance of a solar chimney, J. Sol Energy 74 (2003) 1–17, https://doi.org/10.1016/S0038-092X(03)00114-2.
[16] G. Gan, S.B. Riffat, A numerical study of solar chimney for natural ventilation of buildings with heat recovery,, J. Appl. Therm. Eng 18 (1998) 1171–1187,
https://doi.org/10.1016/S1359-4311(97)00117-8.
[17] K.H. Lee, R.K. Strand, Enhancement of natural ventilation in buildings using a thermal chimney, J. Build. Eng. 41 (2009) 615–621, https://doi.org/10.1016/j.
enbuild.2008.12.006.
[18] M. Bahadori, A. Dehghani-sanij, Wind Towers Architecture, Climate and Sustainability, Springer Cham Heidelberg, New York Dordrecht London, 2014, https://
doi.org/10.1007/978-3-319-05876-4.
[19] M.R. Fanood, The role of four key structures in the creation and survival of cultural landscapes in the desert environment of Iran,, J. Architect. Conserv. 20
(2014) 184–196, https://doi.org/10.1080/13556207.2014.985490.
[20] A.A. Elmualim, H.B. Awbi, Wind Tunnel, Cfd, Investigation of the performance of “windcatcher” ventilation systems, Int. J. Vent. 1 (2002) 53–64, https://doi.
org/10.1080/14733315.2002.11683622.
[21] S.A. Gage, J.M.R. Graham, Static split duct roof ventilators, Build. Res. Inf. 28 (2000) 234–244, https://doi.org/10.1080/09613210050073698.
[22] M.M. Maneshi, A. Rezaei-Bazkiaei, A.S. Weber, G.F. Dargush, A numerical investigation of impact of architectural and climatic parameters of windcatcher
systems on induced ventilation, in: International Mechanical Engineering Congress & Exposition, American Society of Mechanical Engineers Digital Collection,
2013, pp. 1103–1117, https://doi.org/10.1115/IMECE2012-87139. Houston, Texas, USA.
[23] N. Goudarzi, M. Sheikhshahrokhdehkordi, J. Khalesi, S. Hosseiniirani, Airflow and thermal comfort evaluation of a room with different outlet opening sizes and
elevations ventilated by a two-sided wind catcher, J. Build. Eng. 37 (2021) 102112, https://doi.org/10.1016/j.jobe.2020.102112.
[24] M. Farouk, Comparative study of hexagon & square windcatchers using CFD simulations, J. Build. Eng. 31 (2020) 101366, https://doi.org/10.1016/j.
jobe.2020.101366.
[25] M. Ghadiri, The effect of tower height in square plan wind catcher on its thermal behavior, Australian Journal of Basic and Applied Sciences 5 (2011) 381–385.
[26] A.A. Elmualim, Effect of damper and heat source on wind catcher natural ventilation performance, J. Energy Build 38 (2006) 939–948, https://doi.org/
10.1016/j.enbuild.2005.11.004.
[27] S. Attia, A. Herde, Designing the malqaf for summer cooling in low-rise housing, an experimental study, in: 26th Conference on Passive and Low Energy
Architecture, Canada, Quebec City, 2009, pp. 1–6.
[28] M.H. Ghadiri, N. Lukman, N. Ibrahim, M.F. Mohamed, Computational analysis of wind-driven natural ventilation in a two sided rectangular wind catcher, Int. J.
Vent. 12 (2013) 51–62, https://doi.org/10.1080/14733315.2013.11684002.
[29] M. Mahmoudi Zarandi, Analysis on Iranian wind catcher and its effect on natural ventilation as a solution towards sustainable architecture (Case Study: Yazd),
Eng. Technol. 3 (2009) 574–579, https://doi.org/10.5281/zenodo.1077649.
[30] A. Nugroho, Possibility to use solar chimney to improve stack ventilation in tropical climate, J. Alam Bina. 8 (2006). https://www.academia.edu/33432569/
Possibility_to_Use_Solar_Chimney_to_Improve_Stack_Ventilation_in_Tropical_Climate. (Accessed 1 September 2023).
[31] M.E. Mackay, The solar chimney and tower, in: M.E. Mackay (Ed.), Solar Energy: an Introduction, Oxford University Press, 2015, https://doi.org/10.1093/
acprof:oso/9780199652105.003.0007.
[32] J. Foroozesh, S.H. Hosseini, A.J. Ahmadian Hosseini, F. Parvaz, K. Elsayed, N. Uygur Babaoğlu, K. Hooman, G. Ahmadi, CFD modeling of the building integrated
with a novel design of a one-sided wind-catcher with water spray: focus on thermal comfort, J. Sustain. Energy Technol. Assess 53 (2022) 102736, https://doi.
org/10.1016/j.seta.2022.102736.
[33] H.F. Nouanégué, L.R. Alandji, E. Bilgen, Numerical study of solar-wind tower systems for ventilation of dwellings, J. Renewable Energy 33 (2008) 434–443,
https://doi.org/10.1016/j.renene.2007.03.001.
[34] L. Moosavi, M. Zandi, M. Bidi, E. Behroozizade, I. Kazemi, New design for solar chimney with integrated windcatcher for space cooling and ventilation, Build.
Environ. 181 (2020) 106785, https://doi.org/10.1016/j.buildenv.2020.106785.
[35] A. Abdullah, I.B. Said, D.R. Ossen, A sustainable bio-inspired cooling unit for hot arid regions: integrated evaporative cooling system in wind tower, J. Appl.
Therm. Eng. 161 (2019) 114201, https://doi.org/10.1016/j.applthermaleng.2019.114201.
[36] E. Erell, D. Pearlmutter, Y. Etzion, A multi-stage down-draft evaporative cool tower for semi-enclosed spaces: aerodynamic performance, J. Sol Energy 82 (2008)
420–429, https://doi.org/10.1016/j.solener.2007.10.010.
[37] M. Khan, I. Janajreh, Transevaporative cooling performance of a three-sided wind catcher, J. JJMIE 11 (2017) 225–233.
[38] A. Noroozi, Y.S. Veneris, Thermal assessment of a novel combine evaporative cooling wind catcher, J. Energies. 11 (2018) 442, https://doi.org/10.3390/
en11020442.
[39] M.E.H. Bourahli, Caractérisation D’un Composite Verre/Époxy, Doctoral Thesis, Ferhat Abbas University 1, 2014. http://dspace.univ-setif.dz:8888/jspui/
bitstream/123456789/1148/1/Th %c3 %a8se %20BOURAHLI.pdf.
[40] R. Mebrouk, Etude paramétrique des échanges convectifs turbulents dans les configurations D’intérêt pratique, Thèses de doctorat, Frères Mentouri Constantine
1 (2017). https://bucket.theses-algerie.com/files/repositories-dz/6198147040469986.pdf.
[41] S.H. Hosseini, E. Shokry, A.J. Ahmadian Hosseini, G. Ahmadi, J.K. Calautit, Evaluation of airflow and thermal comfort in buildings ventilated with wind
catchers: simulation of conditions in Yazd City, Iran, Energy for Sustainable Development 35 (2016) 7–24, https://doi.org/10.1016/j.esd.2016.09.005.
[42] A. Bohojło-Wiśniewska, Numerical modelling of humid air flow around A porous body, Acta Mech. Automatica 9 (2015), https://doi.org/10.1515/ama-2015-
0027.
[43] D.W. Bearg, Indoor Air Quality and HVAC Systems, first ed., CRC Press, New York, 1993. https://www.routledge.com/Indoor-Air-Quality-and-HVAC-Systems/
Bearg/p/book/9780873715744.

21

You might also like