0% found this document useful (0 votes)
2 views14 pages

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 GHz ISM band applications

This paper discusses the design and simulation of rectangular patch antenna arrays aimed at enhancing bandwidth for 2.4 GHz ISM band applications. The authors present various configurations of patch antenna arrays, achieving bandwidths of 290 MHz and 210 MHz for 2×2 and 2×4 arrays respectively, along with significant improvements in gain and impedance matching. The study utilizes Ansys HFSS for simulations, demonstrating the effectiveness of the quarter-wave transformer method in optimizing antenna performance.

Uploaded by

TELKOMNIKA
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
2 views14 pages

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 GHz ISM band applications

This paper discusses the design and simulation of rectangular patch antenna arrays aimed at enhancing bandwidth for 2.4 GHz ISM band applications. The authors present various configurations of patch antenna arrays, achieving bandwidths of 290 MHz and 210 MHz for 2×2 and 2×4 arrays respectively, along with significant improvements in gain and impedance matching. The study utilizes Ansys HFSS for simulations, demonstrating the effectiveness of the quarter-wave transformer method in optimizing antenna performance.

Uploaded by

TELKOMNIKA
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

TELKOMNIKA Telecommunication Computing Electronics and Control

Vol. 23, No. 3, June 2025, pp. 574~587


ISSN: 1693-6930, DOI: 10.12928/TELKOMNIKA.v23i3.26344  574

Design and simulation of rectangular patch antenna arrays with


high bandwidth for 2.4 GHz ISM band applications

Kymbat Kopbay, Madiyar Nurgaliyev, Ahmet Saymbetov, Nurzhigit Kuttybay, Askhat Bolatbek,
Sayat Orynbassar, Batyrbek Zholamanov
Department of Electronics and Astrophysics, Faculty of Physics and Technology, Al-Farabi Kazakh National University, Almaty,
Kazakhstan

Article Info ABSTRACT


Article history: Ongoing advancements in microstrip patch antenna (MPA) development
research is driven by its compact size, cost-effectiveness and ease of
Received Aug 31, 2024 fabrication. This paper presents a flexible design of patch antenna array
Revised Feb 6, 2025 (PAA) to address bandwidth (BW) limitations within the 2.4 GHz industrial,
Accepted Mar 11, 2025 scientific, and medical (ISM) band, where narrow BW is a common
challenge. To explore the effectiveness of different array configurations, we
designed and evaluated 1×2, 2×2, and 2×4 element rectangular PAA,
Keywords: employing a quarter-wave transformer (QWT) method and parallel scheme
for connecting patches. By utilizing Ansys high frequency structure
Ansys high frequency structure simulator (HFSS) as the modeling environment, we conducted extensive
simulator simulations to refine the antenna parameters and achieve the most optimal
Impedance matching MPA prototype. Our investigation demonstrates sufficiently good results,
Microstrip patch antenna array including BWs of 290 MHz and 210 MHz for 2×2 and 2×4 PAAs
Quarter-wave transformer respectively, which account for 8.75% and 12% of the total value. The
Wide bandwidth parameter return loss (RL) (S_11) reached -51dB for single-element patch
antenna (SPA) and -37.5 dB for 8-element PAA, that shows an ideal
impedance matching. In addition, the designed 2×4 PAA exhibited
impressive performance metrics, accounting for 9.17 dB in gain, 13 dBi in
directivity, and voltage standing wave ratio (VSWR) maintained below 0.5,
ensuring excellent signal transmission and reception.
This is an open access article under the CC BY-SA license.

Corresponding Author:
Madiyar Nurgaliyev
Department of Electronics and Astrophysics, Faculty of Physics and Technology
Al-Farabi Kazakh National University
71 Al-Farabi Ave, Almaty, Kazakhstan
Email: [email protected]

1. INTRODUCTION
In the fast-growing field of wireless communication systems, the need for efficient and compact
antenna designs remains as a paramount issue [1], [2]. Among these, microstrip patch antennas (MPAs) have
gained significant attention due to their compact dimension, easy fabrication, frequency-tuning flexibility,
and seamless integration into various electronic devices [3], [4]. With their planar structure and relatively low
profile, MPAs present an appealing solution for a broad range of applications, spanning from satellite
communications to wireless sensor networks (WSNs) [5]–[8]. Despite numerous advantages, MPA also has
some drawbacks such as narrow bandwidth (BW) and insufficient gain [9]. Additionally, there is a need to
enhance its directivity for specific applications [10].
The diverse application requirements have led to a variety of patch antenna designs [11]–[16], with
shapes ranging from rectangular, circular, and triangular to more complex geometries incorporating slots,

Journal homepage: http://journal.uad.ac.id/index.php/TELKOMNIKA


TELKOMNIKA Telecommun Comput El Control  575

stubs, and truncated grounds. The development of a PAA involves grappling with a series of more intricate
tasks compared to single-element patch antenna (SPA). Among these objectives, perhaps the most important
is to optimize the antenna performance while maintaining optimal geometry and spacing between individual
patch elements to achieve desired BW, directivity, and sidelobe levels. There are numerous works that were
carried out to construct antenna arrays of various sizes and shapes [17]–[21]. Prakasam et al. [22], simulated
and analyzed a 2×4 circular patch antenna array (PAA) with a center probe feed and FR-4 substrate. The
design resulted in a 6.46% BW, a return loss (RL) (S₁₁) -16.5 dB, and a gain 8 dB at 2.4 GHz. One of the
most critical aspects in probe feeding is to accurately place the pin rod connecting the patch with a microstrip
line, and even the slightest deviations can disrupt impedance matching and totally alter the resonant
frequency. However, in larger arrays, probe feeding facilitates uniform electromagnetic wave distribution
through the transmission network. The rectangular 1×2 PAA based sectored antenna for directional WSNs
was proposed and fabricated in [23] with a measured maximum gain of 5.2 dBi, corresponding to the
beamwidth of 45°. In order to connect individual patches, a corporate feeding technique (CFT), specifically a
quarter-wave transformer (QWT) was employed. In article [24], the QWT technique was utilized to serially
connect 2 and 4 element rectangular bow-tie patches, resulting in a 15.83% BW and a 13.6 dB gain for the 4-
element PAA configuration.
In rural areas, there is a great demand for PAA for wireless communication applications. Therefore,
Aji et al. [25] offered a 4×4 conventional rectangular PAA for remote regions using CFT with a probe and
fabricated it using FR-4 substrate. Kaboutari et al. [26] examined how the impedance matching and radiation
patterns (RPs) of arrays are affected by the quantity of elements in the array. They constructed an 8 ×2 square
antenna array with a common feed and showed that its RP exhibits side lobes consistent with the cosecant-
square curve. By employing a square shape and a daisy chain connection, they attained an average gain of
14.22 dBi and a wide impedance BW value of 1.93 GHz ranging from 9.97 to 11.90 GHz. Al-Yasir et al. [27]
presents a new type of antenna filter having a relatively 50% BW, designed based on three types of dielectric
substrates (Rogers RO3003, Rogers RT5880, and FR-4 epoxy), using the same design scheme at 2.4 GHz
band. Although the wide BW is obtained, it can be noticed that the best performance was reached by using
the Rogers RT5880, while antenna developed based on the FR-4 shows relatively little gain and matching.
This work aims to develop and analyze optimized, simplified design for a rectangular microstrip
PAA to achieve wider BW, higher gain and improved impedance matching using cost-effective materials as
FR-4 epoxy. We present the optimal design of the SPA model, as well as 1 ×2, 2×2, 2×4 PAA, where we
achieved remarkable results by increasing the BW to 290 MHz, as well as reaching RL up to -30 dB through
the high frequency structure simulator (HFSS) Ansys 3.0 modeling environment using FR-4 epoxy substrate
material. The operating frequency was set to 2.4 GHz due to its particular relevance in the realm of wireless
communication devices. This frequency band is globally allocated for industrial, scientific, and medical
(ISM) applications, making it an appealing option for wireless communication standards [28]. Opting for FR-
4 epoxy substrate instead of other alternatives as Rogers or RT/duroid is attributed to its cost-effectiveness,
affordability and some mechanical properties, which make it a good solution for a various electronic
applications.

2. METHOD
2.1. Single-element patch antenna geometry
2.1.1. Patch dimensions
The rectangular MPA stands out as the most optimal configuration among all available patch
geometries, making it a popular choice for widespread use. It includes a radiating patch printed on a substrate
made from dielectric material, backed by a ground plane [29]. The simplicity of its construction hides the
potential for high performance across a wide range of frequencies. However, proper impedance matching is
often a challenging task and every change in structure design considerably affects its basic characteristics,
such as RL, RP, and voltage standing wave ratio (VSWR), and gain [30]. Therefore, it is crucial to choose the
correct patch size, feed lines, ground surface, substrate thickness and placement of slots [31].
To develop the optimal MPA design, patch dimensions, substrate material and operating frequency
should be defined. The dimensions of the patch antenna is a key parameter that influences its radiation
diagram, impedance matching, resonant frequency, and so on. The equation that used for calculating width of
the patch is presented in (1) [32]:

𝑐 2
𝑊= √𝜀 (1)
2𝑓𝑟 𝑟 +1

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
576  ISSN: 1693-6930

where, 𝑐 represents the speed of light (3*108 ), 𝑓𝑟 is the antenna’s resonant frequency and 𝜀𝑟 denotes the
substrate’s dielectric permittivity. There are various substrates available for designing MPAs, such as FR-4,
Rogers RT/Duroid, and Teflon (usually 𝜀𝑟 takes value from 2.2 to 12). The choice of substrate relies on
several factors, including material cost, its thermal conductivity, dissipation coefficient, and others. In this
article, we selected FR-4 epoxy, with relative dielectric constant as 𝜀𝑟 =4.4 and substrate height of 1.6 mm.
In MPA, the patch width is typically chosen to be a fraction of the guided wavelength (𝜆𝑔 ) to
manage its resistance. The 𝜆𝑔 of a patch antenna denotes the wavelength of the electromagnetic wave
propagated through the microstrip transmission line formed by the patch and the ground plane. It is usually
shorter than the free-space wavelength (𝜆) due to the dielectric properties of the substrate such as effective
dielectric constant (𝜀𝑒𝑓𝑓 ) and the operating frequency. The 𝜆𝑔 can be written as (2):

𝜆 𝑐
𝜆𝑔 = = (2)
√𝜀𝑒𝑓𝑓 𝑓𝑟 √𝜀𝑒𝑓𝑓

The 𝜀𝑒𝑓𝑓 considers how the substrate properties and the geometry of the patch influence on the overall
electrical characteristics. It represents the average dielectric constant experienced by electromagnetic waves
within the microstrip patch structure and is nearly close value to 𝜀𝑟 , but slightly less. In order to calculate
𝜀𝑒𝑓𝑓 the following empirical equation is usually used:

𝜀𝑟 +1 𝜀𝑟 −1 12ℎ −1/2
𝜀𝑒𝑓𝑓 = + [1 + ] (3)
2 2 𝑊

where, 𝑊 is the width of the patch, and ℎ is the substrate material height, which is usually defined as ℎ << 𝜆
(0.003 𝜆≤ ℎ≤0.005). The patch length can be expressed by (4):

𝑐 𝜆𝑔
𝐿= = (4)
2𝑓𝑟 √𝜀𝑒𝑓𝑓 2

𝜆 𝜆
In the case of rectangular patch, the length (𝐿) of the element typically falls within the range of < 𝐿 < .
3 2
However, as radiation doesn’t emanate from the center but rather from the region between the edges
of the patch and the grounded surface, the phenomenon referred to as the fringing effect arises [33].
Consequently, electrically, the patch appears longer than its typically calculated length, giving rise to the
concept of ∆𝐿 (Figure 1), which represents the extra electrical length.

Figure 1. Physical and effective length of rectangular microstrip patch

A widely used and pragmatic approximate relationship for the normalized length elongation is:
𝑊
∆𝐿 (𝜀𝑒𝑓𝑓 +0.3)( +0.264)

= 0.412 𝑊 (5)
ℎ (𝜀𝑒𝑓𝑓 −0.258)( +0.8)

The patch effective length, denoted as 𝐿𝑒𝑓𝑓 , is given as (6):

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  577

𝐿𝑒𝑓𝑓 = 𝐿 + 2∆𝐿 (6)

Designing a suitable ground plane under the patch antenna will improve radiation efficiency and maintain
stable impedance. The dimensions of the grounding surface can be chosen by the subsequent expression:

𝐿𝑔 = 𝐿 + 6ℎ (7)

𝑊𝑔 = 𝑊 + 6ℎ (8)

2.1.2. Feeding calculation technique


Feeding mechanisms are used to transfer electromagnetic power from transmission line to the
antenna by transforming energy into electromagnetic radiation. It plays a critical role in impedance matching
and effects on basic characteristics such as RP, frequency and polarization. Common types of feeding
mechanisms in patch antennas include microstrip, proximity, aperture coupled and coplanar waveguide
configurations. Each of them has its own pros and cons, and the choice depends on factors such as antenna
design requirements, performance goals, and application constraints.
In this article, the microstrip edge feed was favored for its design simplicity and ease fabrication,
which involves connecting a transmission line straight to the edge of the patch antenna. With any feeding
technique, achieving accurate impedance matching is vital as it minimizes signal reflections and standing
waves, while maximizing the power transfer efficiency. By adjusting the width and length of the transmission
line, impedance matching can be optimized.
The width and length of the patch are calculated by (1) and (6), so that 𝑊𝑝 = 37.1 mm and
𝐿𝑝 = 27.7 mm, respectively. In order to define and calculate the feeding size, it is necessary to determine the
load impedance (Ƶ𝐿 ) of the patch through (9):

𝜀𝑟 2 𝐿
Ƶ𝐿 = 90 ∗ ( )2 (9)
𝜀𝑟 −1 𝑊

Thus, when substituting values into (9), Ƶ𝐿 ≈ 285 ohms. The calculation of the reflection coefficient
Г, provided that the characteristic impedance (Ƶ0 ) is 50 ohms, is carried out using (10):

Ƶ𝐿 −Ƶ0
Г= (10)
Ƶ𝐿 +Ƶ0

From (10), a value of Г=0.7 is determined, which indicates that 70% of the incident power at the
antenna’s feed point is reflected back towards the source, and the remaining 30% is either absorbed or
radiated by the antenna. The value of Г=0.7 indicates poor impedance matching, whereas ideal matching is
characterized by a reflection coefficient close to zero. To minimize the reflection coefficient, we employ
matching technique as QWT, which adjusts the antenna’s impedance to match the characteristic impedance
of the feeding transmission line. The phase and magnitude of the reflected signal is transformed to diminish
the reflection coefficient. To calculate QWT impedance (Ƶ𝑡 ) we will use (11):

Ƶ𝐿 +𝑗Ƶ𝑡 𝑡𝑎𝑛(𝛽𝑙)
Ƶ0 = Ƶ𝑡 (11)
Ƶ𝑡 +𝑗Ƶ𝐿 𝑡𝑎𝑛(𝛽𝑙

2𝜋 𝜆
Since in the matching network 𝑙 = 𝜆/4 (as shown in Figure 2), it implies that β𝑙 = . As a result, it is
𝜆 4
𝜋
evident that 𝑡𝑎𝑛( ) → ∞, leading to (11) being expressed as (12):
2

Ƶ2
𝑡
Ƶ0 = , Ƶ𝑡 = √ Ƶ0 Ƶ𝐿 (12)
Ƶ𝐿

Figure 2. QWT

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
578  ISSN: 1693-6930

𝜆
Hence, we can figure out the length of QWT by expression 𝐿𝑡 = .
4
The width of the QWT and microstrip feed with a 50 Ω characteristic impedance can be determined
using (13):

60 8ℎ 𝑊 𝑊
ln ( + ), <1
√ε𝑟 𝑊 4ℎ ℎ
Ƶ𝐶 = { 120𝜋 𝑊 (13)
𝑊 2 𝑊 , >1
√ε𝑟 (1.393+ ℎ +3 ln( ℎ +1.444)) ℎ

Figure 3 illustrates the geometric structure and key parameters of the rectangular SPA design,
including the dimensions of the patch, transformer, microstrip feed, and substrate. These labeled parameters
form the basis for the modeling and simulation discussed in this article. Table 1 provides the specific width
and length values corresponding to each parameter depicted in Figure 3. Since the design relies on empirical
equations, slight adjustments to these values may be necessary to achieve optimal performance in practical
implementations.

Figure 3. Rectangular single MPA

Table 1. The parameters of rectangular single patch antenna design


Symbols Details Value (mm)
Ws Substrate width 48
Ls Substrate length 66
Wp Patch width 37.1
Lp Patch length 27.7
Wt Width of transformer 1.2
Lt Length of transformer 15
W0 Width of microstrip line feed (50 Ohm) 2.98
L0 Length of microstrip line feed (50 Ohm) 15.5
H Height of substrate 1.6

2.2. Patch antenna array geometry


In specific scenarios, the desired traits of an antenna can be accomplished using just one microstrip
element. Nevertheless, similar to traditional microwave antennas, features like high gain, wide BW, and
beam direction control are only achievable by combining separate radiators into volumetric arrays. The
components may be arranged in series, parallel or by diagonal to form an array. In practical terms, the
selection of its configuration depends on the applications’ characteristics.
In this paper, we have gathered PAA of varying configurations, including 1×2, 2×2, 2×4 elements,
to ensure a more equitable comparison of the results. Each individual patch dimension coincides with the
information indicated in Table 1. In the 1×2 line array, the spacing between patches is determined based on
the guided wavelength 𝜆𝑔 , which is equal to 59.1 mm. A matching network with impedances Ƶ1 , Ƶ2 , Ƶ3 are
used to provide effective and smooth transition. This is essential because employing only a single line with a
fixed impedance value to connect two distinct patches often leads to significant power reflection, thereby
causing inadequate matching. To resolve this concern, we employ the QWT method relying on (12). Also, as

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  579

it can be seen from Figure 4(а), the wavelength between the patches is divided into fractions equal to λ/6, λ/4
and λ/2 (the section lengths used in the simulation may slightly differ from the theoretical data, in order to
reach the best matching). The width and length of matching network transformers are given in Table 2.

Table 2. The parameters of matching network feed


Impedances Width of feed (mm) Length of feed (mm)
Ƶ1 2.98 10.2
Ƶ2 1.54 7.6375
Ƶ3 1.2 23.425
Ƶ0 2.98 15.5

When constructing a 4-element PAA, the identical circuit is used as in a 2-element PAA; 2 similar
patches are added in parallel, which are located at a distance of 48.7 mm from the top ones (Figure 4(b)). To
enhance the performance characteristics of the antenna array, the distance between the upper and lower
antennas should be reduced to less than the wavelength, specifically 0.82 𝜆𝑔 in this case. The radiation
elements are energized through a common feedline, which guarantees an even distribution of power among
them. The same steps are performed when creating an 8-element PAA. In order to connect 2 sides of an array
consisting of 4 elements, we need to use conductor line with a length of 2 𝜆𝑔 (Figure 4 (c)).

(a) (b)

(c)

Figure 4. The front side of; (a) 1×2 PAA, (b) 2×2 PAA, and (c) 2×4 PAA

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
580  ISSN: 1693-6930

3. RESULTS AND DISCUSSION OF ANTENNA DESIGN SIMULATIONS


The simulation of proposed SPA and PAA was accomplished by utilizing the Ansys HFSS 3D
modeling software product. The operating frequency of the antenna is chosen to be 𝑓𝑟 =2.4 GHz, the substrate
material is FR-4 and dielectric constant 𝜀𝑟 =4.4. Following that, we compare the simulation results of 1,2,4, and
8-element patches, namely parameters as RL (𝑆11 ), BW, gain, VSWR, directivity, and 2D, 3D RPs.

3.1. Return loss (𝐒𝟏𝟏 ), bandwidth, and voltage standing wave ratio
In the field of antenna engineering the 𝑆11 value is a crucial metric as it quantifies how effectively
the antenna transforms incoming signal energy into radiated waves versus reflecting it back into the system.
Lower values of 𝑆11 denote superior impedance matching between the antenna and transmission line,
translating to reduced signal loss and enhanced overall performance.
As depicted in Figure 5(a), the 𝑆11 parameter of SPA reached the highest point as -51.7 dB at the
resonant frequency, indicating perfect matching without any signal reflections. Similarly, the 𝑆11 results do
not differ significantly in the 2-element PAA (Figure 5(b)), achieving -47.5 dB at 2.4 GHz. Both types of
antennas have a BW of 60 MHz, which is considered as a relatively narrow frequency band. Subsequently,
by utilizing a parallel patch addition scheme and optimizing patch placements based on the model, a 4-
element PAA (Figure 5(c)) attained a BW of 290 MHz at -10 dB, with an increased number of resonant
frequencies observed at 2.28 GHz (𝑆11 : -17.8 dB), 2.39 GHz (𝑆11 : -25.5 dB), and 2.49 GHz (𝑆11 : -30.6 dB).
The 8-element PAA exhibits multiple resonant frequencies, as shown in Figure 5(d). At 2.46 GHz, 𝑆11 drops
to -37.3 dB, while at 2.25 GHz, it reaches -21.6 dB. The BW covers range from 2.22 to 2.29 GHz and 2.36 to
2.5 GHz, giving a total BW of 210 MHz.

(a) (b)

(c) (d)

Figure 5. The 𝑆11 parameter and BW of; (a) SPA, (b) 1×2 PAA, (c) 2×2 PAA, and (d) 2×4 PAA

VSWR is directly related to RL, because they are inversely proportional. Higher 𝑆11 value
corresponds to lower VSWR, and vice versa. Lower VSWR indicates better impedance matching and less
power being reflected back towards the source. As it can be seen from Table 3, we obtained VSWR values
for SPA which is only 0.04, and also for 8-element PAA values of 0.23, indicating high signal transmission
efficiency.

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  581

Kaboutari et al. [26] constructed a 2×2 circular PAA using a probe feed, where connections between
individual patches are similar to our proposed approach. The distinction lies not only in the shape of the
patch, but also in the connection type, as we employ edge-feed supply instead of probe feed. Compared to the
circular PAA, which achieved only 3.65% of the BW in the frequency range from 2.37 GHz to 2.45 GHz, our
2×2 rectangular PAA increased the BW by 12% in the 2.21-2.53 GHz frequency range. Additionally, 𝑆11 ,
VSWR, and gain value are recorded at -15.19 dB, 1.4885, and 6.75 dBi, correspondingly. In comparison, the
proposed rectangular 2×2 PAA demonstrates results as -30.6 dB for 𝑆11 , 0.5 for VSWR, and 7.08 dBi for
gain. Aji et al. [25] presented a 4×4 rectangular PAA fed by a microstrip line, designed for wireless
applications in rural regions, demonstrating a BW of 130 MHz. Our study proposes a 2×4 PAA that exceeds
this result, reaching a BW of 210 MHz, constituting 8.5% of the total available BW.

Table 3. Various obtained model data of 1, 2, 4, and 8-element PAA


Parameter Number of patch elements
1 2 3 4
Resonant frequency (GHz) 2.4 2.4 2.49 2.46
RL (dB) -51.7 -47.5 -30.6 -37.3
Gain (dB) 2.64 6.03 7.08 9.17
BW (MHz) 60 60 290 210
VSWR 0.04 0.07 0.5 0.23
Directivity (dBi) 5.96 8.8 10.48 13.13
HPBW (°) (E plane) 83 [-37; 46] 89 [-40; 49] 52 [-5; 57] 61 [4; 65]

3.2. Radiation pattern, gain, and directivity


Antenna gain (G) is a crucial parameter that describes the concentration of the antenna’s RP. It
combines the directivity (D) and efficiency (η) of the antenna into a single, convenient value. Directivity
measures the ratio of power density (radiation intensity per unit area) of a real antenna in its main radiation
direction to that of a hypothetical isotropic antenna, which radiates evenly in all directions. In contrast, the
power density of an isotropic emitter is uniformly distributed across the surface of a sphere. A real antenna,
however, exhibits varying degrees of radiation directivity, which is quantitatively represented by the
directivity coefficient. The equation for calculating directivity is as (14):
𝑆
𝐷= (14)
𝑆𝑖

where, 𝑆 represents the power density of the real antenna in its primary radiation direction, and 𝑆𝑖 is the
isotropic antenna’s power density. While antenna directivity is a key aspect of antenna gain, real-world losses
must also be considered. The power radiated by the antenna depends on the power received from the
transmitter, which can be measured at the antenna feed. However, some power is lost along the supply line
due to its ohmic resistance, impacting the antenna’s efficiency. In the case of an ideal, lossless antenna, gain
equals directivity. For real practical antennas, the value of gain is always less than the directivity:

𝐺 = 𝜂𝐷 (15)

In order to see the full picture of the received data, we use the RP extracted through far-field
analysis in Ansys HFSS. We obtained 3D models of gain, encompassing azimuth angle (φ) values ranging
from -180° to 180° and elevation angle (θ) values ranging from 0° to 360° (Figure 6). The 3D RP of SPA
exhibits a uniformly round distribution, which suggests symmetrical radiation characteristics with minimal
variation in intensity around the antenna’s axis. The gain and directivity value of SPA reached 2.64 dB and
5.96 dBi, respectively. In the case of 1×2 PAA configuration, we obtained substantially strong gain of 6.03 dB,
along with directivity of 8.80 dBi. The presented three-dimensional RP in Figure 6(a) demonstrates that the SPA
exhibits a relatively broad main lobe, with a peak gain of approximately 2.64 dB, as indicated by the red region.
This pattern serves as a baseline for evaluating the improvements achieved with phased array configurations. As
shown in the RP of Figure 6(b), a noticeable level of radiation is still present in the back lobe, although the
overall gain has significantly increased. Specifically, the gain reaches 7.08 dB for the 2×2 phased antenna
array (PAA) and further improves to 9.17 dB for the 2×4 PAA, as illustrated in Figures 6(c) and 6(d). These
enhancements are accompanied by an increase in antenna directivity, reaching up to 13.13 dBi, indicating
improved beam focusing and overall performance.
Aji et al. [25] endeavored to enhance gain and directivity by using proximity feeding for 1×2
circular PAA, resulting in a 6.46 dBi gain and a directivity 7.59 dBi. In our study involving a 1×2 PAA,
although we attained a slightly lower gain of 6.03 dBi, the directivity was higher than the circular PAA,
Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
582  ISSN: 1693-6930

reaching 8.8 dBi. Balanis [32] created a sectored antenna with eight sides, each equipped with inset-feed 1×2
PAA’s. They achieved a gain 5.83 dBi, a RL -29.7 dB, and a beamwidth 45°, all while maintaining a λ/2
spacing between patches. In the proposed design of the 1×2 PAA, we accomplished a higher gain 6.03 dBi, a
RL -47 dB and a broader beamwidth due to the spacing between patches, which corresponds to λ.
Also, Kaboutari et al. [26] offers a 1×4 microstrip rectangular PAA with inset-fed feeding technique
for WiMax applications. As a result of employing the corporate feeding arrangement, the authors obtained a
maximum gain of 5.28 dBi and an 𝑆11 parameter of -20.37 dB. In contrast, the 2×2 PAA configuration
proposed in this paper exhibited higher gain and 𝑆11 values of 7.08 dBi and -30.6 dB, respectively.

(a) (b)

(c) (d)

Figure 6. The RP 3D model of; (a) 1×1, (b) 1×2 PAA, (c) 2×2 PAA, and (d) 2×4 PAA

Using a 3D polar plot, we can visualize how an antenna emits beams in 3D space, whereas a 2D plot
is useful for accurately defining directivity and angles of emission since it represents a cross-section of the
3D pattern. The 2D RPof the SPA is described by a half-power beamwidth (HPBW) of 83° (Figure 7(a)),
indicating the angular span within which the radiation intensity decreases to half (-3 dB) of the peak value. In
the 1×2 PAA’s scenario, the 2D plot demonstrates an even gain distribution and wide radiation angle similar
to SPA with HPBW value spanning 88° from -40° to 49° (Figure 7(b)). Additionally, a more pronounced
sidelobe and back lobe was observed from the H plane in the 2-element PAA. It can be noticed that the H
plane in each plot has a similar trend, where increasing the number of patches leads to a reduction in the
width of the main lobe and the appearance of additional sidelobes. For 2×2 and 2×4 element PAA

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  583

configurations, the E plane exhibits a more focused main lobe with a HPBW of 52° ranging from -5° to 57°
and 61° spanning from 4° to 65°, respectively (Figures 7(c) and (d)).
All the data acquired during modeling, including gain, RL, VSWR, directivity, and other
parameters, are showcased in Table 3. The animation of how electric field strength is propagated through the
media in 2×4 PAA is presented in Figure 8. As observed from simulation, the surface current is uniformly
propagated in two directions from the lower side of the patch antenna. However, there is a slight phase shift
between the upper and lower groups of patches.

(a) (b)

(c) (d)

Figure 7. The 2D RP of E-plane (φ=0°) and H-plane (φ=90°); (a) 1×1, (b) 1×2 PAA, (c) 2×2 PAA, and
(d) 2×4 PAA

Figure 8. The E field animation of 2×4 PAA in phase from 0° to 180° degree

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
584  ISSN: 1693-6930

4. CONCLUSION
This work presented a straightforward and effective design of PAA consisting of 2-, 4-, and
8-elements in a rectangular shape for 2.4 GHz ISM band. For this purpose, it was proposed to use a QWT
and parallel adding scheme with edge-feeding technique. As a result, there was a significant improvement in
overall BW by 290 MHz, which is considered as 12% of total BW. The simulation results illustrate that the
proposed architecture exhibited impressive values for RL (𝑆11 ) at -37.3 dB and VSWR at 0.23, indicating
excellent impedance matching. The maximum achievable gain and directivity were measured at -9.17 dB and
13.13 dB respectively for the 2×4 PAA configuration. By examining the RPs in 2D and 3D polar plots, it was
found that directivity enhances with an increase in the number of patch antenna elements. The proposed PAA
design offers seamless integration and is ideal for various wireless communication applications.

ACKNOWLEDGEMENTS
The authors would like to express their sincere gratitude to Al-Farabi Kazakh National University
for providing the resources and support necessary for this research. Special thanks are extended to the
Department of Electronics and Astrophysics for their valuable guidance and collaboration.

FUNDING INFORMATION
This research has been funded by the Science Committee of the Ministry of Science and Higher
Education of the Republic of Kazakhstan (Grant AP19678552).

AUTHOR CONTRIBUTIONS STATEMENT


This journal uses the Contributor Roles Taxonomy (CRediT) to recognize individual author
contributions, reduce authorship disputes, and facilitate collaboration.

Name of Author C M So Va Fo I R D O E Vi Su P Fu
Kymbat Kopbay ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓
Madiyar Nurgaliyev ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓
Ahmet Saymbetov ✓ ✓ ✓ ✓ ✓ ✓
Nurzhigit Kuttybay ✓ ✓ ✓ ✓
Askhat Bolatbek ✓ ✓ ✓
Sayat Orynbassar ✓ ✓ ✓ ✓ ✓ ✓
Batyrbek Zholamanov ✓ ✓ ✓ ✓

C : Conceptualization I : Investigation Vi : Visualization


M : Methodology R : Resources Su : Supervision
So : Software D : Data Curation P : Project administration
Va : Validation O : Writing - Original Draft Fu : Funding acquisition
Fo : Formal analysis E : Writing - Review & Editing

CONFLICT OF INTEREST STATEMENT


We confirm this manuscript has not been published elsewhere and is not under consideration by
another journal. The authors have no conflicts of interest to declare.

DATA AVAILABILITY
The authors confirm that the data supporting the findings of this study are available within the
article.

REFERENCES
[1] D. Piao and Y. Wang, “Tripolarized MIMO antenna using a compact single-layer microstrip patch,” IEEE Transactions on
Antennas and Propagation, vol. 67, no. 3, pp. 1937-1940, Mar. 2018, doi: 10.1109/TAP.2018.2889147.
[2] B. Mishra, R. K. Verma, N. Yashwanth, and R. K. Singh, “A review on microstrip patch antenna parameters of different geometry
and bandwidth enhancement techniques,” International Journal of Microwave and Wireless Technologies, vol. 14, no. 5, pp. 652-
673, Jun. 2022, doi: 10.1017/S1759078721001148.
[3] S. K. Ezzulddin, S. O. Hasan, and M. M. Ameen, “Microstrip patch antenna design, simulation and fabrication for 5G

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  585

applications.” Simulation Modelling Practice and Theory, vol. 116, p. 102497, Apr. 2022, doi:10.1016/j.simpat.2022.102497.
[4] K. F. Lee, and K. F. Tong, “Microstrip patch antennas—basic characteristics and some recent advances,” in Proceedings of the
IEEE, vol. 100, no.7, pp. 2169-2180, Jul.2012, doi: 10.1109/JPROC.2012.2183829.
[5] K. Hati, N. Sabbar, A. El. Hajjaji, and H. Asselman, “A novel multiband patch antenna array for satellite applications,” Procedia
Engineering, vol. 181, pp. 496-502, May 2017, doi: 10.1016/j.proeng.2017.02.422.
[6] A. S. Elkorany et al., “Implementation of a miniaturized planar tri-band microstrip patch antenna for wireless sensors in mobile
applications,” Sensors, vol. 22, no. 2, p. 667, Jan. 2022, doi: 10.3390/s22020667.
[7] D. B. Thi and T. H. T. Phuong, “A narrow beam steering antenna array for indoor positioning systems based on wireless sensor
network,” IEEE Access, vol. 10, pp. 2169–3536, Aug. 2022, doi: 10.1109/ACCESS.2022.3200594.
[8] C. Sun, “A design of compact ultrawideband circularly polarized microstrip patch antenna,” IEEE Transactions on Antennas and
Propagation, vol. 67, no. 9, pp. 6170-6175, Jun. 2019, doi: 10.1109/TAP.2019.2922759.
[9] S. Parasuraman, S. Yogeeswaran, and G. P. Ramesh, “Design of Microstrip Patch Antenna with improved characteristics and its
performance at 5.1 GHz for Wireless Applications,” in IOP Conference Series: Materials Science and Engineering: 1st ICCEMS,
vol. 925, no. 012005, Jul. 2020, doi: 10.1088/1757-899X/925/1/012005.
[10] M. Sağık et al., “Optimizing the gain and directivity of a microstrip antenna with metamaterial structures by using artificial neural
network approach,” Wireless Personal Communications vol. 118, pp. 109-124, Jan. 2021, doi: 10.1007/s11277-020-08004-8.
[11] S. S. Hao, Q. Q. Chen, J. Y. Li, and J. Xie, “A high-gain circularly polarized slotted patch antenna,” IEEE Antennas and Wireless
Propagation Letters, vol. 19, no. 6, pp. 1022-1026, June 2020, doi: 10.1109/LAWP.2020.2987330.
[12] A. Romputtal and C. Phongcharoenpanich, “T-Slot Antennas-Embedded ZigBee Wireless Sensor Network System for IoT-
Enabled Monitoring and Control Systems,” IEEE Internet of Things Journal, vol. 10, no. 23, pp. 20834-20845, Jun. 2023, doi:
10.1109/JIOT.2023.3284005.
[13] H. R. Chowdhury and S. Hussain, “A Compact Quasi-Yagi Microstrip Patch Antenna with High Gain and Bandwidth for UWB
Application,” International Journal of Electronics and Telecommunications, vol. 69. no. 3, pp. 431-437, doi:
10.24425/ijet.2023.144380.
[14] M. D. Fernandez, D. Herraiz, D. Herraiz, A. Alomainy, and A. Belenguer, “Design of a Wide-Bandwidth, High-Gain and Easy-to-
Manufacture 2.4 GHz Floating Patch Antenna Fed with the Through-Wire Technique,” Applied Sciences, vol. 12, no. 24, Dec.
2022: 12925, doi: 10.3390/app122412925.
[15] L. Y. Nie, B. K. Lau, S. Xiang, H. Aliakbari, B. Wang, and X. Q. Lin, “Wideband design of a compact monopole-like circular
patch antenna using modal analysis,” IEEE Antennas and Wireless Propagation Letters, vol. 20, no. 6, pp. 918-922, Jun. 2021,
doi: 10.1109/LAWP.2021.3066985.
[16] J. Anguera, A. Andújar, and J. Jayasinghe, “High-directivity microstrip patch antennas based on TM odd-0 modes,” IEEE
Antennas and Wireless Propagation Letters, vol. 19, no. 1, pp. 39-43, Jan 2020, doi: 10.1109/LAWP.2019.2952260.
[17] Y. Liu, Y. Jia, W. Zhang, Y. Wang, S. Gong, and G. Liao, “An integrated radiation and scattering performance design method of
low-RCS patch antenna array with different antenna elements,” IEEE Transactions on Antennas and Propagation, vol. 67, no. 9,
pp. 6199-6204, Sept. 2019, doi: 10.1109/TAP.2019.2925194.
[18] S. H. Abd Hamid, G. C. Hock, and N. Ferdous, “Optimization of Directivity and Gain Performances on Circular Patch Antenna
Design for 2.4 GHz Applications,” International Journal of Recent Technology and Engineering, vol. 8, no. 4, pp. 6442-6447,
Nov. 2019, doi: 10.35940/ijrte.D5155.118419
[19] N. Ab Wahab, S. A. Nordin, W. N. W. Muhamad, and S. S. Sarnin, “Microstrip rectangular inset-fed patch array antenna for
wimax application,” in 2020 IEEE International RF and Microwave Conference (RFM), IEEE, pp. 1-4, Dec. 2020, doi:
10.1109/RFM50841.2020.9344799.
[20] Z. Gan, and Z. H. Tu, “Dual-mode conjoint patch-pair for 5G wideband patch antenna array application,” IEEE Antennas and
Wireless Propagation Letters, vol. 20, no. 2, pp. 244-248, Feb. 2021, doi: 10.1109/LAWP.2020.3046759.
[21] P. Bouca, J. N. Matos, S. R. Cunha, and N. B. Carvalho, “Low-profile aperture-coupled patch antenna array for CubeSat
applications,” IEEE Access, vol. 8, pp. 20473-20479, Jan. 2020, doi: 10.1109/ACCESS.2020.2968060.
[22] V. Prakasam, K. R. A. LaxmiKanth, and P. Srinivasu, “Design and simulation of circular microstrip patch antenna with line feed
wireless communication application,” 2020 4th International Conference on Intelligent Computing and Control Systems (ICICCS)
IEEE, Jun. 2020, pp. 279-284, doi: 10.1109/ICICCS48265.2020.9121162.
[23] S. Nagaraju, L. Gudino, B. Kadam, V. Khairnar, J. X. Rodrigues, and C. K. Ramesha, “Rectangular microstrip patch antenna
array based sectored antenna for directional wireless sensor networks,” in 2020 12th International Symposium on Communication
Systems, Networks and Digital Signal Processing (CSNDSP) IEEE, Jul. 2020, pp. 1-6, doi:
10.1109/CSNDSP49049.2020.9249489.
[24] K. Mohammed, D. Mehdi, and S. Zeggai, “A 30 GHz Slotted Bow-Tie Rectangular Patch Antenna Design for 5G
Application,” International Journal of Electronics and Telecommunications, vol. 69, no. 4, pp. 669-673, 2023, doi:
10.24425/ijet.2023.147686.
[25] G. M. Aji, M. A. Wibisono, and A. Munir, “High gain 2.4 GHz patch antenna array for rural area application,” in 2016 22nd Asia-
Pacific Conference on Communications (APCC), IEEE, Aug. 2016, pp. 319-322, doi: 10.1109/APCC.2016.7581507
[26] K. Kaboutari, A. Zabihi, B. Virdee, and M. P. Salmasi, “Microstrip patch antenna array with cosecant-squared radiation pattern
profile,” AEU-International Journal of Electronics and Communications, vol. 106, pp. 82-88, July 2019, doi:
10.1016/j.aeue.2019.05.003.
[27] Y. I. Al-Yasir et al., “A new and compact wide-band microstrip filter-antenna design for 2.4 GHz ISM band and 4G
applications,” Electronics, vol. 9, no. 7, p. 1084, Jul. 2020, doi: 10.3390/electronics9071084.
[28] M. F. Zambak et al., “A Compact 2.4 GHz L-Shaped Microstrip Patch Antenna for ISM-Band Internet of Things (IoT)
Applications,” Electronics, vol. 12, no. 9, p. 2149, May 2023, doi: 10.3390/electronics12092149.
[29] S. G. Fang, S. W. Qu, S. Yang, X. Q. Li, H. B. Sun, and Z. P. Zhou, “Low scattering patch array antenna based on grooved
ground,” IEEE Antennas and Wireless Propagation Letters, vol. 20, no. 3, pp. 308-312, Mar. 2021, doi:
10.1109/LAWP.2020.3048779.
[30] M. Alibakhshikenari et al., “Impedance bandwidth improvement of a planar antenna based on metamaterial-inspired T-matching
network,” IEEE Access, vol. 9, pp. 67916-67927, May 2021, doi: 10.1109/ACCESS.2021.3076975.
[31] A. Bansal and R. Gupta, “A review on microstrip patch antenna and feeding techniques,” International Journal of Information
Technology, vol. 12 no. 1, pp. 149-154, Mar. 2020, doi: 10.1007/s41870-018-0121-4.
[32] C. A. Balanis, ”Antenna theory: analysis and design,” John Wiley & Sons, pp. 816, 2016.
[33] M. K. Aghwariya and A. Kumar, “Microstrip patch antenna techniques for wireless applications” in Microstrip Antenna Design
for Wireless Applications, 1st ed., pp. 3-12, CRC Press, 2021.

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)
586  ISSN: 1693-6930

BIOGRAPHIES OF AUTHORS

Kymbat Kopbay was born in 1996. She received the B.S., the M.S. degree in
radio engineering, electronics and telecommunications from Al-Farabi Kazakh National
University, Almaty, Kazakhstan, in 2017 and 2019, respectively. From 2021, she is a Ph.D.
student at the Al-Farabi Kazakh National University. Since 2022, she has been a junior
researcher at the Department of Solid-State Physics and Nonlinear Physics. Her research
interests include wave propagation, antenna modeling, and wireless sensor networks. She can
be contacted at email: [email protected].

Madiyar Nurgaliyev was born in 1994. He received the B.S., the M.S. degree
and the Ph.D. degrees in radio engineering, electronics and telecommunications from Al-
Farabi Kazakh National University, Almaty, Kazakhstan, in 2016, 2018 and in 2022,
respectively. From 2018 to 2022, he was a lecturer at the Department of Solid-State Physics
and Nonlinear Physics at the Al-Farabi Kazakh National University. Since 2022, he has been a
senior lecturer at the Department of Solid-State Physics and Nonlinear Physics. His research
interests include wireless sensor networks, intelligent control systems, and signal processing.
He can be contacted at email: [email protected].

Ahmet Saymbetov was born in 1982. He received Ph.D. degree in electronics


engineering from Physical-Technical Institute of the Academy of Sciences of Uzbekistan,
Tashkent, Uzbekistan, in 2011. From 2009 to 2012, he was a senior researcher at the Physical-
Technical Institute of the Academy of Sciences of the Republic of Uzbekistan. From 2013 to
2018, he was a senior lecturer with the Department of Solid State Physics and Nonlinear
Physics, Al-Farabi Kazakh National University, Almaty, Kazakhstan. Since 2018, he has been
an Associate Professor at the Department of Solid State Physics and Nonlinear Physics, Al-
Farabi Kazakh National University, Almaty, Kazakhstan. His research interests include
wireless sensor networks, intelligent control systems, and information technology. He can be
contacted at email: [email protected].

Nurzhigit Kuttybay was born in Tulkubas, Kazakhstan in 1993. He received the


B.S. in 2015, the M.S. degree in 2017 and the Ph.D. degrees in 2022 in radio engineering,
electronics and telecommunications from Al-Farabi Kazakh National University, Almaty,
Kazakhstan. Since 2017, he has been a senior lecturer at the Department of Solid-State Physics
and Nonlinear Physics at the Al-Farabi Kazakh National University. His research interests
include electronics, intelligent control systems, and radio engineering. He can be contacted at
email: [email protected].

Askhat Bolatbek was born in Almaty, Kazakhstan in 1998. He received the B.S.
in 2020, the M.S. degree in 2022 in radio engineering, electronics, and telecommunications
from the Almaty University of Power Engineering and Telecommunications named after
Gumarbek Daukeyev. Since 2023, he has been studying the Ph.D. degree in radio engineering,
electronics, and telecommunications at the Al-Farabi Kazakh National University. Since 2023,
he has been working as a lecturer at the Department of Solid-State Physics and Nonlinear
Physics at the Al-Farabi Kazakh National University. His research interests include intelligent
control systems, wireless sensor networks, electronics, and signal processing. He can be
contacted at email: [email protected].

TELKOMNIKA Telecommun Comput El Control, Vol. 23, No. 3, June 2025: 574-587
TELKOMNIKA Telecommun Comput El Control  587

Sayat Orynbassar was born in 2000. He received the B.S. in radio engineering,
electronics, and telecommunications from the Al-Farabi Kazakh National University in 2022.
Since 2022, he is a master’s student in electronics and control systems at the Al-Farabi Kazakh
National University. His research interests include electronics, intelligent control systems, and
information technology. He can be contacted at email: [email protected].

Batyrbek Zholamanov was born in Aral, Kyzylorda, Kazakhstan in 1999. He


received the B.S. in 2020, the M.S. degree in 2022 in radio engineering, electronics, and
telecommunications from the Al-Farabi Kazakh National University Since 2022, he has been
studying the Ph.D. degree in radio engineering, electronics, and telecommunications at the Al-
Farabi Kazakh National University. He was a research assistant at the Department of Solid
State Physics and Nonlinear Physics at the Al-Farabi Kazakh National University. Since 2022,
he has been working as a lecturer at the Department of Solid State Physics and Nonlinear
Physics at the Al-Farabi Kazakh National University. His research interests include intelligent
control systems, wireless sensor networks, electronics, and data processing. He can be
contacted at email: [email protected].

Design and simulation of rectangular patch antenna arrays with high bandwidth for 2.4 … (Kymbat Kopbay)

You might also like